You are on page 1of 319

MESOZOIC RIFTING AND

CENOZOIC BASIN INVERSION


HISTORY OF THE
EASTERN CORDILLERA,
COLOMBIAN ANDES
Inferences from tectonic models
VRIJE UNIVERSITEIT

MESOZOIC RIFTING AND


CENOZOIC BASIN INVERSION
HISTORY OF THE
EASTERN CORDILLERA,
COLOMBIAN ANDES
Inferences from tectonic models

ACADEMISCH PROEFSCHRIFT

ter verkrijging van de graad van doctor aan


de Vrije Universiteit Amsterdam,
op gezag van de rector magnificus
Prof.dr. T. Sminia,
in het openbaar te verdedigen
ten overstaan van de promotiecommissie
van de faculteit der Aard- en Levenswetenschappen
op dinsdag 19 maart 2002 om 13.45 uur
in het hoofdgebouw van de universiteit,
De Boelelaan 1105

door

Luis Fernando Sarmiento Rojas

geboren te Santaf de Bogot


Colombia
promotor: prof.dr. S.A.P.L. Cloetingh
copromotor: dr. F. Roure
Luis Fernando Sarmiento, 2001
Copyright
Ecopetrol 2001
Vrije Universiteit Amsterdam

The research reported in this thesis was carried out at


the Tectonics Department
Faculty of Earth and Life Sciences, Vrije Universiteit Amsterdam
De Boelelaan 1085
1081 HV Amsterdam
The Netherlands

Netherlands Research School of Sedimentary Geology (NSG)


Publication number: 2002.01.01

Financial support was provided by ECOPETROL Empresa Colombiana de Petrleos


And the ECOPETROL-ICETEX fund.

ISBN 9287-25-5
Si un hombre nunca se contradice,
ser porque nunca dice nada
Miguel de Unamuno.

(If a man never contradicts


himself, it will be because
he never says anything)

To my wife Gladytas
My mother, my sons Cesar and Daniel
and to the memory of my father Filiberto
CONTENTS

ACKNOWLEDGEMENTS xii

SUMMARY xiv

SAMENVATTING xx

RESUMEN xxvi

CHAPTER 1: INTRODUCTION AND OUTLINE 1

1.1 AIM AND METHODS 1


1.2 BASIN FORMATION AND INVERSION MODELS 1
1.3 THE EASTERN CORDILLERA OF COLOMBIA 2
1.4 TECTONIC SETTING OF THE EASTERN CORDILLERA 2
1.5 THESIS OUTLINE 3

CHAPTER 2: MESOZOIC RIFTING HISTORY OF THE EASTERN


CORDILLERA, COLOMBIAN ANDES 7

1. INTRODUCTION 7
2. TECTONIC SETTING 8
2.1 PLATE-TECTONIC INTERPRETATIONS 8
2.1.1 Triassic and Jurassic 8
2.1.2 Cretaceous 12
3. STRATIGRAPHY 12
3.1 TRIASSIC AND JURASSIC SYN-RIFT SEDIMENTATION 12
3.2 CRETACEOUS SEDIMENTATION 18
3.2.1 Early Cretaceous Syn-Rift Sedimentation 18
3.2.2 Cretaceous Post-Rift Sedimentation 26
4. SUBSIDENCE ANALYSIS 31
4.1 TECTONIC SUBSIDENCE DURING TRIASSIC AND JURASSIC TIME 34
4.1.1 Basin Compartments 37
4.1.2 Triassic and Jurassic Subsidence Events 37
4.2 TECTONIC SUBSIDENCE DURING CRETACEOUS TIME 42
4.2.1 Basin Compartments 42
4.2.2 Cretaceous Fast Subsidence Events 44
5. CORRELATION OF FAST SUBSIDENCE EVENTS WITH MAGMATIC, EUSTATIC AND
PLATE-TECTONIC EVENTS 48
5.1 CORRELATION BETWEEN FAST SUBSIDENCE EVENTS AND SUBDUCTION
RELATED MAGMATIC ARCS 48
5.2 CORRELATION BETWEEN FAST SUBSIDENCE EVENTS, PLATE-TECTONIC EVENTS
AND EUSTATIC EVENTS 52
6. FORWARD MODELLING OF BASIN EVOLUTION 54
6.1 NUMERICAL MODEL 54
6.2 MODELLING PROCEDURE 55
6.3 MODEL RESULTS: STRETCHING FACTORS 56

vii
6.3.1 Mesozoic Lithosphere Stretching Phases 56
6.3.2 Total Stretching 61
7. DISCUSSION 62
7.1 GEOMETRY OF RIFT BASINS 62
7.2 RELATIONSHIPS BETWEEN MESOZOIC RIFTING AND MAGMATISM 66
7.3 SUGGESTIONS FOR FUTURE STUDIES 70
8. CONCLUSIONS 70

CHAPTER 3: PALAEOGENE INCIPIENT BASIN INVERSION HISTORY OF


THE EASTERN CORDILLERA, COLOMBIAN ANDES 72

1. INTRODUCTION 72
2. TECTONIC SETTING 74
2.1 PLATE TECTONIC INTERPRETIONS 74
2.1.1 Latest Cretaceous and Tertiary 74
3. STRATIGRAPHY 76
3.1 LATE MAASTRICHTIAN-EARLY PALEOCENE 76
3.2 LATE PALEOCENE 78
3.3 PALAEOGENE (EOCENE TO EARLY MIOCENE) 80
3.3.1 Llanos and Eastern Cordillera 81
3.3.2 Magdalena Valley (MV) 84
4. SUBSIDENCE ANALYSIS OF THE PALAEOGENE SEDIMENTARY RECORD 93
4.1. RESULTS 93
4.1.1 Events of Tectonic Subsidence 94
4.1.2 Maps of Tectonic Subsidence 96
5. MODELLING OF REMAINING THERMAL SUBSIDENCE AFTER MESOZOIC RIFTING 97
5.1 RESULTS 98
6. FLEXURAL BEHAVIOUR OF THE LITHOSPHERE 99
6.1 MODEL DESCRIPTION 101
7. MODELLING OF FLEXURAL SUBSIDENCE PRODUCED BY THE TOPOGRAPHIC LOAD
OF THE PALAEO-CENTRAL CORDILLERA DURING PALAEOGENE TIME 103
7.1 2D FLEXURAL MODELLING 103
7.2 3D FLEXURAL MODELLING 105
7.3 RESULTS 105
7.3.1 2D Models 105
7.3.2 3D Models 107
8. MODELLING OF FLEXURAL SUBSIDENCE PRODUCED BY THE TOPOGRAPHIC LOAD
OF THE PARTIALLY INVERTED EXTENSIONAL BASIN DURING PALAEOGENE TIME 108
8.1 RESULTS 109
9. MODELLING OF SUBSIDENCE PRODUCED BY THE COMBINED EFFECT OF (A)
REMAINING THERMAL SUBSIDENCE AFTER MESOZOIC RIFTING, (B) FLEXURAL
SUBSIDENCE PRODUCED BY THE TOPOGRAPHIC LOAD OF THE PALAEO-CENTRAL
CORDILLERA AND (C) FLEXURAL SUBSIDENCE PRODUCED BY LOCAL TOPOGRAPHY
DUE TO PARTIAL INVERSION OF THE MESOZOIC EXTENSIONAL BASIN 109
9.1 2D MODELS 109
9.1.1 Results 111
9.2 3D MODEL 117
10. COMPARISON OF MODEL RESULTS WITH FISSION TRACK DATA AND OTHER
EVIDENCE OF UPLIFT AND DEFORMATION 119
10.1 FISSION TRACK DATA AND OTHER EVIDENCE OF EXHUMATION AND
DEFORMATION 119
10.2. COMPARISON WITH MODEL RESULTS 124
11. STRUCTURAL KINEMATIC MODEL OF PALAEOGENE DEFORMATION, INCLUDING
SEDIMENTATION AND EROSION EVENTS 124
11.1 STRUCTURAL, SEDIMENTARY AND EROSION MODEL 124
11.2 RESULTS 125
12. DISCUSION 128
12.1 CONSTRAINING DIFFERENT TECTONIC SCENARIOS FOR THE PALAEOGENE 128

viii
12.2 PALAEOGENE TECTONIC HISTORY AND PLATE-TECTONICS 129
12.3 MECHANICAL ASPECTS OF BASIN INVERSION 130
12.3.1 Stresses 130
12.3.2 Rheology of the Lithosphere 130
12.3.3 Pre-Existing Crustal Discontinuities 132
12.4 LARGE-SCALE BASIN WIDE FLEXURE AND STRUCTURAL GEOLOGY 132
12.5 SUGGESTIONS FOR FUTURE STUDIES 133
13. CONCLUSIONS 133

CHAPTER 4: NEOGENE BASIN INVERSION HISTORY OF THE EASTERN CORDILLERA,


COLOMBIAN ANDES 135

1. INTRODUCTION 135
2. TECTONIC SETTING 135
2.1 PLATE TECTONIC INTERPRETATIONS 135
2.1.1 Middle Miocene 135
2.1.2 Late MiocenePliocene 135
3. STRATIGRAPHY 137
3.1 LLANOS ORIENTALES (LLA) 137
3.2 MAGDALENA VALLEY (MV) 140
3.2.1 Middle Magdalena Valley 142
3.2.2 Upper Magdalena Valley 142
3.3 SABANA DE BOGOT 145
4. TECTONIC SUBSIDENCE DURING THE NEOGENE 145
4.1 RESULTS 146
4.1.1 Llanos Orientales (LLA) 146
4.1.2 Magdalena Valley (MV) 146
5. NEOGENE REMAINING THERMAL SUBSIDENCE AFTER MESOZOIC RIFTING 146
5.1 RESULTS 147
6. NEOGENE FLEXURAL SUBSIDENCE 148
6.1. FLEXURAL MODELS PRODUCED BY THE PRESENT-DAY TOPOGRAPHIC LOAD OF
THE EASTERN AND CENTRAL CORDILLERA 149
6.1.1 Methods 149
6.1.2 Results 150
6.2 FLEXURAL MODELS PRODUCED BY THE PRESENT-DAY TOPOGRAPHIC LOAD OF
THE EASTERN CORDILLERA ONLY 151
6.2.1 Flexural Deflection of the Lithosphere as a Mechanism of Tectonic Subsidence During
Neogene Time 151
6.2.2 Calculated Bending Stress 155
6.3 FLEXURAL SUBSIDENCE MODELS WITH A BROKEN PLATE UNDER THE EASTERN
CORDILLERA 155
6.3.1 Methods 155
6.3.2 Results 156
6.4 LATERAL DISTRIBUTION OF EFFECTIVE ELASTIC THICKNESS AT PRESENT 157
7. FLEXURAL SUBSIDENCE PRODUCED BY GRADUAL SURFACE-UPLIFT OF THE EASTERN
CORDILLERA DURING NEOGENE TIME 158
7.1 DATA 159
7.2 METHOD 160
7.3 RESULTS 161
8. UPLIFT EVOLUTION FROM FLEXURAL MODELING, FISSION TRACK AND GEOLOGICAL
DATA 162
8.1 SURFACE-UPLIFT EVOLUTION OF THE EASTERN CORDILLERA AS INFERRED FROM
FLEXURAL MODELLING 162
8.2 COMPARISON OF THE UPLIFT AND EXHUMATION EVOLUTION INFERRED FROM
FISSION TRACK AND GEOLOGICAL DATA 167
9. DISCUSSION 171
9.1 NEOGENE TECTONIC HISTORY AND PLATE-TECTONICS 171
9.2 LITHOSPHERE STRENGTH EVOLUTION AS INFERRED FROM FLEXURAL
MODELLING 172

ix
9.2.1 Effect of Lithosphere Thermal Age on Strength 174
9.2.2 Coupling/Decoupling State of the Crust-Mantle, Thickness and Proportions of
Mechanically Competent Crust and Mantle 175
9.2.3 Effect of Local Curvature of the Plate on Lithosphere Strength 177
9.2.4 Effect of Pre-Existing Discontinuities on Lithosphere Strength 178
9.2.5 Other Effects on Lithosphere Strength 178
9.2.6 Temporal Changes of Lithosphere Strength 178
9.2.7 Suggestions for Further Studies 179
10. CONCLUSIONS 179

CHAPTER 5: MAP VIEW RESTAURATION OF TRANSPRESSIONAL


BASIN INVERSION IN THE EASTERN CORDILLERA 181

1. INTRODUCTION 181
2. TECTONIC SETTING 181
2.1 PRESENT DAY PLATE-TECTONIC SETTING 181
3. STRUCTURE OF THE EASTERN CORDILLERA 184
4. REGIONAL MAP VIEW RESTORATION OF THE NW CORNER OF SOUTH AMERICA 190
4.1 METHOD 190
4.2 RESULTS 191
5. MAP VIEW RESTORATION OF THE EASTERN CORDILLERA 192
5.1 METHOD 192
5.2 RESULTS 197
5.3 DISPLACEMENT OF BLOCKS RELATIVE TO STABLE SOUTH AMERICA AND
ROTATIONS ABOUT VERTICAL AXES 201
6. DISCUSION 201
6.1 ADVANTAGES OF THE MAP VIEW RESTORATION 201
6.2 COMPARISON WITH PREVIOUS SHORTENING ESTIMATES 201
6.3 COMPARISON WITH OUTCROP STRUCTURAL STUDIES, STRESS INFERRED FORM
BOREHOLE BREAKOUT DATA AND PLATE MOTIONS FROM GEOPHYSICAL DATA 203
6.3.1 Shortening Perpendicular to the Regional Structural Grain of the EC 203
6.3.2 Conjugate (?) Left-lateral and right-lateral strike-slip faults 204
6.3.3 Clockwise rotation of the Central Cordillera, Magdalena Valley and Western Flank of the
Eastern Cordillera 211
6.3.4 Deformation of the Eastern Cordillera by Transpressive Basin Inversion 212
6.3.5 Minor Structures Recognized in Outcrops, but not in the Map View Restoration 213
6.4 COMPARISON OF DISPLACEMENT OF BLOCKS RELATIVE TO STABLE SOUTH
AMERICA WITH PLATE VELOCITY VECTORS FROM GEOPHYSICAL DATA 213
6.5 ANDEAN NEOGENE DEFORMATION, BASIN INVERSION OF THE EASTERN
CORDILLERA AND LITHOSPHERE RHEOLOGY 215
7. CONCLUSIONS 215

CHAPTER 6: RHEOLOGICAL EVOLUTION OF THE LITHOSPHERE OF THE


EASTERN CORDILLERA AND HYPOTHESES ABOUT ITS deep STRUCTURE 217

1. INTRODUCTION 217
2. GEOPHYSICAL DATA CONSTRAINING THE DEEP STRUCTURE OF THE EASTERN
CORDILLERA 217
2.1 SEISMIC VELOCITY MODEL BASED ON REFRACTION IN THE SOUTH WEST OF
COLOMBIA 217
2.2 GRAVITY AND MOHO DISCONTINUITY DEPTH 217
2.3 SEISMICITY 220
2.4 GEOTHERMAL REGIME 227

x
3. RHEOLOGICAL EVOLUTION OF THE LITHOSPHERE OF THE EASTERN CORDILLERA 227
3.1 RHEOLOGY OF THE CONTINENTAL LITHOSPHERE 227
3.1.1 Rheology of the Continental Lithosphere from Rock Mechanics Data 227
3.1.2 Rheological Profiles and Integrated Strength of a Stratified Lithosphere 231
3.1.3 Thermal Structure of Continental Lithosphere 232
3.2 RHEOLOGY MODELS OF THE EASTERN CORDILLERA THROUGH TIME 233
3.2.1 Rheology Models 233
3.2.2 Local Isostasy vs Regional Isostasy Effects 235
3.2.3 Model Results 235
3.2.4 Comparison of Lithosphere Strength Estimates form Rheological Models and Flexural
Models 235
3.2.5 Models of Evolution of Lithosphere Rheology and Basin Extension and Inversion in the
EC 241
4. EVOLUTION OF STRESSES AFFECTING THE LITHOSPHERE OF THE EASTERN
CORDILLERA 242
4.1 MESOZOIC 242
4.2 CENOZOIC 242
5. MESOZOIC-CENOZOIC TECTONIC EVOLUTION OF THE EASTERN CORDILLERA 243
5.1 MESOZOIC EXTENSIONAL BASIN FORMATION 243
5.2 CENOZOIC TRANSPRESSIONAL INVERSION OF MESOZOIC EXTENSIONAL BASINS 243
6. COMPARISON OF THE EASTERN CORDILLERA WITH SIMILAR MONTAIN BELTS,
ANALOGUE AND NUMERICAL MODELLING EXPERIMENTS AND HYPOTHESES ABOUT
THE DEEP STRUCTURE OF THE EASTERN CORDILLERA 244
6.1 ASYMMETRY OF THE EASTERN CORDILLERA 244
6.2 COMPARISON AND SIMILARITIES WITH THE PYRENEES AND THE MRIDA ANDES
OF VENEZUELA 245
6.3 FLEXURAL MODELLING OF BROKEN PLATE AND SUBDUCTION OF THE MANTLE
LITHOSPHERE 248
6.4 COMPARISON WITH ANALOGUE MODEL EXPERIMENTS 248
6.5 COMPARISON WITH NUMERICAL MODELS OF BEAUMONT ET AL. (2000) AND ELLIS
AND BEAUMONT (1999) 250
6.5.1 Model Features 250
6.5.2 Results 250
6.6 NEOGENE VOLCANISM AND DEEP-INTERMEDIATE SEISMICITY: TWO ENIGMATIC
FEATURES OF THE EASTERN CORDILLERA 257
6.6.1 Neogene Magmatism of Paipa Iza 257
6.6.2 The Bucaramanga Earthquake Nest 257
6.6.3 Similarities Between the Bucaramanga Earthquake Nest and the Vrancea Seismic Cluster 258
6.6.4 The Slab Break-off Model 259
6.6.5 Need of Deep Seismic Refraction and Reflection Data 261
7. CONCLUSIONS 261

REFERENCES 263

xi
ACKNOWLEDGEMENTS

The completion of this Ph.D. thesis would have been impossible without the help and cooperation of
various people, organizations and institutions. It is a pleasure to acknowledge all people who made
possible this thesis, with apologies to anyone I may inadvertently omitted. In the first place I wish to
thank my promotor Sierd Cloetingh, for inviting me to do a Ph.D. at the Vrije Universiteit and for
helping me in scientific, economical and even personal matters. Also in the first place I wish to thank
my copromotor Franois Roure who encouraged me with the Ph.D. project and offered me his
hospitality during three weeks of my work at the IFP. I wish to thank his guidance and his help in many
aspects of the project. Jan Diederik van Wees introduced me with the topic of lithospheric stretching
and rheological models, he facilitated me the use of the program Whizmod developed by him. He also
ran the rheological models discussed in last chapter. Reini Zoetemeijer introduced me with flexural
modelling and facilitated me the use of the program Cobra developed by her. Dick Nieuwland
introduced me to the analogue modelling topic, made it possible for me to attend to the Thrust
Tectonics Conference in London and shared his knowledge and experience on structural geology.
Furthermore, I wish to thank all of them for their guidance, stimulating discussions, and constructive
criticism on the manuscript. Special thanks for detailed reviews of earlier versions of the manuscript
are also due to Harry Doust, Fernando Etayo-Serna, Tomas Villamil, Pedro Restrepo-Pace, and Roberto
Linares.

I wish to thank my former boss at ICP Kurt Bayer, and Jaime Cadavid Calvo former Director of the
Instituto Colombiano del Petrleo (ICP) who approved me a study commission to complete a Ph.D; and
Alirio Hernandez Director of the ICP who helped me to get approval for the final expenses. Without
the economic support from Ecopetrol and the Ecopetrol-Icetex Fund the completion of this thesis
would have been impossible. Egon Castro put special interest in creating possibilities for me to do a
Ph.D. and helped me to gather data from Ecopetrol. My former boss at ICP Bernardo Silva, my present
boss Yolanda Aguiar, Maria Ximena Mantilla and Alberto Ortiz made possible for me to get additional
support for the final corrections, edition and printing of this thesis. Oscar Daz, Juan Alvaro Gonzalez
and Olga Lucia Hernndez Pineda helped me with the final edition. Many other friends at Ecopetrol
helped me to gather high quality data sets or in many other different ways. Among them Jaime Muz,
John Ceron, Cesar Mora, Fabio Crdoba, Shajid Kairuz, Ariel Solano, Juan Pablo Reyes, Yolanda
Aguiar, Hans Bartels, Myriam Caro, Antonio Rangel, Blanca Nubia Giraldo, Diego Garca, Felix
Texeira, Ivan Olaya and his wife Martha, Oscar Daz, Alberto Ortz, Andrs Fajardo, Andrs Reyes,
Martn Mantilla, Fernando Munar and many others. Andrs Fajardo supplied me a software for the final
correction of the figures. Martn Mantilla supplied me with some seismic lines from the Magdalena
Valley. I wish to thank specially to my friends Jorge Rubiano and his wife Myriam for their continuous
help in many personal matters in Colombia and for their friendship. Gladys Rocio Ramrez and Bertha
Nereida Gmez and many others employees of Ecopetrol, ICP and Icetex made my Ph.D. possible.

I acknowledge the Netherlands Research School of Sedimentary Geology and the Vrije Universiteit of
Amsterdam for their kind invitation to do a Ph.D. research. I wish to thank in special Anco Lankreijer,
for his friendship and kind support, solving me administrative or financial difficulties. I am grateful to
all my colleagues in Amsterdam who provided a fruitful and nice working atmosphere and were always
ready to help me. Especially my roommates Bernd Andeweg and Ernst Willingshofer solved me many
doubts on various matters. Additionally Bernd translated the Dutch summary. I also appreciate the
support of my friends Jorge, Eduardo, Sandra and Anton. The support of Fred Beeckman and F.
Canemeijer on computer troubles is highly appreciated. Randell Stephenson is thanked for his fruitful
comments. Furthermore, I want to thank Aline, Ingrid, Jolante, Jos, Marlies, Sevgi, Stphanie, Cees,
Daniel, Captain Dick, Gabor, Gerco, Giovani, Harm, Harry, Henk, Joaquim, Markus, Mathias,
Christophe, Ritske, Rudie, Taco and Tore for their friendship and help. Margot Saher, Anouk Creusen
and Arjan van Vliet helped me with the figures of the last chapter. They were also my roommates
during my last months in Amsterdam along with Gideon. I also acknowledge their pleasant company.
Thanks also to Alwien Prinsen, Monique Gerbrands, Ellen Salmom, Liesbeth Aardema, Hetty Turley

xii
and Marjo Duifs who also solved me many practical problems. Sjaak Heezen helped me a lot on
logistic matters in Amsterdam before my arrival here.

I am indebted with the Institute Franais du Petrole (IFP), especially with Franois Roure who made it
possible for me to work there, use their facilities, and participate in a Subtrap meeting at Avignon.
Thanks also to William Sassi who helped me using the program Thrustpack and Bernard Colletta for
his interest and many suggestions about the structural geology of the Colombian Eastern Cordillera. I
like to express my gratitude to Jaime Toro, Nathalie Bordeaux and the Iraqi. All of them made my stay
in Paris very pleasant. Thanks also to the employees of the IFP who helped me in financial or
administrative matters.

I deeply appreciate the unconditional moral support and friendship from Jos Estevez and his wife
Lucy. Jos Estevez also helped me with some figures. This thesis would have been impossible without
the constant support on administrative matters in Colombia from my sister in law Martica Sanchez de
Osorio, my friends Jorge Rubiano, his wife Myriam, Hans Bartels and Luis Enrique Cruz. I also deeply
acknowledge my mother Lolita, and my sisters and brother. Moral support and love from my wife
Gladytas and my sons Cesar and Daniel, is beyond words can describe. I want to dedicate this thesis to
them and to the memory of my father.

xiii
SUMMARY

The focus of this thesis is on the tectonic basin forming and inversion history of the Eastern
Cordillera (EC) of Colombia in terms of the geodynamic processes that govern deformation of the
lithosphere. This goal is pursued through compilation of local data into a regional geological model,
analysis and quantitative modelling of tectonic subsidence and basin-formation mechanisms. To
constrain the alternative possible tectonic scenarios related to the Palaeogene tectonic history, I used
quantitative models to test plausible different tectonic scenarios. More than 100 stratigraphic columns
and wells of the EC, Llanos Orientales (LLA) and Magdalena Valley (MV) were compiled, as well as
previous stratigraphical and paleogeographic interpretations, to draw a series of paleogeographic and
original thickness maps, for several time intervals, covering the Mesozoic and Cenozoic. These maps
help to unravel the complex rifting and basin inversion history of the EC of Colombia.

Mesozoic rifting history of the Eastern Cordillera.

During the Triassic and Jurassic tensional/transtensional stresses, probably initially related to
the break-up of Pangea and later to backarc extension, produced lithosphere stretching and generated
narrow rifts (< 150 km wide), located in the places of the present day MV and the western flank of the
EC (Magdalena-Tablazo sub-basin). During the Early Cretaceous tensional/transtensional stresses
probably related to backarc extension produced new episodes of lithosphere stretching and generated a
wide (> 180 km wide) system of asymmetric half-rift basins.
Subsidence analysis through backstripping and forward modelling of these stratigraphic
columns and wells allow to identify five stretching events during the Mesozoic. Using forward
modelling techniques, assuming uniform and two layered stretching, I calculated for each stratigraphic
column or well the coeval crustal and sub-crustal lithosphere stretching factors. For the Triassic and
Jurassic episodes I could only calculate stretching factors for the whole lithosphere due to limited
resolution on time data. The main rifted events comprise:
(1) Triassic (comprised between 248 and 235 Ma, according the geological time scale proposed
by Gradstein and Ogg, 1996). This stretching event produced narrow rifts (< 150 km wide),
located in the place of the present day Upper Magdalena Valley (with lithosphere stretching
factor up to 1.17), the Serrana de San Lucas (lithosphere stretching factor up to 1.23), and less
subsiding rift basin at the current location of the western flank of the EC (lithosphere stretching
factors up to 1.13).
(2) Latest Triassic to Middle Jurassic (comprised between 208 and 185 Ma.) and stretching
factors up to 1.12 in the Upper Magdalena Valley.
(3) Middle Jurassic (comprised between 180 and 176 Ma.) with stretching factors up to 1.39 in
the western flank of the EC. These two latest Triassic to Middle Jurassic and Middle Jurassic
stretching events produced a relatively narrow rift basin (< 150 km wide) at the current location
of the north-western flank of the EC, west of Bucaramanga. (Moreover, assuming the thickness
of the post-rift section was large along the western flank of the EC north-west of Bogota, part
of Jurassic Cretaceous subsidence is likely to relate to the evolution of these early thermal
events). In addition, if the thickness of the post-rift Cretaceous section in the western flank of
the EC northwest of Bogota is large, it is possible from a two-layered model to hypothesize that
this area was thermally weakened by a Jurassic-stretching event.
(4) Berriasian to Hauterivian (144 to 127 Ma.). This Early Cretaceous stretching event generated a
wide (> 180 km wide) asymmetric half rift basin with a depocenter located along the palaeo-
eastern flank of the EC and a major normal fault system in its eastern border. A system of horst
blocks was also located in the area of the Santander and Floresta massifs. A less developed
second order half rift occurred in the place of the southern western flank of the Cordillera. Two
layered stretching models with stretching factors up to 1.66 for the crust and up to 3.49 for the
subcrustal lithosphere suggest that some decoupling occurred between the crust and subcrustal
lithosphere, or that an increased thermal thinning affected the mantle lithosphere. In places of
maximum crustal and subcrustal lithosphere stretching (greater than 1.4), small mafic intrusions
were emplaced during the Cretaceous (Fabre and Delaloye ,1982).

xiv
(5) Aptian to early Albian (121 to 102.6 Ma.). This stretching event affected mainly the palaeo-
western flank of the EC north-west of Bogot, and the palaeo-Upper Magdalena Valley
whereas a major normal fault system was active along the western border of the Cordillera.
Subsidence curves suggest that during Late Cretaceous, subsidence was produced by thermal
relaxation of the lithosphere.
Periods of rift activity correlate in time with gaps of subduction related magmatic arc activity as
suggested by Aspden et al. (1987) especially for Jurassic time supporting the hypothesis of back arc
extension, which is also supported by volcaniclastic input from the west as indicated by paleocurrent
data and local westerly onlap terminations on the basement. If backarc extension continued during the
Early Cretaceous by oblique plate convergence, it probably had a strong strike-slip component, as
suggested by Aspden et al.. (1987).
Initial Triassic rift basins were narrow but increased in width during Triassic and Jurassic
times. Cretaceous rifts were wider, and were more asymmetrical than Triassic-Jurassic rift basins.
During Berriasian-Hauterivian time the eastern side of the rift possibly was developed by reactivation
of an older Palaeozoic rift system associated to the Guaicramo fault system (c.f. Hossack, et al., 1999).
The western side probably developed by reactivation of an earlier normal fault system developed
during Triassic-Jurassic rifting. Probably the increasing width of the rift system was the result of
progressive tensional reactivation of pre-existing upper crustal weakness zones. Lateral changes of
Mesozoic sediment thickness suggest that the reverse or thrust faults that now define the eastern and
western borders of the EC are largelly controlled by former normal faults that became inverted during
the Cenozoic Andean orogeny. The oblique orientation of most of them relative to the Mesozoic
magmatic arc of the Central Cordillera may be the result of oblique slip extension during Mesozoic, or
alternatively can be inherited from the pre-Mesozoic structural grain. However, not all the Mesozoic
extensional faults were inverted, some normal faults were passively transported with short-cut
basement blocks during Cenozoic inversion (e.g. Esmeraldas Fault, ESRI and Ecopetrol, 1994, Cooper
et al., 1995).
Repeated Mesozoic stretching events in the same area suggest strain localisation affecting a
weak lithosphere. Thermal heating associated with stretching and reactivation of crustal discontinuities
probably contributed to such a strain localisation.
Thermal processes were more dominant than mechanical stretching during Late Triassic-Early
Jurassic phase, in contrast to the Cretaceous rifting phase. During Late Triassic-Early Jurassic abundant
volcaniclastic rocks suggest a positive thermal anomaly in the lithosphere but a moderate lithosphere
stretching. Triassic-Jurassic age unconformities could have been produced by thermal uplift (active
rifting?). In contrast during the Cretaceous less abundant volcanic rocks, absence of tectonically
controlled unconformities, and the large amount of tectonic subsidence indicates absence of thermal
doming. The presence of minor mafic intrusions coinciding with places of maximum crustal and mantle
subcrustal stretching, suggest that a modest magmatism took place as a consequence of extension of the
lithosphere (passive rifting).

Palaeogene incipient basin inversion history of the Eastern Cordillera.

During Palaeogene accretion of the oceanic terranes that now form the Western Cordillera
correlated with incipient inversion of the Mesozoic extensional basin, where continental to coastal plain
and estuarine sediments were deposited.
For several time intervals from late Maastrichtian to Early Miocene the tectonic subsidence was
calculated and compared to the thermally driven subsidence that would be produced after the Mesozoic
stretching events that occurred in the area, assuming a scenario of tectonic quiescence. The great
difference between the two scenarii suggests that the hypothesis of tectonic quiescence is not valid for
the Palaeogene of the EC. Numerous evidences such as the occurrence of Palaeogene unconformities,
the regional Eocene unconformity which locally truncates structures and other local unconformities,
lateral changes of facies and thickness, local erosion indicated by detrital composition of sandstone and
limited fission track data, all suggest that an incipient inversion of Mesozoic extensional basins
occurred during Palaeogene time.

xv
Subsidence in the pre-Andean LLA, palaeo-EC and MV during Palaeogene time cannot be only
explained by the topographic load of the palaeo-Central Cordillera as if the basin were a simple
foreland basin as interpreted by some authors (e.g. Cooper et al., 1995). Palaeogene subsidence could
only be modelled assuming three different subsidence components: (1) residual thermal subsidence
after Mesozoic rifting, (2) flexural subsidence of the lithosphere due to topographic load of the Central
Cordillera, and (3) flexural subsidence of the lithosphere produced by incipient topography generated
during the Palaeogene in the vicinity of the EC flanks. Flexural subsidence models assuming the
lithosphere behaves as an elastic plate of laterally variable thickness, including these three subsidence
components along four regional cross sections, suggest that during the Palaeogene local topography (up
to 500 m) was developed close to the borders of the former Mesozoic extensional basin probably by
inversion along the border extensional faults. However such topography was probably discontinuous
and low enough to significantly disturb the sedimentary and palaeocurrent pattern in the Palaeogene
basin. Structural kinematic modelling along a regional cross section suggests that the amount of
shortening during Palaeogene necessary to produce such Palaeogene topography is small and
dependent of the dip angle of the Palaeogene contraction faults. Assuming a dip angle close to 30 and
using the structural cross section interpretation by Cooper et al. (1995) the modelled total amount of
shortening at the end of Early Miocene was less than 10 km. This early inversion episode accounts for
Palaeogene deformation. A direct consequence of this interpretation is the possibility of generation of
hydrocarbon traps during Palaeogene, a time when petroleum generation and migration occurred in the
place of the EC according to petroleum system modelling results published in the literature (e.g. Mora,
1996).
Palaeogene basin inversion was related to the collision of oceanic plateau terranes with the
north-western margin of South America (Nivia, 1987; Kerr et al., 1996, 1997; Sinton et al., 1998).
Right-lateral transpressional deformation likely lead to a pre-Andean orogeny in the Central Cordillera
during the Palaeogene. Periods of development of basin-inversion and compressional structures seem to
correlate with times of high convergence rate. Development of compressional/transpressional structures
suggests some mechanical coupling between the orogenic wedge, represented mainly the Central
Cordillera, and the regions east of it (MV, EC, and LLA) during Palaeogene time.
Modelling results require very low values of effective elastic thickness (EET) in the area of the
former extensional basin. The Mesozoic rifting events reduced significantly the strength of the
lithosphere, making it very prone to Palaeogene deformation and to further Andean deformation. In the
EC area during Palaeogene the reduced EET values (<10 km) inferred from flexural models suggest
that lithosphere strength was represented only by the upper crust strength (c.f. Cloetingh and Burov,
1996). This compressional intraplate deformation was restricted to crustal levels involving simple
shear-type detachment of the crust at the level of the rheologically weak lower crust from the mantle
lithosphere. Incipient inversion of Mesozoic extensional basins occurred by reactivation of previously
existing discontinuities, such as Mesozoic extensional faults. The occurrence of earlier deformation in
the western side: the MV (e.g. George et al., 1997; Restrepo-Pace et al., 1999a,b) as compared to the
EC or LLA foothills suggests basin inversion progradation. This requires that basins located more
proximal to the collision front were characterized by a mechanically weaker pre-inversion lithosphere
than that of more distal basins.

Neogene basin inversion history of the Eastern Cordillera.

During the Neogene the Mesozoic extensional basin was completely inverted. During Middle
Miocene to Pliocene Andean deformation and uplift occurred mainly involving inversion of the original
Early Cretaceous extensional basin. Uplift of the EC separated the LLA and MV basins. Backstripping
analysis of the stratigraphic record of some wells or stratigraphic columns from the EC, LLA and MV
and flexural models assuming the lithosphere behaves as an elastic plate has successfully tested the
hypothesis that uplift of the EC loaded the lithosphere, thus creating accommodation space in the LLA
and MV basins, which were mainly filled with molassic sediments. Flexural subsidence produced by
the topographic load of the EC explains the observed Neogene subsidence in the LLA and MV.
Irrealistic high bending stresses predicted by flexural models under the EC and the better fit obtained
by flexural models assuming a broken plate under the EC support the hypothesis of mantle lithosphere

xvi
subduction (or at least thinning of the lithospheric mantle due to the rise of the isotherms) under the
EC. Flexural models results for different stratigraphic markers within the Neogene sedimentary record,
and comparison with available fission track data and other evidence of deformation/uplift suggest that
during Miocene time important surface uplift (> 1000 m) and exhumation occurred mainly at the
locations of the margins of the EC. In the middle axial Tunja-Sabana de Bogot region where the
Palaeogene sedimentary record has been partially preserved, important uplift only occurred later during
Pliocene time as recorded by the exceptional palynological record of the Sabana de Bogot. Such
uplifting history can be explained in terms of basin inversion: contractional reactivation of Mesozoic
extensional faults initiated during Palaeogene times leading to an initial slight basin inversion, then
during Neogene time the compressional deformation increased in rate and magnitude leading to a
complete inversion of the original extensional basin. Complete inversion of the master normal fault
systems delimiting the former extensional basins probably lead to extrusion of the sedimentary fill of
half graben basins that now form the eastern and western flanks of the EC.
Maximum dip-slip displacement and shortening occurred on the thrust faults connected to
inverted Mesozoic normal faults that now approximately delimit the Lower Cretaceous outcrops of
both flanks of the EC (e.g. Colletta et al., 1990; Cooper et al., 1995). In these uplifted flanks Lower
Cretaceous or older exposed rocks indicate that all Late Cretaceous or younger sediments have been
eroded. Preservation of the Palaeogene sedimentary record in the axial Bogot-Tunja zone of the EC, as
well as the Neogene sedimentary record of the Sabana de Bogot area suggest that this axial region
remained low during Miocene time but was uplifted since the Pliocene.
One of the major controls on the effective elastic thickness of the study area is the thermal age
of the lithosphere. Based on the thermal age and the values obtained for elastic thickness, three regions
can be distinguished in the study area: (1) The LLA basin where effective elastic thickness has a
maximum value of 50 to 55 km, the thermal age of its eastern lithosphere being Palaeozoic (this
correlation suggests a relatively stabilized strong lithosphere in this area); (2) The EC-MV, where
effective elastic thickness has values of 25 km or less, indicating a weak lithosphere, the thermal age of
which being about 120 20 Ma (this weak thermal destabilized lithosphere is prone to deformation;
weakness there is inherited from the former Mesozoic extensional basins); (3) The Central Cordillera
(including westernmost MV) with EET values of 5 km and very young thermal age as indicated by the
presence of a recent volcanic arc in this area.
EET values lower than 10 km occur in the northern and southern parts of the EC, where major
strike-slip faults such as the Santa Marta-Bucaramanga, Bocono and Altamira faults are located.
Regionally Andean deformation affected weak lithosphere; however, the weakest lithosphere is
associated with strike-slip faults. Such faults probably reach deep enough into the lithosphere to reduce
its total strength. Results of flexural modelling indicate crust-mantle decoupling. In the EC mantle
began to contribute to lithospheric strength probably only since Neogene time. In the weak EC region,
lithospheric strength increased during Neogene time.
Two major controls on lithosphere dynamics of the extensional basin formation and inversion
history of the EC have been (1) lithosphere rheology, and (2) plate-related tectonic stresses. Standard
rheological models and flexural models indicate that during the Mesozoic and Palaeogene a weak
lithosphere resulted from lithosphere stretching in the area of Mesozoic extensional basins. However,
flexural models indicate a weak lithosphere under the EC in open disagreement with standard
rheological models calculated for the present time, that suggest a strong lithosphere in the EC. Two
important effects not considered in the standard rheological models may explain the difference: (1) The
presence of crustal discontinuities that weaken the lithosphere; and (2) The presence of an anomalous
heat input into the lithosphere, as indicated by local Neogene volcanic rocks, and several hot springs in
the EC.

Map view restoration on Neogene transpressional basin inversion of the Eastern Cordillera.

The Northern Andes of Colombia and Venezuela (where the EC is located) represent a broad
zone of deformation resulting from the interaction of the South American, Nazca and Caribbean plates.
Such interaction has fragmented the NW corner of the South American Plate into a number of micro-
plates or tectonic blocks: Guajira, Santa Marta, Maracaibo and Central Colombia (Colombian Central

xvii
Cordillera). Relative movements of the micro-plates resulted in deformed belts with a combination of
compressional thick/thin-skinned thrusting, folding and strike-slip faulting which form several
mountain ranges or strike-slip fault zones in the Northern Andes. I have manually restored in map view
a mosaic of fault-bounded blocks using all the kinematic constraints available in literature, such as
strike-slip displacement estimates and amounts of shortening from available structural balanced cross-
sections. In order to simplify the model I have used a limited number of blocks. This technique is
useful to constrain shortening and strike-slip displacement estimates as well as to detect strike-slip
motions and rotations about vertical axes not revealed by balancing cross-sections. According to this
regional map restoration, the following relative movements of regional blocks leading to deformation
belts were interpreted: (1) Eastward convergence and clockwise rotation of the Central Colombia
micro-plate relative to South America resulted in transpression of the EC, with northward increasing
shortening in this mountain range; (2) Eastward convergence and left-lateral strike-slip movement of
Central Colombia relative to Maracaibo produced the northern part of the EC in the Santander Massif
region; (3) North-eastward convergence of the Santa Marta Block relative to Maracaibo generated the
Perij Mountain Range; and (4) South-eastward convergence and right-lateral strike-slip motion of the
Maracaibo Block relative to South America generated the Mrida Andes of Venezuela.
A more detailed map view restoration of the EC of Colombia was done using balanced cross-
sections from the literature. According to the restoration the amount of shortening during Andean
deformation is approximately one half of the present-day width of the Cordillera. Both width and
amount of shortening increase northward. Results of the map view restoration are in general supported
by kinematic indicators (fault striae) and other outcrop structural data, borehole breakout data and the
focal mechanism solutions of upper crustal earthquakes. These results indicate: (1) ENE-WSW
shortening perpendicular to the regional structural grain of the EC; (2) Conjugate (?) or pseudo-
conjugate (?) left-lateral and right-lateral strike-slip faults; (3) Clockwise rotation of the Central
Cordillera, MV and western flank of the EC; (4) The SE flank of the EC (Cocuy and eastern
Cundinamarca sub-basin) was right-lateral transpressively deformed; (5) The NW flank of the EC
(Magdalena-Tablazo sub-basin) was left-lateral transpressively deformed; (6) Andean deformation
generated the EC through transpressive inversion of Mesozoic extensional basins. The western part of
the EC was affected by N-S sinistral transpression, while the eastern part was affected by E-W dextral
transpression.
Comparison of displacement of blocks relative to stable South America obtained from the map
view restoration, with plate velocity vectors from geophysical data, suggest that the Andes Block as a
whole is at present time moving with a dominant right-lateral strike-slip component along the faults at
its easternmost boundary. This strike-slip component probably was less important during the whole
period of Neogene Andean deformation, as suggested by the map view restoration.

Hypothesis about the deep structure of the Eastern Cordillera.

Based on surface geological data, limited geophysical data (gravity, seismicity), comparison
with similar mountain belts and analogue and numerical models from the literature, it is possible to
narrow the uncertainity range on the deep structure of the EC. Probably it resembles that of the
Pyrenees: Lithospheric shortening has been accommodated in the upper brittle crust by development of
a double vergent asymmetric wedge, while the mantle lithosphere accommodated shortening by west-
dipping subduction of the cooler and denser LLA mantle lithosphere under the buoyant and hotter
Andean lithosphere. The lower ductile crust probably accommodated shortening by thickening.
Probably the EC has been strongly affected by transpression with important transcurrent components.
Intermediate seismicity beneath the EC suggest that a subducted slab fragment is present below
its NW margin (Taboada et al., 1999, 2000). A small, but very active zone of intermediate-deep
seismicity, the Bucaramanga earthquake nest (Schneider et al., 1987), may result from deep oblique
convergence of Palaeo-Caribbean plate fragment connected to the Panam Block and the Caribbean
Plate from the north. This may explain the NW-SE lineation of the nest (Schneider et al., 1987). Two
particularities of the EC, i.e. the local presence of Neogene volcanic rocks at Paipa and Iza, and a
intermediate seismicity, may be the expression of slab break-off of an eastward-subducting Caribbean-
type mantle lithosphere under the EC. Subduction resistance of the relatively buoyant Caribbean type

xviii
crust (Burke, 1988; Kerr et al., 1997) may have produced break-off of its denser mantle lithosphere
under the Colombian Andes or a fragment of a normal denser lithosphere of the old Farallon Plate. The
deep slab suggested by tomographic images (Taboada et al., 2000) may be a subducted slab, which is
no longer connected to the surface plates. This hypothesis would explain the east-dipping slab fragment
suggested by seismological studies (Pennington, 1981; Schneider et al., 1987; Taboada et al., 1999,
2000) and the Paipa-Iza volcanic rocks. These volcanic rocks may be related with upwelling hot
asthenosphere where slab break-off occurs. Possibly partial melting of the lower crust beneath the EC
may have generated rising magmas, generated a thermal anomaly under the EC and weakened its
lithosphere.
Deep seismic studies could offer reliable data in order to support or contradict the suggested
hypotheses.

xix
SAMENVATTING

Het doel van dit proefschrift is een bijdrage te leveren aan het begrip van hoe verschillende
geodynamische processen, die vervorming van de lithosfeer bepalen, hebben bijgedragen aan het
ontstaan van een bekken en de verandering van het bekken in een gebergte (inversie) in het gebied van
de Cordillera Oriental (Eastern Cordillera, EC) van Colombia.
Dit doel is nagestreefd door (1) een compilatie van lokale gegevens in een regionaal geologisch
model, (2a) analyse en (2b) kwantitatieve modellering van daling door tektonische processen en
bekkenvormende mechanismen. Verschillende mogelijke tektonische scenarios zijn getoetst om te
bepalen welke daarvan het meest waarschijnlijk is voor de minder goed bekende tektonische
geschiedenis van de EC tijdens het Paleogeen. Meer dan 100 stratigraphische kolommen en putten in
de Cordillera Oriental (EC), de Llanos Orientales (LLA) en de Valle del Magdalena (Magdalena
Valley, MV) zijn samengevoed met al bestaande stratigrafische en paleogeografische interpretaties om
een serie kaarten te construeren voor verschillende tijdsintervallen tijdens het Mesozo cum en
Cenozo cum van de paleogeografie en oorspronkelijke dikte van de verschillende gesteentepakketten.
Deze kaarten helpen om de complexe ontstaans- en vervormingsgeschiedenis van het bekken te
ontrafelen.

Mesozo sche rek geschiedenis van de Cordillera Oriental.

Gedurende de Trias en Jura veroorzaakten tensionele en transtensionele spanningen het


ontstaan van z.g. "narrow rifts" (<150km breed) in de huidige MV en de Westflank van de EC
(Magdalena-Tablazo sub-bekken). Deze spanningen zijn vermoedelijk gerelateerd aan het uiteenvallen
van Pangea en vervolgens back-arc extensie. Tijdens het Vroege Krijt veroorzaakten back-arc
erelateerde spanningen nieuwe periodes van lithosfeer verdunning in een systeem van z.g. "wide rifts"
(>180km breed), met asymmetrische half-grabens.
Vijf verschillende Mesozo sche perioden van rek (rifting) kunnen worden onderscheiden uit
een analyse van de dalingsgeschiedenis (backstripping) en door voorwaartse modellering van de
stratigrafische kolommen of boringen. Met de voorwaartse modellering, waarbij werd uitgegaan van
uniforme rek in twee lagen, kon voor elke stratigrafische kolom of boring bepaald worden met welke
factor zowel de korst als het sub-korst gedeelte van de lithosfeer werden verdund (stretching factors).
Voor de perioden in Trias en Jura kon door een gebrek aan goede resolutie van de gegevens alleen een
algemene stretching factor voor de hele lithosfeer berekend worden.
De belangrijkste perioden van rek zijn:
(1) Trias (248 tot 235 Ma). Deze periode van rek produceerde smalle rek-bekkens (< 150 km breed),
gelokaliseerd op de huidige plaats van de MV (met lithosfeer stretching factor tot 1.17), de
Serrana de San Lucas (lithosfeer stretching factor tot 1.23), en een minder hard dalend rek-bekken
op de huidige locatie van de westelijke flank van de EC (lithosfeer stretching factors tot 1.13);
(2) Laatste Trias tot Midden Jura (208 tot 185 Ma.) met stretching factors tot 1.12 in de Upper MV;
(3) Midden Jura (180 tot 176 Ma.) met stretching factors tot 1.39 in de westelijke flank van de EC.
Deze twee perioden van rek leidden tot een relatief smal bekken (< 150 km wiid) op de positie van
de huidige noordwestelijke flank van de EC, ten westen van Bucaramanga. (Als aangenomen
wordt dat de dikte van de post-rift section aanzienlijk was langs de westelijke flank van de EC in
dit gebied ten noord-westen van Bogota, kan met een model dat uitgaat van twee sterke lagen in
de lithosfeer, de hypothese worden opgesteld dat dit gebied thermisch verzwakt werd door een
riftperiode in de Jura. Een deel van de daling in Jura en Krijt is waarschijnlijk gerelateerd aan de
ontwikkeling van deze vroege thermische events).
(4) Berriasian tot Hauterivian (144 tot 127 Ma.). Deze Vroeg Krijt riftperiode genereerde een breed (>
180 km breed) asymmetrisch half rek-bekken met een depocenter langs de paleo-oostelijke flank
van de EC en een groot systeem van afschuivingsbreuken langs de oostelijke rand. Andere
gebieden waarin een systeem van horsten ontstond zijn de tegenwoordige massieven van
Santander en Floresta. Een model van rek in twee lagen leverde stretching factors op van 1.66
voor de korst en tot 3.49 voor het subkorst deel van de lithosfeer. Dit suggereert dat een zekere
mate van ontkoppeling plaats vond tussen de korst en het lithosfeer gedeelte onder de korst, of dat

xx
een verhoogde thermische activiteit het mantel gedeelte van de lithosfeer verdund heeft. Op de
plekken met maximale stretchings factors ontstonden kleine mafische intrusies tijdens het Krijt
(Fabre and Delaloye ,1982);
(5) Aptian tot Vroeg Albian (121 tot 102.6 Ma.). Deze periode van rek be nvloedde voornamelijk de
paleo-westelijke flank van de EC ten noord-westen van Bogot, en de paleo-Valle Superior del
Magdalena, terwijl een belangrijk systeem van afschuivingsbreuken actief was langs de westelijke
grens van de EC. Dalings curves suggereren dat tijdens Laat Krijt daling werd veroorzaakt door
thermische relaxatie van de lithosfeer.

Perioden in rift-activiteit correleren met tijdsintervallen waarin de activiteit van de vulkaanboog


afwezig was volgens Aspden et al. (1987), in het bijzonder tijdens de Jura. Dit ondersteunt de
hypothese dat de rift ontstond als een back-arc extensie, rek achter een subductie gerelateerde boog,
wat ook blijkt uit input van volcaniclastic uit het westen (zoals aangetoond door gegevens over paleo-
stroomtrichting) en lokaal westwaarts onlap terminations over het basement. Als back-arc extensie
gedurende het Vroeg Krijt doorging door zijdelingse convergentie van platen, dan had het daardoor een
grote zijwaartse component, zoals gesuggereerd door Aspden et al. (1987).
De rek-bekkens waren oorspronkelijk in het Trias smal, maar verbreedden zich gedurende
Trias-Jura. De Krijt bekkens waren nog breder en meer asymmetrisch dan de Trias-Jura rek-bekkens.
Tijdens Berriasian-Hauterivian ontwikkelde de oostzijde van het bekken zich waarschijnlijk door
reactivatie van een Paleozoisch rift systeem, geassocieerd met het Guaicramo breuk systeem (c.f.
Hossack, et al., 1999). Ook de westelijke zijde ontwikkelde zich waarschijnlijk door reactivatie, in dit
geval reactivatie van een systeem afschuivingsbreuken dat was ontstaan tijdens de rek in Trias en Jura.
De verbreding van het systeem was een resultaat van progressieve reactivatie onder rek van reeds
bestaande zwaktezones in de bovenkorst. Laterale veranderingen in de dikte van de Mesozo sche
sedimenten wijzen er op dat de over- en opschuivingsbreuken die nu de oostelijke en westelijke rand
vormen van de EC voornamelijk zijn ontstaan uit oudere afschuivingen die werden ge nverteerd tijdens
de Andean gebergte vorming in het Cenozo cum. Het hoekverschil tussen de meeste breuken en de
a) zou het resultaat kunnen zijn van oblique slip rek
tijdens het Mesozo cum, of als alternatief, overgenomen van al bestaande structuren in de korst voor
het Mesozo cum. Niet alle Mesozo sche afschuivingbreuken werden overigens gereactiveerd,
sommigen werden passief meegevoerd in basement blokken tijdens de inversie in het Cenozo cum.
Herhaalde perioden van rek in het Mesozo cum in dezelfde gebieden suggereren dat de
vervorming gelokaliseerd werd in zwakke delen van de lithosfeer. Opwarming die samenhangt met rek
en reactivatie van discontinu teiten in de korst hebben waarschijnlijk bijgedragen aan deze lokalisatie
van vervorming.
Het veelvuldig voorkomen van Laat Trias tot Vroeg Jura volcaniclastische gesteenten
suggereren een positieve thermische anomalie in de lithosfeer, maar slechts beperkte lithosferische rek.
Unconformiteiten van Trias-Jura ouderdom zouden gevormd kunnen zijn door thermische opheffing
(active rifting?). Thermische processen waren dominanter dan mechanische rek tijdens Laat Trias tot
Vroeg Jura, in tegenstelling tot de rekfase in het Krijt toen thermische opheffing waarschijnlijk afwezig
was, zoals blijkt uit de volgende feiten: minder veel voorkomen van volcaniclastische gesteenten, de
afwezigheid van unconformiteiten -veroorzaakt door tektoniek- en de grote tektonische daling. De
aanwezigheid van kleine mafische intrusies komen overeen met regios van maximale korst en mantel
rek, wat er op wijst dat beperkte magmatische activiteit plaats vond als gevolg van rek in de lithosfeer
(passieve rifting).

Paleogene start van bekken-inversie in de EC.

Tijdens het Paleogeen correleert de accretie van fragmenten oceanische korst, die nu de
Westelijke Cordillera vormt, met het begin van de inversie van het Mesozo sche rek-bekken waarin
continentale tot kustvlakte en estuariene sedimenten werden afgezet.
Voor verschillende tijdsintervallen van Laat Maastrichtian tot Vroeg Mioceen is de tektonische daling
berekend en vergeleken met de daling gedreven door thermische activiteit die zou worden
geproduceerd na de rekperiode die plaatsvond in het gebied, uitgaand van een scenario met tektonische

xxi
rust na de rek. De grote verschillen tussen de twee scenarios suggereren dat de hypothese van
tektonische rust niet geldig kan zijn voor het Paleogeen in de EC. Verschillende feiten wijzen ook op
een beginnende inversie van het Mesozo sche rek-bekken tijdens het Paleogeen: (1) Paleogene
unconformiteiten, (2) de regionale Eocene unconformiteit die lokaal oude structuren en
unconformiteiten afsnijdt, (3) laterale verandering van facies en dikte, (4) lokale erosie als aangetoond
door de detritische samenstelling van zandsteen, en (5) beperkte gegevens van een splijtsporen-analyse.
Daling in de pre-Andean LLA, EC en MV tijdens het Paleogeen kan niet alleen worden
verklaard door topografische lading van de Centrale Cordillera (CC), alsof het bekken een eenvoudig
voorlandbekken zou zijn, zoals wel ge nterpreteerd is door sommige auteurs e( .g. Cooper et al., 1995).
Paleogene daling kan alleen gemodelleerd worden door drie verschillende componenten van de daling
mee te nemen: (1) een residu van thermische daling na de Mesozo sche rek (2) daling ten gevolge van
flexuur van de lithosfeer door topografische lading door de CC en (3) daling ten gevolge van topografie
die ontstond tijdens het Paleogeen in de buurt van de flanken van de EC. Dalingsmodellen die flexuur
gebruiken nemen aan dat de lithosfeer zich gedraagt als een elastische plaat van een eventueel lateraal
variabele dikte. Als we de drie genoemde componenten van daling mee nemen in vier regionale
profielen, suggereert dat dat de Paleogene locale topografie (tot 500m hoog) ontstond dicht bij de
randen van het vroegere Mesozo sche rek-bekken, waarschijnlijk door inversie van de randbreuken. De
topografie was waarschijnlijk discontinu en laag genoeg om het sedimentaire patroon en paleo-
stroomrichting van het Paleogene bekken behoorlijk te be nvloeden. Kinematische modellering langs
een regionale sectie suggereren dat er maar weinig verkorting nodig was om deze Paleogene topografie
te vormen, afhankelijk van de hellingshoek van de breuken waarlangs de verkorting plaatsvindt.
Uitgaande van een hellingshoek dicht bij 30 graden, en gebruikmakend van de interpretatie van een
structureel profiel door Cooper et al (1995), is de gemodelleerde hoeveelheid verkorting aan het eind
van het Vroeg Mioceen minder dan 10 km. Een direct gevolg van deze interpretatie is de mogelijkheid
van ontwikkeling van hydrocarbon traps tijdens het Paleogeen, een tijdsinterval waarin volgens de
literatuur ontwikkeling en migratie van petroleum plaatsvond in de EC (e.g. Mora, 1996).
Paleogene bekken-inversie was gerelateerd aan de botsing van oceanische plaatfragmenten met
de noordwestelijke marge van Zuid Amerika (Nivia, 1987; Kerr et al., 1996, 1997; Sinton et al., 1998).
Rechtsom zijwaartse transpressieve vervorming leidde vermoedelijk tot een pre-Andean
gebergtevorming in de CC tijdens het Paleogeen. Perioden van ontwikkeling van bekken-inversie en
compressieve structuren lijken overeen te komen met perioden met een hoge convergentie-snelheid. De
ontwikkeling van compressieve of transpressieve structuren suggereert een zekere mechanische
koppeling tussen de orogene wig, bestaand uit voornamelijk de CC, en de regios oostelijk van de CC
tijdens het Paleogeen (MV, EC en LLA).
De modelresultaten vereisten erg lage waarden voor de Effectieve Elastische Dikte (EET) in het
gebied van het voormalige rek-bekken. De Mesozo sche rekperioden reduceerden de sterkte van de
lithosfeer aanzienlijk, waardoor het gebied zowel in het Paleogeen als de Andean gebergtevorming
betrokken raakte. Tijdens het Paleogeen werd de sterkte van de lithosfeer bepaald door alleen de sterkte
van de bovenkorst (c.f. Cloetingh and Burov, 1996), volgens de waarden voor de EET die voortkomen
uit de flexuurmodellen. Deze compressieve intraplate vervorming was beperkt tot de korst niveaus door
middel van simple shear type ontkoppeling van de korst en mantel lithosfeer langs het niveau van de
zwakke onderkorst. Het eerder voorkomen van vervorming in de westelijke zijde (MV, e.g. George et
al., 1997; Restrepo-Pace et al., 1999a,b) in vergelijking met de EC of LLA suggereren een progradatie
van bekken-inversie. Dit vereist dat de bekkens die dichter bij de botsing waren gelokaliseerd,
gekarakteriseerd worden door een mechanisch zwakkere lithosfeer voor de inversie dan de verder weg
gelegen bekkens.

Neogene bekken-inversie in de Cordillera Oriental.

Tijdens het Neogeen werd het Mesozo sche rek-bekken volledig ge nverteerd. Andean
vervorming en opheffing gebeurde van Midden Mioceen tot Plioceen vooral door de inversie van het
originele Vroeg Krijt rek-bekken. Opheffing van de EC scheidde het LLA- van het MV-bekken. Zowel
backstripping van de stratigrafische opeenvolging van een aantal putten en stratigrafische kolommen
van de EC, LLA en MV als flexuurmodellen laten zien dat de opheffing van de EC de lithosfeer laadde

xxii
zodat er accomodatie ruimte ontstond in de LLA en MV bekkens die voornamelijk werden gevuld met
molasse sedimenten. Onrealistisch hoge buigspanningen die door flexuurmodellen voorspeld worden
onder de EC en de betere overeenstemming met de waarnemingen behaald met flexuurmodellen die
uitgaan van een gebroken plaat onder de EC, steunen de hypothese van mantel lithosfeer subductie (of
op zijn minst verdunning van het mantel gedeelte van de lithosfeer door een stijging van de
isothermen). De resultaten van de flexuurmodellen voor de verschillende stratigrafische niveaus in de
Neogene sedimentaire opeenvolging en vergelijking met beschikbare splijtingssporen gegevens en
andere aanwijzingen voor deformatie en/of opheffing suggereren dat tijdens Mioceen belangrijke
opheffing en exhumatie (tot >1000m) plaatsvond, voornamelijk langs de marges van de EC. In de
axiale regio (Tunja-Sabana de Bogot), waar de Paleogene opeenvolging gedeeltelijk bewaard is
gebleven vond belangrijke opheffing pas plaats tijdens het Plioceen, zoals is achterhaald uit de
buitengewone palynologische gegevens van de Sabana de Bogot. Een dergelijke geschiedenis van
opheffing kan worden uitgelegd in termen van bekken inversie: reactivatie onder verkorting van de
Mesozo sche rekbreuken begon tijdens het Paleogeen en leidde tot een initiel slecht beperkte inversie
van het bekken, met daarna tijdens het Neogeen een volledige inversie van het bekken. Volledige
inversie van de grote systemen van afschuivingsbreuken die de randen vormden van het vroegere
bekken, leidde waarschijnlijk tot extrusie van de sedimentaire vulling van de half-graben bekkens die
nu de oostelijke en westelijke flanken van de EC vormen.
Een maximum verplaatsing langs de helling van de breuk en verkorting vond plaats langs de
opschuivingen die behoorde bij de ge nverteerde Mesozo sche afschuivingen die nu ongeveer de
voorkomens van Onder Krijt ontsluitingen in beide flanken van de EC begrenzen (e.g. Colletta et al.,
1990; Cooper et al., 1995). In deze opgeheven flanken zijn Onder Krijt en oudere gesteente zichtbaar
die aangeven dat alle Laat Krijt en jongere sedimenten zijn gerodeerd. Dat de Paleogene opeenvolging
is bewaard gebleven in de axiale Bogot-Tunja zone van de EC, als ook de Neogene sedimenten in de
Sabana de Bogot, wijst erop dat dit gebied laag gebleven is tijdens het Mioceen en pas werd
opgeheven vanaf het Plioceen.
Een van de belangrijkste factoren die de EET bepaalt in het bestudeerde gebied is de
thermische ouderdom van de lithosfeer. Gebaseerd op de thermische ouderdom en de waarden
verkregen voor de EET, kunnen drie regios onderscheiden worden: (1) het LLA bekken met EET
waardes tot 50-55km en een thermische ouderdom van Paleozoicum. Dit wijst samen op een relatief
stabiele en sterke lithosfeer, (2) EC en MV, waar de EET ongeveer 25km bedraagt en de thermische
ouderdom om en nabij 120 Ma is. Dit duidt op een thermische gedestabiliseerde lithosfeer die gevoelig
is voor vervorming; en (3) de CC (inclusief het meest westelijke deel van MV) met EET waardes van
slechts 5 km en erg lage thermisch ouderdom, zoals blijkt uit de aanwezigheid van een recente
vulkanische boog in het gebied.
Waarden van minder dan 10km voor de EET komen voor in de noordelijke en zuidelijke
gedeelten van de EC, op plekken waar grote zijschuivingen zoals de Santa Marta-Bucaramanga,
Bocono en Altamira breuken te vinden zijn. Zulke breuken reiken waarschijnlijk diep genoeg in de
lithosfeer om die te kunnen verzwakken. Resultaten van de flexuur modellering wijzen op ontkoppeling
tussen korst en mantel. In de EC nam de sterkte van de lithosfeer toe in het Neogeen, meest
waarschijnlijk begon de mantel toen pas mee te tellen in de sterkte van de hele lithosfeer.
Twee grote factoren die invloed hebben op de lithosfeer dynamiek tijdens de vorming van rek-
bekkens en de inversie van de EC zijn: (1) de rheologie (materiaal eigenschappen) van de lithosfeer, en
(2) tektonische spanningen, samenhangend met plaat-interactie. Standaard rheologie- en
flexuurmodellen geven aan dat een zwakke lithosfeer tijdens Mesozo cum en Paleogeen het resultaat
was van de Mesozo sche rek. De flexuurmodellen wijzen echter op een zwakke lithosfeer onder deEC
voor de huidige situatie, wat in tegenspraak is met de uitkomsten van standaard rheologische modellen
voor dezelfde situatie: die suggereren een sterke lithosfeer. Twee belangrijke effecten die niet
meegenomen zijn in de standaard rheologische modellen zouden de verschillen kunnen verklaren: (1)
de aanwezigheid van discontinu teiten in de korst die de lithosfeer verzwakken, en (2) de aanwezigheid
van een anomale hittebron, zoals aangetoond door Neogene vulkanische gesteenten en verschillende
hete bronnen in de EC.

xxiii
Reconstructie in kaartbeelden van de Neogene transpressieve inversie van het bekken van de EC.

De noordelijke Andes van Colombia en Venezuela (waar de EC zich bevindt), vormt een brede
zone van vervorming als gevolg van interactie tussen de Zuid-Amerikaanse, Nazca en Caribbische
platen. Deze interactie heeft ervoor gezorgd dat de NW hoek van de Zuid-Amerikaanse plaat is
opgebroken in verschillende fragmenten: Guajira, Santa Marta, Maracaibo en Centraal Colombia
(Centraal Colombiaanse Cordillera). Relatieve bewegingen van deze micro-platen zorgde in de
noordelijke Andes voor gebergteketens en zones met een combinatie van compressieve thick/thin-
skinned thrusting, plooi ng en zijschuivingsbreuken. Om de complexiteit van het model binnen de
perken te houden, is maar een beperkt aantal blokken die door de breuken van elkaar gescheiden
worden gebruikt in een handmatige reconstructie. Deze reconstructie is gemaakt aan de hand van
kinematische randvoorwaarden die in literatuur gevonden konden worden, zoals de verplaatsing van de
zijschuivingen en schattingen van verplaasting uit gereconstrueerde profielen door het gebied. Deze
techniek is zeer nuttig om betere schattingen te kunnen maken van de verkorting en verplaatsing en om
rotaties te ontekken die niet in profielen kunnen worden vastgesteld. Volgens deze regionale
reconstructie kunnen de volgende bewegingen van de blokken die hebben bijgedragen aan de
vormingen van de gebergte- gordels, worden vastgesteld: (1) oostwaartse convergentie en rotatie met
de klok mee van de Centraal Colombiaanse micro-plaat ten opzichte van Zuid-Amerika, wat
resulteerde in transpressie in de EC met naar het noorden toenemende verkorting; (2) oostwaartse
convergentie en linksom zijwaartse beweging van Centraal Colombia ten opzichte van Maracaibo, wat
het noordelijk deel van de EC in de regio van het Santander massief vormde; (3) noord-oostelijke
convergentie van het Santa Marta Blok ten opzichte van Maracaibo zorgde voor het ontstaan van het
Perij gebergte; en (4) zuid-oostwaartse convergentie en rechtsom zijwaartse beweging van het
Maracaibo Blok ten opzichte van Zuid-Amerika. Dit veroorzaakte de Mrida Andes van Venezuela.
Een meer gedetailleerde reconstructie op kaartbeelden is uitgevoerd voor de EC,
gebruikmakend van profielen uit de literatuur. Volgens deze reconstructie is de totale verkorting tijdens
de Andes vervorming ongeveer gelijk aan de helft van de huidige breedte van de Cordillera. Zowel
breedte als verkorting nemen toe richting het noorden. Deze resultaten worden over het algemeen
ondersteund door bewegings-indicatoren (bewgingsrichting op breukvlakken) en andere stuctureel
geologische gegevens uit ontsluiting, vervorming van boorgaten, en gegevens over de
verkortingsrichting verkregen uit aardbevingen in de bovenkorst. De reconstructie toont dat: (1) ONO-
WZW verkorting plaatsvond, loodrecht op de regionale structuur van de EC; (2) conjugate (?) of
pseudo-conjugate (?) linksom en rechtsom zijschuivingen zich ontwikkelden; (3) de CC, MV en
westelijke flank van de EC met de klok meegedraaid zijn; (4) de zuidoostelijke flank van de EC (Cocuy
en oostelijke Cundinamarca sub-bekken) vervormd werd door transpressie met een grote component
rechtsom; (5) de noordwestelijke flank van de EC (Magdalena-Tablazo sub-bekken) juist vervormd
werd door transpressie met een grote component linksom; (6) vervorming van de Andes zorgde voor
transpressieve inversie van de oorspronkelijke Mesozo sche rek-bekkens. Het westelijke deel van de
EC werd be nvloed door noord-zuid georienteerde transpressie met linksom beweging, terwijl het
oostelijk deel oost-west georienteerde transpressie met een component rechtsom onderging.
Vergelijking van waarden voor de verplaatsing van blokken ten opzichte van het stabiele Zuid-
Amerika die zijn verkregen door de kaartrecontructies met plaatsnelheden van geofysische gegevens,
suggereren dat het Andes Blok als geheel op het moment met een dominante rechtsom zijwaartse
beweging langs de oostelijke delen beweegt. Deze component van zijwaartse beweging was
waarschijnlijk minder belangrijk tijdens de gehele periode van Neogene vervorming van de Andes,
volgens de reconstructie op kaartbeeld.

Hypothese over de diepe structuur van de EC.

Gebaseerd op geologische gegevens, beperkte geofysische date (zwaartekracht en seismiciteit),


vergelijking met vergelijkbare gebergten en analoge en numerieke modellen uit de literatuur, is het
mogelijk de mogelijke variatie in diepe structuur van de EC. Waarschijnlijk lijkt de diepe structuur van
de EC nogal op die van de Pyreneen. Verkorting van de lithosfeer heeft plaatsgevonden in de brosse
boven korst door een dubbel-vergente asymmetrische wig, terwijl op het niveau van de mantel

xxiv
verkorting mogelijk is geworden door westwaartse subdcutie van de koudere en zwaardere mantel
lithosfeer van de LLA onder de bouyant en warmere Andean litosfeer. De ductiele onderkorst heeft de
verkorting waarschijnlijk opgevangen door verdikking. Waarschijnlijk is de EC sterk beinvloed door
transpressie met een belangrijke transcurrent component.
Seismiciteit van gemiddelde diepte onder de EC suggereert dat een gesubduceerd deel van de
lithosfeer nog aanwezig is onder de NW rand (Taboada et al., 1999, 2000). Een kleine, maar actieve
zone met aardbevingen van gemiddelde diepte (de Bucaramanga earthquake nest (Schneider et al.,
1987), kan het resultaat zijn van scheve convergentie van de oude Caribische plaat, verbonden aan de
Panama en de recente Caribische plaat. Dit kan de NW-ZO verspreiding van deze aardbevingen
verklaren, (Schneider et al., 1987). Twee opvallende eigenschappen van de EC, n.l. de aanwezigheid
van Neogene vulkanische gesteenten (in Paipa en Iza ) en de seismiciteit van gemiddelde diepte,
kunnen een aanwijzing vormen voor het afbreken van het mantel gedeelte van de oostwaards
gesubduceerde Caribische plaat onder de EC. Subductie weerstand van de relatief lichte Caribische
korst (Burke, 1988; Kerr et al., 1997) kan hebben geleid tot het afbreken van de dichtere lithosfeer van
de oude Farallon plaat.
De diepe slab, gesuggereerd op basis van tomgrafische beelden (Taboada et al., 2000) kan een
gesubduceerde plaat zijn, die niet langer is gekoppeld aan oppervlakte plaaten. Deze hypothese kan de
oostduikende plaat uit seismologische studies van (Pennington, 1981; Schneider et al., 1987; Taboada
et al., 1999, 2000) en de Paipa-Iza vulkanische gesteenten verklaren. De vulkanische gesteenten kunnen
gerelateerd zijn aan de stijgende hete asthenosfeer op plaatsen waar de plaat gebroken is. Mogelijke
gedeeltelijke smelt van de onderkorst onder de EC kan hebben geleid tot magma vorming, thermische
anomalien onder de EC en plaatselijke verzwakking van de lithosfeer.
Diep seismische studies zouden betrouwbare gegevens kunnen opleveren om deze hypotheses
te falsifiren of verifiren.

xxv
CHAPTER 1

INTRODUCTION AND OUTLINE

1.1. AIM AND METHODS

The main aim of this thesis is to contribute to understand the tectonic basin forming and
inversion history of the Eastern Cordillera (EC) of Colombia in terms of the geodynamic processes that
govern deformation of the lithosphere. This goal is pursued through compilation of local data into a
regional geological model, analysis and quantitative modelling of tectonic subsidence and basin-
formation mechanisms. To constrain the alternative possible tectonic scenarios related to the not-well
constrained Palaeogene tectonic history, I used quantitative models to test plausible different tectonic
scenarios. Details of the methods are explained in the chapter in which they are applied.

1.2 BASIN FORMATION AND INVERSION MODELS

Sedimentary basin formation mechanisms are basically:

(1) Thermal mechanisms producing subsidence or uplift due to their effects on the lithosphere density
structure (e.g.. Sleep, 1971) and its isostatic balance. An example of this is the cooling and
subsidence of oceanic lithosphere as it moves away from spreading centers.
(2) Changes in lithosphere thickness and structure produced by stretching, erosion or emplacement of
dense material into the lithosphere column leading to isostatic readjustment. An example of this is
mechanical stretching of the crust accompanied by normal faulting occurring in extensional basins
(e.g.. McKenzie, 1978).
(3) Loading of the lithosphere causing lateral subsidence due to regional isostatic effects (flexure) by
mountain building processes, as occurring in foreland basins (c.f. Price, 1973).

In many cases where the lithosphere is affected by rapid stretching, as the case of rifting, lithospheric
thinning and upwelling of hot asthenosphere change its density and thermal structure. The consequent
isostatic readjustment causes fast tectonic subsidence followed by slower thermal subsidence as the
lithosphere thickness gradually increases by heat conduction (McKenzie, 1978). The simple stretching
model of McKenzie (1978) has been followed by more sophisticated quantitative models, including
differential stretching finite rates of extension (Jarvis and McKenzie, 1978), lateral heat conduction
(Royden and Keen, 1980), depth dependent rheology (Sawyer, 1985), and other complex effects.
Assumption of local isostasy implies lack of lateral strength of the lithosphere. More realistic models
assume the lithosphere behaves to applied loads as an elastic plate overlying a weak asthenospheric
mantle. This has led to development of the concept of flexural subsidence that can explain subsidence
observed in foreland basins (Beaumont, 1981).
Initial flexural models applied to foreland basins assumed only topographic loads (Beaumont,
1981). However, later models have suggested other possible subsurface loads. The strength of the
lithosphere controls its flexural behaviour. The distribution of strength in the lithosphere varies
vertically, due to its petrological and thermal structure (Ranalli and Murphy, 1987; Ranalli, 1995;
Cloetingh and Burov, 1996), and laterally as a result of its pre-orogenic history (De Jong, 1991; Van
Wees et al., 1992; Van Wees and Cloetingh, 1993; Van Wees, 1994; Ziegler et al., 1995, 1998).
Extrapolation of rock mechanical, geophysical (including deep seismic) and geological data support the
concept of a rheologically stratified lithosphere characterized by a relatively strong (mostly brittle)
upper crust, a weak (mostly ductile) lower crust, and a strong (brittle and ductile) subcrustal lithosphere
(Ranalli, 1995). This will be discussed in last chapter.
During basin inversion, basin controlling faults reverse their movement due to compressional
tectonics, and basins become positive topographic features (e.g.. Bally, 1984). Numerical models for
thermal compression and exhumation of the lithosphere (England and Thompson, 1986; Davy and
Gillet 1986) have been applied to compressional settings such as those of mountain building and basin
inversion. However, models for the rheological evolution of the lithosphere indicate that the predicted
Introduction Outline
rheological evolution cannot explain repeated reactivation of faults (Van Wees, 1994; Ziegler et al.,
1995, 1998). Observed reactivation of upper crustal faults not aligned with respect to the stress field,
deformation patterns in rifted and inverted basins and other evidence, suggest that in addition to the
standard rheology of the lithosphere, also reactivation of inherited weak fault zones control the
response of the lithosphere to tectonic stress fields (Van Wees, 1994; Ziegler et al., 1995, 1998). To a
large extent, the mechanical processes responsible for sedimentary basin formation and evolution are
thermo-mechanical, inducing deformation in the lithosphere, and related to the presence of pre-existing
discontinuities that have a crucial role in weakening the lithosphere (Van Wees, 1994; Ziegler et al.,
1995, 1998).
In a sedimentary basin the stratigraphic record, including sedimentary fill, unconformities and
other indicators of uplift, reflects the history of vertical movements. The study of the sedimentary fill
and its associated tectonic subsidence signal gives important information about basin formation
mechanisms (Ter Voorde, 1996; references in Cloetingh et al., 1993). Several palaeo-thermal
indicators, such as fission tracks, vitrinite reflectance and mineral palaeo-thermometers are available to
record thermal history.

1.3 THE EASTERN CORDILLERA OF COLOMBIA

The Eastern Cordillera (EC) of Colombia in an excellent example of a Mesozoic extensional


basin inverted during Cenozoic that can be used as a natural laboratory to study these processes. In the
EC the stratigraphic record is better constrained than other uplift or thermal indicators. For this reason
in the EC and neighbouring areas the stratigraphic record constitutes the natural starting point to study
vertical movements induced by tectonic processes responsible for basin formation, with the aim to infer
from them the thermo-mechanical behaviour of the lithosphere.
Although several conceptual tectonic models have been proposed for the evolution of the EC
(Julivert, 1970; Fabre, 1987; Colletta et al., 1990; Dengo and Covey, 1993; Cooper et al., 1995; Roeder
and Chamberlain, 1995; Casero et al., 1997; Branquet, 1999), the tectonic basin forming and inversion
history of the EC of Colombia has not been studied in terms of the geodynamic processes that govern
deformation of the lithosphere. The deep structure of the EC is largely unknown. Although there is
some geological evidence of basin inversion during the Palaeogene, such inversion is poorly
documented and not regionally understood. Additionally important basic data are not available for the
EC: For example there are no thickness maps for the Meso-Cenozoic sedimentary fill in the area.

1.4 TECTONIC SETTING OF THE EASTERN CORDILLERA

The EC is the eastern branch of the Colombian Andes (Fig. 1.1). The latter comprise three
mountain ranges: the Eastern, Central and Western Cordilleras, which merge southward into a single
range. The Magdalena Valley (MV) separates the Eastern and Central Cordilleras and the Cauca Valley
separates the Central and Western Cordilleras. The EC and its bounding basins, the Llanos Orientales
(LLA) in the east and MV in the west define the area studied in this thesis.
During the Mesozoic the area of the EC was an extensional basin. During the Palaeogene some
authors (e.g.. Van der Hammen, 1961; Roeder and Chamberlain, 1965; Restrepo-Pace et al., 1999a,b;
Gomez et al., 1999) have suggested the possibility of upthrusted blocks and/or incipient inversion of
the Mesozoic extensional basin in the area of the EC. However a commonly accepted view is that of a
unique simple foreland basin related to the topographic load of the Central Cordillera (e.g. Cooper et
al., 1995). This problem will be discussed in Chapter 3. During the Neogene there is a general
agreement that the Mesozoic extensional basin became inverted, deformed and uplifted to form the EC
(Cooper et al., 1995).
In the study area, during Triassic and Jurassic, continental and volcanic facies were deposited
in narrow rifts developed behind a magmatic arc related to the subduction of the Pacific plates under
the western border of South America. During Early Cretaceous marine facies were accumulated on a
wide rift system, probably a back-arc basin related to subduction of Pacific plates west of Colombia
(Aspden et al., 1987). Shallow marine sedimentation ended at the end of the Cretaceous due to the

2
Chapter 1
0 0 0 0
78 76 74 72 0 70 68 0
0
12 a ra
aji
Se ji r
a Gu
n ua n
ea G a si
bb B
a ri

ta
an ev a a

ar
ta d a
N e rr

M
C

ja
S

ey

ri
S

Pe
al l
e
0

d
l le y

rV
10

de
Va

sa

nia
Ce
Be lt

rra
a
le n

Se
Pa
i n to

gd a

M
na
Ja c

a
ma

r M
Venezuela
lt
Be
Sa n
s
de
we
0
An
nu

8 Lo
Si

Bari nas

( M len a
a
id
er B as i n

da
V)
M
S err an ia

Va M ag
ll ey
C)
d le
(E
le y
d e Ba uc

Mid

ra
a
er a

e
Va l

ll er

0 i ll
6 rd
Paciffic Ocean

d i ll

Co
r di
A t ra to

Co r

rn
Co
o

s te
Ea
Llanos O rient ales
J ua n

Basin ( LLA)
L in
eo
l le y

f S
Sa n

e c ti
l le y

0
Va

on
4 Fi g
.
a Va

1.2
a
uc
Ca

nn
l
ra

al e
nt
Ce

gd
Ma
n

St udy ar ea (Fig. 2. 1)
er
s in

st

0
Ba

2 Guyana Shield
pe
We
f ic

Up
ci f
Pa

Putumay o

Ba si n Am a zo na s
0 100 200 Km
0 Ba s i n
0 Ecuador

Figure 1.1: Location Map

accretion of the oceanic terranes of the Western Cordillera (Fig. 1.2, Cooper et al., 1995). According to
Cooper et al., (1995), accretion of the Western Cordillera created the early foreland basin during the
Palaeogene (Fig. 1.2). A different scenario for the Palaeogene is one of incipient inversion of the
Mesozoic extensional basin (Restrepo-Pace et al., 1999a,b; Gomez et al., 1999). During the Neogene
(since middle Miocene) the onset of Andean deformation in the EC isolated the MV basin from the
LLA basin (Fig. 1.2). The deformation was dominated by inversion of the basin controlling faults to
create the EC (Cooper et al., 1995; Fajardo-Pea, 1998).

1.5 THESIS OUTLINE

This thesis gives an overview of the stratigraphy and structural geology of the EC and
summarises from literature its geological evolution. In order to use the stratigraphic record of the EC
and its neighbouring areas to study tectonic vertical movements and basin forming mechanisms, I

3
Introduction Outline
compiled from the available literature more than 100 stratigraphic columns and wells. I also used
previous stratigraphic and palaeogeographic interpretations to create a series of palaeogeographic and
restored original thickness maps for several time intervals, covering the Mesozoic and Cenozoic.
Chapter 2 summarises the Mesozoic stratigraphy and discusses the vertical (tectonic
subsidence) and horizontal (lithosphere stretching) movements that occurred in the Colombian basin
during the Mesozoic. At that time the Colombian basin was an extensional basin. The reconstruction is
accomplished by using the documented sedimentary record as an indicator of subsidence. Because
basin subsidence is produced both by tectonic processes responsible for basin formation, as well as by
isostatic compensation of surface loads produced by sediment and water filling the basin, it is necessary
to separate the tectonic signal from the total subsidence signal. I applied 1D-backstripping techniques
(Bond and Kominz, 1984) in order to do this. Theoretically, if the eustatic variation through time is
known, it is possible to quantify the contribution to subsidence produced by water loads. However,
quantification of the sea level curve is difficult and debatable (e.g. Schlager, 1993). I identified in the
curves of tectonic subsidence several pulses of fast subsidence, and I interpreted each of these in terms
of tectonic or eustatic processes. Only those pulses of fast tectonic subsidence correlatable to normal
fault activity were related to lithosphere extension, while the slower subsidence that followed them was
linked to thermal re-equilibration (cooling and contraction) of the lithosphere after stretching.
Horizontal extensional movements responsible for the observed subsidence were calculated through
forward modelling, assuming that lithosphere stretching produced the Mesozoic subsidence. Using a
lithosphere extension model, which considers a two-layers lithosphere (Royden and Keen, 1980), I
calculated crustal and subcrustal mantle lithosphere factors that best fit the observed tectonic
subsidence signal. The geometry of the basin is reviewed in the context of spatial variations in
extensional factors and the location of faults limiting the extensional basin, which controlled
subsidence. Basins produced by lithosphere stretching usually are associated with magmatism;
therefore, I have also summarized magmatic activity during the Mesozoic and discussed possible
relationships between lithosphere stretching and magmatism in the study area.
Chapter 3 summarises the Palaeogene stratigraphy and studies the contemporary vertical and
horizontal compressional movements that occurred in the Colombian basin during the Palaeogene.
According to some authors (e.g. Cooper et al., 1995; Villamil and Restrepo, 1997) at that time the
Colombian basin was a foreland basin. Other authors (e.g. Van der Hammen, 1961; Roeder and
Chamberlain, 1965; Restrepo-Pace et al., 1999a,b; Gomez et al., 1999) have suggested the possibility
of upthrusted blocks and/or incipient inversion of the Mesozoic extensional basin in the area of the EC.
I discussed and tested these alternative hypotheses using flexural models. I calculated the tectonic
subsidence through backstripping techniques. Several hypotheses about tectonic vertical movements
are tested through 1D-thermal subsidence modelling and 2D-flexural modelling to explain the observed
tectonic subsidence. Because Mesozoic lithosphere stretching generated the basin, the first hypothesis
tested (like tested in the Pyrenees by Desgaulx et al., 1991) was whether some remaining thermal
subsidence induced by Mesozoic stretching affected the Palaeogene evolution. The second hypothesis
tested was one of lithosphere flexure produced by a topographic load in the palaeo-Central Cordillera,
as proposed in literature (e.g. Cooper et al., 1995). The third hypothesis tested was flexural subsidence
produced by some local uplifted blocks within the study area. This hypothesis was based on evidence
of active deformation, slight inversion and uplift in the hanging-walls of Mesozoic extensional faults in
the Colombian basin during Palaeogene. Although the observed subsidence could not be explained by
any individual hypothesis, a combination of all gave a satisfactory result. The observed local flexural
component of subsidence is interpreted as produced by uplift of local faulted blocks. To better
understand the relationships between vertical and horizontal movements, horizontal compressional
movements responsible for uplift of local faulted blocks were calculated through forward modelling of
structural deformation kinematics, sedimentation and erosion, assuming an incipient inversion of the
Mesozoic extensional basin during the Palaeogene (Sassi et al., 1998). This model allowed an estimate
to be made of the amount of basin inversion and shortening that occurred during Palaeogene.
Chapter 4 summarises the Neogene stratigraphy and investigates the vertical movements in the
study area during the Neogene, when the Mesozoic Colombian basin was then inverted, deformed and
uplifted to form the EC. This study was accomplished by studying the basin subsidence of the Llanos
and Magdalena Valley basins using the known sedimentary record as an indicator of subsidence. These

4
Chapter 1
basins became independent sedimentary depocenters in Neogene times. The hypothesis that tectonic
subsidence of these basins resulted from lithosphere flexure produced by the increasing topographic
load of the uplifting EC during Neogene time is tested through modelling. The potential contribution of
Neogene thermal subsidence following Mesozoic stretching was also tested.
Chapter 5 gives an overview of the structural geology of the EC and studies the horizontal
compressional movements that were responsible for basin inversion and uplift of the EC during
Neogene Andean deformation through map view structural restoration of the EC and neighbouring
areas. I manually restored in map view a mosaic of fault bounded blocks using all the kinematic
constraints available in the literature, such as strike-slip displacement estimates and amounts of
shortening from available balanced cross-sections. This technique is useful to constrain estimates of
shortening and strike-slip displacements as well to detect strike-slip motions and rotations about
vertical axes not revealed by balancing cross-sections.
Chapter 6 presents rheological models of the EC lithosphere and based on previous results of
the thesis provides a synthesis of the lithosphere scale Meso-Cenozoic tectonic evolution of the EC in
terms of stresses affecting the lithosphere and lithosphere rheololgy. Additionally Chapter 6 provides a
synthesis of the geological and limited geophysical data, previous results of this thesis and makes
comparisons with similar mountain belts and analogue and numerical models from the literature,
leading to a hypothesis on the deep structure of the EC.

5
W SE
WESTERN 'OCEANIC' ACCRETED TERRANES
EASTERN 'CONTINENTAL' DOMAIN

Pacific Ocean Baudo Arc Atrato Basin Western Cordillera Amaime Central Cordillera Magdalena Eastern Cordillera Llanos Foreland Basin
Terrane Valley Basin
Western Eastern
domain domain
Pliocene-Recent
Miocene
Romeral Suture Palestina F La Esmeralda F Oligocene
Cauca River Fault zone Mulatos F Lengup F
Honda F Eocene
Baud River Baud F Cauca F Cambao F Tesalia F Paleocene
Bituima F Agua Clara F
0 km
Upper
0 km
-10
Cretaceous
A T
-10
-20 . -20
-30 Moho A T -30
Lower
-40 Cretaceous
-40 Evaporites
-50 Moho -50 Jurassic

Triassic

Palaeozoic
PRESENT
Proterozoic

0 km
-10
Dip
-20 Foliation
-30
A T Moho Igneous Intrusion
-40
-50 Salt Intrusion

Back-arc extension

Trench roll back

CRETACEOUS 0 50 100 km

Figure 1.2: Cartoon showing the tectonic evolution of the Eastern Cordillera since Mesozoic time according to the hypothesis
proposed by Cooper et al., 1995. (From Cooper et al., 1995).
CHAPTER 2

MESOZOIC RIFTING HISTORY OF THE EASTERN CORDILLERA, COLOMBIAN ANDES

1. INTRODUCTION

Plate boundary forces, deviatoric tensional stresses developing over zones of upwelling
asthenospheric convection and frictional stresses on the base of the lithosphere caused by the
convective asthenosphere are all driving mechanisms for lithosphere extension (Ziegler, 1994).
Therefore, lithosphere stretching events responsible for development of extensional basins should in
principle correlate in time and space with periods of increased tensional stress resulting from plate-
tectonic interaction (e.g. Janssen et al., 1995). The main proposed geotectonic scenarios for rifting are:
(1) Atlantic-type rifts developed during the break-up of major continental plates (e.g. Ziegler, 1993);
and (2) Back-arc rifts developed during decrease of the convergence rate and/or divergence of colliding
plates (Cross and Pilger, 1982; Tamaki and Honza, 1991). Many features of extensional basins, such as
their associated magmatism, geometry and structural style, are controlled by the rheological structure of
the lithosphere, the presence of pre-existing faults favourable to be tensionally reactivated and the
amount of extension (Ziegler, 1994).
The extensional basins developed during the Mesozoic in the area of the Colombian EC (Fig.
2.1) are good examples to study these processes. However, many features of these extensional basins
and their underlying mechanisms are practically unknown.
SERRANIA MARACAIBO
DE PERIJA BASIN
ES
1400 ND
AS

A
SA

A
L UC

NTA

ID BARINAS

ER
ND

BASIN
SAN

ER

M
DE

MAS
(MVL EN
ANIA

SIF

1300
E RR

SA F LO
A

NT RES
L L GD
S

A N TA
S

DE
L
VA MA
EY

HIL

R
OT

ARAUCA
F L MA S
FO

REGION
HIGH
E

OR SIF
DL

EY

ES
LL

TA

1200
MID

VA

NA
N

LE
IO

DA O
G

G Z IN
RE

MA BLA AS
TA B -B L
IA
LS

CU SU AX
A

IL

ND
TH

SO
LER

IN

IN A
O

AM NJ MO
S
FO

BA

AR TU GA
CASANARE REGION
B-

1100
OS

CA SO
SU
N
RDIL

A
IN

SU
LL

B-
UY
AS

G TA DE

BA
)
OC

LLA
AL S UB -B

SI
N
RE GO A

IN (
BO AN
A

N
L CO

EN

IO

AS
SA

B
DA

ES
GD
N

TAL
HO
MA

1000
TRA

RIEN
SO
NO
CEN

B A OT
IN

LLA
B - RD
S
SU IR A

LEY
G

VAL

SIF

900
AS
NA

META REGION
M
ALE

E
M
TA
AGD

UE

GUYANA
Q
R M

SHIELD
PE
UP

SIF
800 IN M
AS 0 100 200 Km
S N
VA
A ZO
EI B -B AR
R

N U G
E

S
S

800 900 1000 1100 1200 1300 1400 1500

Figure 2.1: Location map of study area with geographic regions referred in the text. Plane coordinates in kilometres
are referred to a local origin at Bogota (X=1.000, Y=1.000), as used by the Colombian Geographic Institute IGAC.
Mesozoic Rifting of the Eastern Cordillera, Colombian Andes
In this chapter I investigate the tectonic basin formation mechanisms of the Mesozoic extensional basin
developed in the area of the EC. I have studied the tectonic subsidence signals that give important
information about basin formation mechanisms. For this purpose, I analysed temporal and spatial basin
subsidence patterns, quantitatively analysed tectonic subsidence and forward modelled it, to explain
them in the framework of geodynamic processes which formed the Mesozoic EC basin. In doing so I
will address questions such as the relationship between basin development, extensional episodes, plate-
tectonic events, magmatic events and basin geometry.
Fabre (1983a,b, 1985b, 1987) and Hbrard (1985) studied the subsidence of the eastern flank of
the EC during the Cretaceous. They identified the basin as being produced by lithosphere extension,
calculated tectonic subsidence curves and, using the uniform instantaneous stretching model developed
by McKenzie (1978), calculated lithosphere stretching factors close to 2. They distinguished an Early
Cretaceous phase of subsidence produced by rifting and Late Cretaceous subsidence produced by
thermal decay after rifting. This chapter studies the tectonic subsidence assuming several events of
lithosphere stretching of finite duration and examines the possibility to differentiate between crustal
and subcrustal stretching, which occurred in the Colombian basin throughout the Mesozoic. An
extensive data set of more than 100 stratigraphic columns and wells from the EC, MV and LLA areas
(Fig. 2.2 and Table 2.1) extracted from literature plus well data from Ecopetrol were used.

2. TECTONIC SETTING

2.1. PLATE-TECTONIC INTERPRETATIONS

2.1.1. Triassic and Jurassic

According to plate-tectonic reconstructions (Pindell and Dewey, 1982; Burke et al., 1984;
Duncan and Hardgraves, 1984; Ross and Scotese, 1988; Jaillard et al., 1990; Pindell and Barret, 1990;
Pindell, 1993, Pindell and Erikson, 1993; Meschede and Frisch, 1998), North and South America were
part of the supercontinent Pangea during Triassic time. For the evolution of this period in Colombia the
plate-tectonic interpretations rely on two hypotheses (Fig. 2.3):
(1) Intra-continental rifting related to the break-up of Pangea (Pindell and Dewey, 1982; Jaillard et al.,
1990) occurred during Triassic and Early Jurassic time (Fig. 2.3a). This hypothesis probably is more
applicable to the northern part of Colombia and Venezuela and to their separation from North America.
(2) Back-arc rifting behind a subduction related magmatic arc (Bourgois et al., 1982; Maze, 1984;
McCourt et al., 1984; Pindell and Erikson 1993; Pindell and Tabut, 1995; Toussaint, 1989, 1995a,b;
Meschede and Frisch, 1998). According to this hypothesis, the study area was located at the margin of
the continent, when active subduction of oceanic Pacific plates was occurring (Fig. 2.3b). Such an
interpretation explains the Triassic and Early Jurassic rift basins in the study area as backarc basins.
The existence of a magmatic arc made up of calc-alkaline plutonic bodies with batholiths of Jurassic
age in the area of the Central Cordillera (Aspden et al., 1987) supports the interpretation of back-arc
rifting. If the nature of the Triassic-Jurassic plutonic bodies of the Santander Massif (Ward et al., 1973;
Restrepo-Pace, 1995) is that of a magmatic arc (Restrepo-Pace, personal communication), then the
Santander Massif can be considered as a minor arc segment compared with the longer Central
Cordillera arc. Intrusion of magmatic bodies in this small Santander arc is favoured by the existence of
the palaeo-Bucaramanga fault-zone (Kammer, 2001). According to Smith and Landis (1995) some arcs
migrate either toward or away from the trench in response to changing subduction angles and other
processes. Because of arc migration, a single site may change among fore-arc, intra-arc and back-arc
settings (Smith and Landis, 1995). Most of the basin may be considered as backarc (relative to the
Central Cordillera arc). In addition the portion of the basin located in the Santander Massif minor arc
segment may be described as intra-arc (descriptive term sensu Smith and Landis, 1995). Similar
difficulties in classifying convergent-margin basins have been discussed by Jordan and Alonso (1987),
Flint et al., (1993) and Smith and Landis (1995).

8
Chapter 2

13
SERRANIA DE MARACAIBO
PERIJA BASIN

S
LUCA
1400

DE
AN
BARINAS

SAN
W

A
CUCUTA BASIN

ID
ER
DE

M
RANIA
SER

Fig. 2.
1300 BUCARAMANGA 0 100 200 Km

24 b
Fi ARAUCA
g.
2 .2
4a

1200 MEDELLIN
E RA
DILL

2
COR

TUNJA

1100
Fig YOPAL
. 2 .2
1
MANIZALES LLANOS ORIENTALES
BASIN (LLA)
F ig
.2
.2 3
BOGOTA Fi a
1000 g.
2 .2
IBAGUE 3b

VILLAVICENCIO

900 5
L
RA
NT
CE

NEIVA

800 GUYANA SHIELD

R
SAN JOSE DEL
E
S

GUAVIARE 12
800 900 1000 1100 1200 1300 1400 1500

Figure 2.2: Location of stratigraphic columns and wells (see Table 2.1) and stratigraphic regional sections. Numbers
along sections refer to labelling of stratigraphic transects (Figs. 2.7, 2.8 and 2.9). Lines E and W denote location of
stratigraphic longitudinal sections (Figs. 2.10 and 2.11)

According to Pindell and Erikson (1993), in addition to continued subduction of Pacific plates
in the western margin of South America, separation of North and South America started during the
Middle Jurassic. As a result a new proto-Caribbean oceanic basin began to open between northwestern
South America and the Chortis and Yucatan blocks (Fig. 2.3b). Calc-alkaline intermediate volcanism
registered in the Middle Jurassic Saldaa Fm in the south of the basin records arc-related magmatism
(Bayona et al., 1994).
For the Cretaceous there are basically three alternative hypotheses (Fig. 2.4):
(1) Backarc rifting (Toussaint and Restrepo, 1989; Bourgois et al., 1982; McCourt et al., 1984; Fabre
1987; Cooper et al., 1995; Meschede and Frisch, 1998). Key evidence for this hypothesis is the
existence of a subduction-related magmatic arc (Fig. 2.4a). However, while, there are Cretaceous
plutonic rocks in the Central Cordillera, Cretaceous plutonism is only sporadically developed in the
northern part of the Central Cordillera, it is practically absent in southern Colombia and Ecuador. It is
very extensive in Peru (Cobbing and Pitcher 1972 in Aspden et al., 1987). Aspden et al. (1987)
suggested that oblique convergence and an offset in the trench along a major dextral transform fault
could explain the notable absence of Cretaceous plutonism in southern Colombia and Ecuador.

9
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes

No. STRAT. COLUMN REFERENCE


1 Abejorral Burgl and Radelli (1962); Botero-Arango (1963)
2 Alpujarra Mojica et al. (1985); Salazar (1992); Florez and Carrillo (1994)
3 Apulo Cceres and Etayo-Serna (1969); Etayo-Serna (1979); Sarmiento (1989); Martnez (1990).
4 Arcabuco Renzoni (1967); Renzoni et al. (1967) ; Galvis and Rubiano (1985)
5 Ataco Cediel et al. (1981); Geyer (1982); Mojica (1982); Mojica and Llinas (1984)
6 Caqueza Hubach (1957); Julivert (1962, 1968) Renzoni (1965a,b, 1968); De La Espriella and Cortes
(1982); Pimpirev et al. (1992)
7 Cocuy Etayo (1985); Fabre(1985a,1986,1987); Mayorga and Vargas (1995)
8 Chima Julivert (1958a,b); Pulido (1979); Pulido et al. (1979a,b, 1980); Rolon and Carrero (1995)
9 Chita Fabre(1985a,1986,1987)
10 Cimitarra Alfonso and Ballesteros (1987); Alfonso (1989)
11 R Luisa Tellez and Navas (1862); Mojica et al. (1985)
12 Coello Tellez and Navas (1862); Mojica et al. (1985)
13 El Sudn Geyer (1976, 1982)
14 Floresta Mojica and Villarroel (1984); Alzate and Bueno (1994); Alzate and Roln (1996)
15 Fusagasug Laverde and Ramrez (1987); Laverde (1989)
16 Girardot Burgl and Dumit (1954); Raasvelt (1956)
17 Melgar Burgl and Dumit (1954), Raasvelt (1956); Diaz (1994a)
18 Guaca Maughan et al. (1979); Fabre (1985a, 1986, 1987)
19 Guataqu Porta (1965); Julivert (1968); Schammel and Buttler (1981, in ESRI - ECOPETROL, 1994)
20 Guateque Burgl (1958); Ulloa et al (1975); Ulloa and Rodrguez (1976a); Mayorga and Vargas (1995)
21 Itaibe Mendivelso (1982)
22 La Calera Van der Hammen (1958); Ujueta (1961); Renzoni (1965a); Julivert (1968); Cuervo and
Ramrez (1985)
23 Labateca Carillo (1982); Boinet et al. (1982,1985)
24 Medios Renzoni (1967); Renzoni et al. (1967); Diaz and Sotelo (1995)
25 Santos Julivert (1958); Zamarreo de Julivert (1963); Julivert et al. (1964); Cediel (1969); Laverde
(1985); Alfonso (1985)
26 Matanza Julivert (1959)
27 Mojicones Daconte and Salinas (1980); Fabre (1981,1987)
28 Nazareth Fabre (1987)
29 Neiva Burgl (1959); Beltran and Gallo (1968); Howe (1969); Wellman (1970); Anderson
(1970,1972); Waddell (1982)
30 Onzaga Daconte and Salinas (1980)
31 Ortega Burgl (1961); Cediel et al. (1981); Mojica et al. (1985); Carrillo and Flrez (1994); Daz
(1994); Amezquita and Montes (1994)
32 Pajarito Miller and Etayo-Serna (1979); Hebrard (1985); Naar and Coral (1993); Cardona and
Gutierrez (1995); Mayorga and Vargas (1995)
33 Aguazul Miller and Etayo-Serna (1979); Hbrard (1985)
34 Paz de Ro Alvarado and Sarmiento (1944); Walthill and Berry (1979); Guatame and Lara (1995)
35 Prado Vsquez and Ros (1979); Mojica and Maca (1981); Salazar (1992); Daz (1994a)
36 Q Calambe Flrez and Carillo (1994)
37 Q El Cobre Barrero (in Julivert, 1968)
38 Q Olini Villamil and Arango (1998); Flrez and Carrillo (1994); Villamil (1994)
39 Quipile Sarmiento et al. (1985); Sarmiento (1989); Martnez (1990); Gmez and Pedraza (1994)
40 R Cusay Ecopetrol(1970); Fabre (1987)
41 R Saldaa Cediel et al. (1981); Geyer (1982); Mojica (1982); Mojica and Llins (1984)
42 San Lucas Bogot and Aluja (1981); Geyer (1982)
43 San Pablo Hall et al. (1972); Etayo-Serna et al. (1976)
44 Servita Julivert (1960); Vargas et al. (1981,b, 1984a); Fabre(1987)
45 San Felix Etayo-Serna (1985); Rodriguez and Rojas (1985)
46 San Luis de Gaceno Ulloa et al. (1975,1976b,c); Ulloa and Rodriguez (1976a); Robertson Reseach Inc. (1983,
1986)
47 Simacota Pulido (1979a,b, 1980); Rolon and Carrero (1995)

Table 2.1: Stratigraphic columns used in this thesis

10
Chapter 2

No. STRAT. COLUMN REFERENCE


48 Simijaca Ulloa and Rodrguez (1979, 1987, 1991)
49 Sogamoso Mayorga and Vargas (1995)
50 Sutamarchan Ulloa and Rodrguez (1979, 1987, 1991)
51 Tabio Burgl (1958); Julivert (1962)
52 Tablazo Morales et al. (1956); Cediel (1968); Ward et al (1973); Renzoni (1985); Clavijo (1985)
53 Lebrija Morales et al. (1956); Cediel (1968); Ward et al (1973); Renzoni (1985); Clavijo (1985)
54 Tibasosa Renzoni (1967); Renzoni et al. (1967); Alzate and Bueno (1994)
55 Tunja Renzoni (1967); Renzoni et al. (1967)
56 Chivat Renzoni (1967); Renzoni et al. (1967)
57 Vado Real Pulido (1979a,b, 1980)
58 Vlez Ulloa and Rodrguez (1979)
59 Villa de Leiva Etayo-Serna (1968), Galvis and Rubiano (1985); Cardozo and Ramrez (1985); Villamil
(1994); Villamil and Arango (1998)
60 Villeta Cardozo and Sarmiento (1987); Sarmiento(1989); Martnez (1990); Moreno (1990)
61 Yacop Moreno and Rubiano (1987); Rubiano (1989); Moreno (1990a); Rodrguez and Ulloa
(1994a,b)
62 Apiay-4P Ecopetrol and Beicip (1995)
63 Bolivar-Corrales Well log; Cspedes and Pea (1995)
64 Cormichoque-1 Well log
65 Chichimene-4 Ecopetrol and Beicip (1995)
66 Chitasug-1 Well log
67 Cusiana-1X-2 Fajardo et al. (1993)
68 Leticia-1 Fajardo et al. (1993)
69 Medina-1 Fajardo et al. (1993)
70 Sabalo-1K Ecopetrol et al. (1994); Ecopetrol-ICP (1996); Ecopetrol, Vicepresidencia de Exploracin,
(1997);
71 Llanito-1 Ecopetrol, Vicepresidencia de Exploracin, (1997)
72 Casabe-199 Ecopetrol, Vicepresidencia de Exploracin, (1997)
73 Infantas-1613 Ecopetrol, Vicepresidencia de Exploracin, (1997)
74 Cascajales-1 Ecopetrol, Vicepresidencia de Exploracin, (1997)
75 Suba-2 Well log
76 Suesca-1 Well log
77 Suesca Norte-1 Well log
78 Tunja-1 Well log
79, 80 Wells Llanos Ecopetrol and Beicip (1995)
81 Arimena-1 Fajardo et al. (1993)
82 Balastrera-1 Ecopetrol and Beicip (1995)
83 Cachama-1 Fajardo et al. (1993)
84 Camungo-1 Ecopetrol and Beicip (1995)
85, 86 Wells Llanos Fajardo et al. (1993)
87 Cao Cumare-1 Ecopetrol and Beicip (1995)
88, 89 Wells Llanos Fajardo et al. (1993)
90 Cao Limn-1 Ecopetrol and Beicip (1995)
91 Cao Rondn-1 Ecopetrol and Beicip (1995)
92 to 102 Wells Llanos Fajardo et al. (1993)
103 Joropo-1 Ecopetrol and Beicip (1995)
104 to 107 Wells Llanos Fajardo et al. (1993)
108 La Tortuga-1 Ecopetrol and Beicip (1995)
109 to 124 Wells Llanos Fajardo et al. (1993)
125 San Juan-1 Ecopetrol and Beicip (1995)
126 ST-0-04 Fajardo et al. (1993)
127, 128 Wells Llanos Ecopetrol and Beicip (1995)
129 to 132 Wells Llanos Fajardo et al. (1993)
133 SV-9 Ecopetrol and Beicip (1995)
134 to 139 Wells Llanos Fajardo et al. (1993)
140 Paime Rubiano (1989)
141 Chenche-1 Ecopetrol, Vicepresidencia de Exploracin, (1997)

Table 2.1: Continued

11
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes

No. STRAT. COLUMN REFERENCE


142 Valle de San Juan Barrero (in Julivert, 1968)
143 Cambao Schamel and Butler (1981 in ESRI ECOPETROL, 1994)
144 Simiti Dickey (1941); Morales et al. (1956); Etayo et al. (1976)
145 Palmichal Ulloa et al. (1975,1976b,c); Ulloa and Rodriguez (1976a); Robertson Reseach Inc.
(1983, 1986); Bartels (1986); Sarmiento (1992, 1993)
146 Tausa Follmi et al. (1992), Sarmiento (1992, 1993)
147 Honda Porta (1965, 1966); Julivert (1968); Van Houten and Travis (1968)
148 Sabana de Bogot Helmens (1990)
149 Chicoral-1 Ecopetrol, Vicepresidencia de Exploracin, (1997)
150 Sapo-1 Ecopetrol, Vicepresidencia de Exploracin, (1997)
151 Surez-1 Ecopetrol, Vicepresidencia de Exploracin, (1997)
152 Venganza-1 Ecopetrol, Vicepresidencia de Exploracin, (1997)
153 Revancha-1 Ecopetrol, Vicepresidencia de Exploracin, (1997)
154 Manueln Howe (1969, 1974); Wellman (1970)
155 Q Micha Howe (1969, 1974); Wellman (1970)
156 Desengao Howe (1969, 1974); Wellman (1970)
157 Palogrande Howe (1969, 1974); Wellman (1970)
158 Campoalegre Howe (1969, 1974); Wellman (1970)
159 Q Guadinosita Howe (1969, 1974); Wellman (1970)
160 to 162 Wells Llanos Fajardo et al. (1993)
Table 2.1: Continued

2.1.2. Cretaceous

(2) Passive margin (Pindell and Erikson 1993; Pindell and Tabut, 1995). The scarcity of magmatic
rocks in the basin seems to support this hypothesis (Fig. 2.4b). However, the presence of some plutonic
rocks of Cretaceous age in the Central Cordillera is difficult to explain with this hypothesis.
(3) Intracontinental rifting related to the break-up of Pangaea. Some authors (e.g. Geotec, 1992)
following this hypothesis suggested a NW-SE graben developed in the northern Central Cordillera
during Early Cretaceous. The presence of some plutonic rocks of Early Cretaceous age in the Central
Cordillera (i.e. San Diego, Altavista, Cambumbia and Mariquita stocks; Restrepo et al., 1991;
Toussaint and Restrepo, 1991) interpreted as part of a subduction-related magmatic arc (Aspden et al.,
1987) is difficult to explain with this hypothesis. Most plate-tectonic assume that during Early
Cretaceous western Pangaea was already broken (Fig. 2.4).
During the latest Cretaceous (post-Santonian) all plate-tectonic interpretations propose a
convergent margin west of Colombia. The Caribbean plate was moving eastward relative to South
America, while the Farallon Plate (Fig. 2.4) was subducting west of Southern Colombia (Pindell and
Erikson, 1993 and Pindell and Tabut, 1995).

3. STRATIGRAPHY

3.1. TRIASSIC AND JURASSIC SYN-RIFT SEDIMENTATION

The Triassic and Jurassic sedimentary record is present in several isolated outcrops (Fig. 2.5).
Continental deposits with red beds and volcanic effusive and pyroclastic deposits are dominant,
although there are locally some marine facies. Triassic and Jurassic rocks were deposited in rift
extensional basins mainly located in the Upper Magdalena Valley, Serrania de San Lucas and the
western flank of the EC (Mojica et al., 1996).
Figure 2.6 shows a stratigraphic synthesis modified from Mojica et al. (1996, their Fig. 2.6).
Triassic and Jurassic sedimentary rocks formed a sequence bounded by unconformities. The lower
contact is marked by an unconformity, which is dominantly angular. The upper contact is dominantly
unconformable but locally it is conformable. Jurassic deposits consisting of clastic facies deposited in
dominantly continental environments are widely distributed. In those sections where there are some
marine facies, they are underlain and overlain by continental clastic facies. The fine-grained muddy
marine facies record local marine incursions. Volcaniclastic, pyroclastic and volcanic lavas are mainly
restricted to the upper part of the Upper Triassic to the lower part of the Middle Jurassic (Mojica et
al.,1996; their Fig. 2.6).

12
an

J
t
A B
uc

N N

Yu
Ch
or
tis
C
ho
rt i
s Y

CC
ca

EC
EC
P

CC
S
S

P
J P

13
Yu G
Chapter 2

A
Ch
or
tis

Ch
or
tis
Y

CC
EC
S
CC
EC

S S
S

R P

1 1
5

Figure 2.3.: Plate-tectonic reconstructions for Triassic-Jurassic time: A. Hypothesis of intracontinental rifting (from Pindell and Dewey, 1980;
Ross and Scotesse, 1988; Jaillard et al., 1990). B. Hypothesis of backarc rifting related to subduction of Pacific oceanic plates west of Colombia
(from Jaillard et al., 1990; Pindell and Erikson 1993; Pindell and Tabut, 1995). EC: Palaeo-Eastern Cordillera extensional basin, CC Palaeo-Central
Cordillera. Inset: location of study area (Fig. 2.1).
E
A B

N
G

C
C

A
nt
F

il l a
CC
EC

s-A
S

CC
ma

EC
S

ime
S

14
Y

C
C
An
till
a
CC

s-
EC

CC
A

EC
m
ai m
e

S
P

5
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes

Figure 2.4. Plate-tectonic reconstructions for Cretaceous time: A. Backarc rifting hypothesis (from Meschede et al., 1998). B. Passive margin
hypothesis (from Pindell and Erikson 1993; Pindell and Tabut, 1995). EC: Palaeo-Eastern Cordillera extensional basin, CC Palaeo-Central
Cordillera. Inset: location of study area (Fig. 2.1).
Chapter 2

Cosinas Gp
Cojoro Gp

Corual Fm

Guatapuri Fm

en
ab
en
h
Los Indios Fm

ug

Gr
ab
La Quinta Fm

Tro

ro
Gr

za
ke

La
s
que

La

n
Sa
al
ch i

ntr
Ma

Ce
Cru day
st)

Morrocoyal Fm
n
tine Bou

s
Serrania Catatumbo
ntal

uca
La Quinta Fm
Conestem

5) de Perija region

San L
1 99

BARINAS
(W

BASIN
in t,

ia de
ous sa

Giron Gp
Serra
e (T

h
ug
n

Tro
e rr a

te
mi T

SECTION

an
1

Urib
a

Triassic-Ju rassic
Tah

Arauquita-1
Matanegra-5
W

and Guafita-5X
wells
ION

Rusia, Montebel
SECT

and Palermo Fms


re
l Su tu

LLANOS ORIENTALES
era

BASIN (LLA)
Rom

Bata Fm

Brechas de Buenavista Fm
E
N
IO
CT
SE

GUYANA SHIELD
Payande Gp
0
-
100 200 Km
Study area (Fig. 2.1)

Outcrops of Triassic and Jurassic strata (Maze, 1984; Toussaint, 1995)

Boundaries of Triassic and Jurassic troughs (Julivert, 1968; Maze, 1984)

TRIASSIC-JURASSIC BASIN COMPARTMENTS


Motema Fm
(Refered in Figure 2.20)

Payande, San Lucas and Sierra Nevada Terranes acording


to Etayo-Serna et al (1986), western part of Chibcha Terrane
(Toussaint, 1995).

Eastern part of Chibcha Terrane (Toussaint, 1995) and


Guyana Shield.

Figure 2.5:Location of Triassic-Jurassic strata outcrops and stratigraphic sections. Labeling of stratigraphic
sections according to Fig. 2.6. The stratigraphic section E represents the Triassic-Jurassic sedimentary record of
the eastern part of the Chibcha Terrane according to Toussaint (1995). The stratigraphic section W represents the
Triassic-Jurassic sedimentary record of the western part of the Chibcha Terrane equivalent to the Payand, San
Lucas and Sierra Nevada Terranes of Etayo-Serna et al. (1986). (modified from Toussaint, 1995). Inset: location
of Figs. 2.1 and 2.2.

15
.
NNE SECTION W (western part of Chibcha Terrane) SSW

Payande San Sierra Guajira


Lucas Nevada

Oca F

Ataco
Q. El Cobre
Coello
San Lucas
El Sudan

Neiva
Rio Saldaa
Age (Ma)

EARLY

Palanz

CRETACEOUS
142.0
LATE Cuiza
Caju
MIDDLE Cheterlo
Uitpana

JURASSIC
EARLY La Mojana Norean
Mbo Prado Guatapuri
Saldaa
205.7 Morrocoyal Los Indios Corual La Quinta
Mbo Chicala El Sudan
LATE
Payande
MIDDLE
Luisa

TRIASSIC
EARLY
248.2
SSW
NNE SECTION E (eastern part of Chibcha Terrane)

16
Arcabuco

Coello
Onzaga

Neiva
Tablazo
Age (Ma)

Cumbre
EARLY Saame
Los Santos Palanz

CRETACEOUS
142.0 Buenavista
Arcabuco
LATE Cz. Guavlo Giron
Jordan Uitpan
MIDDLE
La Rusla
Bocas
Montebel
Motema Tiburon

JURASSIC
EARLY

205.7 Bata Palermo La Quinta


LATE Rancho Grande

MIDDLE

TIRASSIC
EARLY
248.2

Figure 2.6: Triassic-Jurassic stratigraphic sections. Location of transects in Fig. 2.5. Vertical axis represents geological time and horizontal
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes

axis represents present day horizontal distance (km) without palinspastic restoration (modified from Mojica et al., 1996). In difference with
this interpretation Geyer (1982) has suggested that the poorly fossiliferous El Sudn Fm. in the Cinaga de Morrocoyal area (north of
Serrana de San Lucas) is Triassic and correlative with the Luisa Fm. of the Payand area (see data limitations in section 7.3).
Chapter 2
SECTION 1
WNW ES E

W e s t er n East ern Guya na Shield


C hi bc ha Chib cha
Terr ane Terrane

Bituima L a Salin a Gu aicramo


palaeo f ault system palaeo- f au lt Arauquita-1

El S u da n
system Matan egr a-5

T a bl a z o

Sa n t os
and G uafita-5 X
wells
Age (Ma)
CRETACEO US

La
Cu m b r e Ma ca n al
E AR L Y
Magdalena
142.0 palaeo f au lt Los
S ant os
syst em
J U R AS S I C

LAT E
G ir o n

MID D LE Jor d an

E ARL Y L a M o ja n a Bocas
Tiburon
205.7 Morrocoyal
TI R A S S I C

LAT E Co r u a l
0 100 200 Km
MID D LE
E AR L Y
248.2

DEPOSITIONAL ENVIRONMENT

Volcanic and volcaniclastic rocks


Alluvial fan and fluvial sandstones and mudstones
Coastal plain and lacustrine mudstones and sandstones
Littoral to inner shelf sandstones

Shallow marine inner shelf carbonates


Shallow marine inner shelf mudstones and siltstones
Outer shelf shales or carbonates

Turbidites sandstones and mudstones

Figure 2.6: Continued

Facies and thickness similarities related to their geographic position suggest that Triassic-
Jurassic sedimentation occurred in two separate basin compartments, each one with its own subsidence
history and sedimentary fill (Figs. 2.5 and 2.6):
(1) Upper Magdalena and Cienaga de Morrocoyal (region A in Fig. 2.5). It corresponds with the
western part of the Chibcha Terrane as defined by Toussaint (1995a,b) or the Payande and San
Lucas Terranes proposed by Etayo-Serna et al., 1983; Fig. 2.5). In this area a marine incursion
with a correlatable maximum marine flooding surface is recognisable near the Triassic-Jurassic
boundary (Chicala member of the Saldaa Fm in the UMV, the Morrocoyal and Los Indios Fms,
ages according to Mojica et al., 1996; their Fig. 2.6). Continental sedimentation followed by shallow
marine limestone deposition occurred during Triassic (Norian?-Rhetian) in the Upper MV.
Volcanic-related facies are volumetrically more important in this area than in the Eastern
Cordillera area (Region B in Fig. 2.5).
(2) Eastern Cordillera (Region B in Fig. 2.5). This corresponds to the eastern side of the Chibcha
Terrane as defined by Toussaint (1995a,b; Fig. 2.5). In this area a marine incursion with a
correlatable maximum marine flooding surface is located within the Lower Jurassic (upper part of
Bata and Montebel Fms, ages according to Mojica et al., 1996; their Fig. 2.6).

17
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes

3.2. CRETACEOUS SEDIMENTATION

The majority of exposed rocks in the EC are Cretaceous in age. Figures 2.7 to 2.11 show two
longitudinal and 3 traverse time-stratigraphic cross-sections of the basin. Location of these sections is
shown in Figure 2.2. They have been constructed from earlier versions by Fabre (1985a, 1986, 1987)
and Cooper et al (1995), modified according to sequence stratigraphy interpretations from literature
(Pimpirev et al., 1992; Fajardo et al., 1993; Etayo-Serna, 1994; Ecopetrol et al., 1994; Villamil, 1994;
Roln and Carrero, 1995; Villamil and Arango, 1998).
Cretaceous rocks, including locally the uppermost Jurassic and Paleocene deposits, form a
mega-sequence bounded by regional unconformities that are at least locally angular. On a broad scale,
Cretaceous rocks represent a major transgressive-regressive cycle with the maximum flooding surface
close to the Cenomanian-Turonian boundary, corresponding to the maximum Cretaceous, and even
Mesozoic, eustatic level (Fabre, 1985a; Villamil 1993; Figs. 2.7 to 2.11). Superimposed on this large-
scale trend, several smaller transgressive-regressive cycles are present, suggesting an oscillating
relative tectono-eustatic level. Subsidence was rapid (Fabre, 1983a,b, 1985b, 1987; Hbrard, 1985), but
shallow water sedimentation suggests that deposition kept pace with it. The basin was a wide rift
system oriented approximately NNE-SSW, divided into two sub-basins (western Tablazo sub-basin and
eastern Cocuy sub-basin) by the Santander-Floresta palaeo-Massif. Toward the north, these sub-basins
continued to the Machiques Trough in the Mrida Andes of Venezuela and the Uribante Trough in the
Serrana de Perij (Julivert, 1968; Fabre, 1985a, 1987). Toward the south these sub-basins joined as the
Cundinamarca sub-basin (Burgl, 1961), where the thickness of the Cretaceous sections reaches a
maximum (Figs 2.1, 2.12 and 2.13). Fabre (1987) and Sarmiento (1989) suggested that the
Cundinamarca Trough was limited to the south by a NW-SE transfer palaeo-fault. South of it, there was
an extensional relay system (Restrepo-Pace, personal communication). N-S lateral changes of thickness
suggest the existence of other NW-SE trending transfer faults (Fig. 2.12 and 2.13). Based on the
presence of Lower Cretaceous sedimentary rocks in the northern part of the Central Cordillera (Figs.
2.12, 2.14 and 2.16) Geotec (1992) suggested the existence of a NW-SE graben in this area, it was
connected with the Cundinamarca sub-basin. Compiled available thickness of these sediments of the
Central Cordillera for similar chronostratigraphic intervals, however, is significantly smaller than those
of the EC (Figs. 2.13, 2.15 and 2.17). If such a graben existed, in terms of subsidence it was a minor
feature compared with the grabens in the area of the EC.

3.2.1. Early Cretaceous syn-rift sedimentation

The Early Cretaceous sedimentary history is illustrated in Figures 2.12 to 2.17. Sedimentation
started in the Tablazo sub-basin in Jurassic time and continued during Early Cretaceous locally without
a tectonic-related angular unconformity (e.g. at the Rio Lebrija section, Cediel, 1968). In other areas
Cretaceous sedimentary rocks rest with an angular unconformity on earlier Mesozoic, Palaeozoic or
even Pre-Cambrian rocks. In the Tablazo sub-basin the first facies deposited were mainly sandstones
(Los Santos, Tambor, and Arcabuco Fms) deposited in fluvial environments (Renzoni, 1985a,b, c;
Clavijo, 1985; Vargas et al., 1985; Laverde and Clavijo, 1985; Galvis and Rubiano, 1985; Etayo-Serna
and Rodrguez, 1985). Brgl (1960, 1964, 1967) suggested that an initial marine incursion in the
Cundinamarca sub-basin flooded a continental area with a desert climate, which provided conditions
for evaporite formation during the early stages of marine transgression. McLaughlin (1972) cited
palaeontological evidence of Berriasian-Valanginian age for some evaporite occurrences. During the
Berriasian the sea flooded the basin from the northern part of the Central Cordillera toward the
Cundinamarca sub-basin (Etayo-Serna et al., 1976). Then the sea advanced from the Cundinamarca
sub-basin toward the north into the two sub-basins while the Santander-Floresta palaeo-Massif
remained emergent (Etayo-Serna et al., 1976; Fabre, 1985a, 1987; Sarmiento, 1989; Cooper et al.,
1995; Figs. 2.12 to 2.17).

18
WNW SECTION 2 ESE
MIDDLE MAGDALENA VALLEY EASTERN CORDILLERA LLANOS ORIENTALES

Guaicramo Fault
La Salina Fault

Simit
Gu aca
C. Rondon-1
St ella 1

Joropo-1

R Servita
Cocu y
R. Ele-1
ST- CN- 8

In fantas 161 3
R Cu say

Mojicones

Ta bl azo
Los Sant os
Age (Ma)
0 Q 0
PLI L
Mesa Guayabo
10 10
M Real Leon
20 MIO E Carbonera 20
Colorado Concentracion
L
30 OLI E Mugrosa Mirador 30
L Picacho
40 Esmeraldas
M La Paz 40

TERTIARY
EOC
50 E 50
6 L Lisama Ar Socha Arc Socha Barco Cu ervos
PAL E 60
0 Guaduas
70 M
Umir Tierna 70
C Colon MitoJuan Labor Guadalupe
80 80

LATE
S La Luna La Luna Gacheta
90 C
T

19
90
C Chipaque
100 L Simiti 100
A M Une
E
Chapter 2

110 Tablazo

CRETACEOUS
Apon DEPOSITIONAL ENVIRONMENT 110
A
120 Paja Tibu-Mercedes 120
B

EARLY
130 H
Las Juntas Alluvial fan and fluvial sandstones and mudstones
Rosablanca Rio Negro 130
V
14 0 R B Macanal
Los Santos Coastal plain and lacustrine mudstones and sandstones 140
V T Giron Guaicramo
150 K Boyac Servit 150

LATE
O Palaeo-fault Palaeo-fault Littoral to inner shelf sandstones
160 Palaeo-fault
C 160
B
170 Shallow marine inner shelf carbonates

JURASSIC
B La Salina COCUY SUB-BASIN 170
Suarez

MIDDLE
180 A Palaeo-fault Palaeo-fault 180
Shallow marine inner shelf mudstones and siltstones
MAGDALENA-TABLAZO
SUB-BASIN Outer shelf shales or carbonates
0 100 200
Km
Turbidites sandstones and mudstones

Figure 2.7: Cretaceous and Tertiary stratigraphic section 2. Location in Figure 2.2. Vertical axis represents geological time according to the scale
of Gradstein and Ogg (1996), horizontal axis represents present day horizontal distance (km) without palinspastic restoration (after Fabre, 1985,
1987; Villamil 1994; Cooper et al., 1995)
SECTION 5
WNW ESE
MIDDLE MAGDALENA VALLEY EASTERN CORDILLERA LLANOS ORIENTALES
Guaicaramo Fault
La Salina Fault

Age (Ma)

Sutamarchan

Matadero-2
Yacopi
Velez
Sim ijaca
Moniquira
Tunja-1
Cormchoque-1
Tu nja
Pajarito
Cusiana-1
La Mara 1
Gloria N-1
Entrerrios-1
Guarimena-1
Arimen a-1
C. Barulia-1
Planas-1

Villa de Leiva
ST-GU-15

0 0
PLI Q Mesa Tilata Guayabo
10 L Real 10
M Leon
20 MIO E 20
L Concentracion
30 OLI Colorado Carbonera
E 30
L Mugrosa Picacho Mirador
40 40
Esmeraldas

TERTIARY
EOC M
50 E 50
60 L Bogota Arc Socha Ar Socha Cuervos Barco
PAL E 60
Lisama Guaduas
70 M Tierna 70
C Umir Labor Guadalupe
80 80

LATE
S Plaeners Dura Gacheta
C La Luna
90 T San Rafael Chipaque 90
C
100 L Hilo Simiti Une 100
San Gil Sup DEPOSITIONAL ENVIRONMENT
A M
110 E Tablazo San Gil Inf 110

CRETACEOUS
A Fomeque Alluvial fan and fluvial sandstones and mudstones
120 Paja 120
B

EARLY

20
130 H Tibasosa Las Juntas 130
V
Rosablanca Coastal plain and lacustrine mudstones and sandstones
140 R B Arcabuco Macana 140
l Littoral to inner shelf sandstones
150 V T Giron
K 150
Boyac

LATE
O
Soapga
160
C
Palaeo-fault Palaeo-fault(?) Guaicramo Shallow marine inner shelf carbonates 160
B La Salina Suarez Palaeo-fault
170 170

JURASSIC
B Palaeo-fault Palaeo-fault

MIDDLE
A
Shallow marine inner shelf mudstones and siltstones
180 180
MAGDALENA-TABLAZO COCUY SUB-BASIN
Outer shelf shales or carbonates
SUB-BASIN
0 100 200 Km Turbidites sandstones and mudstones

Figure 2.8: Cretaceous and Tertiary stratigraphic section 5. Location in Figure 2.2. Vertical axis represents geological time according
to the scale of Gradstein and Ogg (1996), horizontal axis represents present day horizontal distance (km) without palinspastic
restoration (after Fabre, 1985, 1987; Villamil 1994; Cooper et al., 1995).
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes
WNW SECTION 12 ESE

MIDDLE MAG VALLEY EASTERN CORDILLERA LLANOS ORIENTALES

Bituima Fault Guaicramo Fault

Villeta

Viani
Tabio
Manacaias-1

Beltran -1
Suba-2
La Calera
Chichimene-1
Negritos-1
Voragine-1
Ganibay-1

Caqueza
Apiay-1
SV-9

Chitasuga-2
Age (Ma)
0
0 Mesa
PLI Q Guayabo 10
10 L Honda Marichuela Sabana Subachoque Leon
M
MIO Santa Teresa 20
20 E La Carbonera
L San Juan de Usme 30
30 OLI Mirador
E La Regadera
L Rio seco 40
40

TERTIARY
EOC M
50
50 E DEPOSITIONAL ENVIRONMENT
Hoyn Bogota Ar. Cacho 60
60 L
PAL E Seca Guaduas
Tierna Alluvial fan and fluvial sandstones and mudstones 70
70 M
Labor 80
80 C Gacheta Coastal plain and lacustrine mudstones and sandstones

21
Dura

LATE
S 90
90 C
T Chipaque
C Frontera Littoral to inner shelf sandstones 100
10 0
Chapter 2

A L Hilo
M Une 110
110 E Socot Shallow marine inner shelf carbonates

CRETACEOUS
A 120
12 0 Trinch eras Fomeque
B 0 100 200 Km

EARLY
Utica Shallow marine inner shelf mudstones and siltstones 130
13 0 H Cambao Caqueza
V
Palaeo-fault Macanal 140
14 0 R B Murca Buenavista Outer shelf shales or carbonates
V 150
150 T
K Turbidites sandstones and mudstones

LATE
O 160
16 0 C Bituima Guaicramo
B 170
17 0 Palaeo-fault Canoas Palaeo-fault

JURASSIC
B

MIDDLE
A Palaeo-fault 180
18 0

Figure 2.9: Cretaceous and Tertiary stratigraphic section 1.2 Location in Fig. 2.2 vertical axis represents geological time according to
the scale of Gradstein and Ogg (1996), horizontal axis represents present day horizontal distance (km) without palinspastic restoration
(after Fabre, 1985,1987; Villamil 1994; Cooper et al, 1985))
SEC TION E
SSW NNE
U PPER M AG D ALEN A VA LLEY SOU TH ER N EASTER N C OR D EASTER N C OR D ILL ER A

N azar eth
Faul t
C oc uy
R C u sa y
La b a te c a

M esa

N ei va
Alp uj ar ra
Gir ar do t
C a q ue za
Gu a t eq u e
Paj ar it o

N a za re th
Ag e (M a) Gig ant e
0 Q H on da
PL I
10 L H on da 0
M IO M Barzal oza 10
20
OLI L Gua lan day 20
30
L 30
40 EOC M
50 40
E
PA L L Gua du 50
60 E
70 MAA 60
CAM 70
80
L SAN 80
C ON
90 T UR
CEN 90
100 L
ALB M
E 100
110
APT
120
E 110
BAR

22
HAU 120
130 VAL N azar eth Palae o-
140 R B 130
V
tr ansfe r fa ult
L T 140
150 K
O Pa laeo - Pa laeo - 150
160
C tr ansfe r fa ult
B tr ansfe r fa ult C OC U Y SU B-BA SIN 160
170 M
B SOU TH ER N
180 A C U N D IN AM AR C A 170
C U N D IN AM AR C A
SU B-BASI N 180
SU B-BASI N

Alluv ial fan and fluv ial sand ston es a nd m udst ones

C oast al p lain and lac ustr ine muds tone s an d san d-


sto nes
0 100 200 Km Litto ral to i nner shel f sa ndst ones

Shallo w mar ine i nner shelf carb onate s

Shall ow ma rine inner shel f mud stone s and silt stone s

Outer she lf shale s or car bonat es


Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes

Turb idit es s ands tone s a

Figure 2.10: Cretaceous and Tertiary stratigraphic section E. Location in Fig. 2.2. Vertical axis represents geological time according to the scale of Gradstein
and Ogg (1996), horizontal axis represents present day horizontal distance (km) without palinspastic restoration (after Fabre, 1985, 1987; Villamil 1994;
Cooper et al., 1995)
SSW SECTION W NNE

UPPER MAGDALENA VALLEY SOUTHERN EASTERN CORDILLERA EASTERN CORDILLERA -WESTERN FOOTHILLS

Nazareth
Palaeo-fault

Neiva
Q. Olini
Ortega
Ap ulo
Villeta
Yacop
Cimitarra
Cascajales 1
Infantas-1613
Tablazo

Florentina-1
Girardot
Matadero-2
Age (Ma) Gigante Mesa
0
0
PLI Q
Honda 10
10 L
MIO M Barzaloza Barzaloza Santa Teresa 20
20 E
Colorado
L Mugrosa 30
30 OLI E
Gualanday San Juan de Rioseco
Esmeraldas-La Paz 40
L
40

TERTIARY
EOC M 50
50 E Hoyon 60
L
Guaduala Seca Lisama
60 PAL Cimarrona
E 70
70 M Monserrate Lutitas y Arenas Umir
Lutitas y Arenas 80
C Olini
80 90
S Frontera Frontera La Luna
L C
90 T Hilo
C Simiti 100
10 0 Caballos Tablazo
A ML Alpujarra Ocal Hilo 110
110 E Yavi

CRETACEOUS
A
Trincheras Paja 120
Naveta
12 0

23
E B Rosablanca 130
DEPOSITIONAL ENVIRONMENT Socot Utica Rosablanca
13 0 H Murca
V 140
140 R B Cumbre
Chapter 2

Alluvial fan and fluvial sandstones and mudstones Arcabuco Los Santos-Tambor 150
150 L V K T Nazareth Palaeo Palaeo Palaeo 160
16 0 O Coastal plain and lacustrine mudstones and sandstones transfer fault transfer fault
C transfer fault 170
B
17 0 M MAGDALENA-TABLAZO

JURASSIC
B Littoral to inner shelf sandstones SOUTHERN CUNDINAMARCA 180
A
18 0 CUNDINAMARCA SUB-BASIN SUB-BASIN
Shallow marine inner shelf carbonates
SUB-BASIN
Shallow marine inner shelf mudstones and siltstones 0 100 200 Km

Outer shelf shales or carbonates


Turbidites sandstones and mudstones

Figure:. 2.11. Cretaceous and Tertiary stratigraphic section W. Location in Fig 2.2. Vertical axis represents geological time according to the
scale of Gradstein and Ogg (1996), horizontal axis represents present day horizontal distance (km) without palinspastic restoration (after
Fabre, 1985, 1987; Villamil 1994; Cooper et al., 1995).
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes
Early Cretaceous sedimentation on the Tablazo sub-basin. In the Tablazo sub-basin latest
Jurassic to Valanginian fluvial sedimentation was followed by mudstone deposition in marginal marine
environments recording the marine transgression (Cumbre Fm, Mendoza, 1985; Renzoni, 1985c;
Ritoque Fm Ballesteros and Nivia, 1985; Roln and Carrero, 1995). Later, tidal and shallow water
marine shelf carbonates (Rosablanca Fm, Cardozo and Ramirez, 1985) were deposited during
Valanginian-Hauterivian time, followed by shallow marine shales (Paja Fm) during Hauterivian-
Barremian time (c.f. Etayo-Serna et al., 1976). Although transgression was progressive from the center
of the basin, two periods of relative sea retreat occurred, during Hauterivian and Aptian (Roln and
Carero, 1995; Ecopetrol et al., 1994; Figs. 2.12 to 2.17).
Later during Aptian a relative tectono-eustatic sea-level rise occurred as suggested by a deeper
marine facies of the upper part of the Paja Fm (Forero and Sarmiento, 1985; Ecopetrol et al., 1994;
Roln and Carrero, 1995).
Berriasian to Aptian sedimentation on the Cocuy sub-basin. In the Cocuy sub-basin marine
transgression started in its southern part during the end Jurassic to earliest Cretaceous as recorded by
the Brechas de Buenavista Fm (Dorado, 1984) and the Calizas del Guavio Fm (Ulloa and Rodrguez,
1976; Fabre 1985a; Mojica et al., 1996). In the northern part of the Cocuy sub-basin, facies changes
record the transition from continental to shallow marine sedimentation (Lutitas de Macanal Fm) during
Berriasian to Valanginian times (Fabre, 1985a). During the Hauterivian to Barremian wave-dominated
deltaic sandy environments of deposition developed (Arenisca de Las Juntas Fm, Fabre 1985a; Figs.
2.12 to 2.17). In Hauterivian time, the deposition of prograding sands in a rapidly subsiding basin
(Fabre, 1985a) was probably facilitated by relative tectono-eustatic base level fall.
Early Cretaceous sedimentation on the Santander-Floresta palaeo-Massif. The Santander-
Floresta palaeo-Massif remained emergent until Hauterivian time when deposition of continental
sandstones began, followed by progradation of deltaic sandstones (Rionegro Fm, lower part of Tibasosa
Fm) in turn by shallow water marine carbonates. The two sub-basins started to form a single wide basin
during Hauterivian time due to flooding of the palaeo-Massif (Fabre, 1985a; Moreno, 1990a,b, 1991)
and base level rise. This intrabasinal high was, however, a significant barrier to sediment movement
until Aptian time (Cooper et al., 1995; Figs. 2.12 to 2.17). The succession of sandstone (Tambor and
Los Santos Fms), limestone (Rosablanca Fm) and dark shale (Paja Fm) facies, recorded in the Tablazo
sub-basin, is laterally younger toward the east on the Santander-Floresta palaeo-Massif (sandstone:
Rionegro Fm, limestone and shale: Tib and Mercedes Fms) and in the Cocuy sub-basin (sandstone:
Arenisca de Las Juntas Fm, limestone and shale: Apn Fm; Fabre, 1985a). This lateral change in age of
facies occurred as a result of the oscillating and progressive marine transgression toward the east during
Valanginian to Aptian times (Figs. 2.12 to 2.17).
Berriasian to Aptian sedimentation on the Cundinamarca sub-basin. Towards the south both
the Tablazo and Cocuy sub-basins show a gradual increase in dark shale deposited in poorly
oxygenated shallow marine environments. (Caqueza Gp, Villeta Gp; Fabre, 1985a; Rubiano, 1989;
Sarmiento, 1989). In the Cundinamarca sub-basin Cretaceous sedimentation started during Tithonian?-
Berriasian-Valanginian with turbidite deposits in both the eastern (lower Caqueza Gp, Pimpirev et al.,
1992) and western (lower part of Utica Sandstone, Murca Fm, Sarmiento, 1989; Moreno, 1990b, 1991)
flanks (Figs. 2.12 to 2.17). Turbidite deposition prevailed up to the Hauterivian in the eastern border of
the basin (Caqueza Gp, Pimpirev et al., 1992).
During the earliest Cretaceous, basin subsidence exceeded sediment supply, resulting in
retrogradation of the turbidite fan system, so that distal fan sediments covered middle fan mouth
channel deposits. In post-Berriasian time, sediment supply increased overwhelmed basin subsidence,
resulting in progradation of the turbidite fan system (Pimpirev et al., 1992) and locally by progradation
of deltaic sands during Hauterivian time (upper part of Utica Sandstone, Sarmiento 1989; Moreno,
1990b). Towards the south the shallow marine sandstones and limestones of the Naveta Fm mark the
development of a shoreline during Hauterivian-Barremian time (Cceres and Etayo-Serna, 1969;
Sarmiento, 1989). Differential subsidence related to syn-sedimentary normal faulting caused unstable
slopes on basin margins. These processes favoured turbidite deposition during early Cretaceous up to
the Aptian (lower part of Utica Sandstone, Murca Fm, Socota Fm, Polana and Rodrguez, 1978;
Sarmiento, 1989; Moreno, 1990b, 1991; Caqueza Gp, Pimpirev et al., 1992; Figs. 2.12 to 2.17).

24
Chapter 2
Serrania
de Perija Maracaibo
Basin

ucas
1400 Cucuta
Barinas

s
de
an L
Basin

An
de S

Floresta High

a
rid
Santander

Me
ania
1300 Bucaramanga

Serr

lt
fau
Arauca

o
lae
a
ra

pa
sub am ar c
di lle

as alen rez
sin
b- b gd Sua
- ba
e

Co r

din

asin
1200

lt
Cun

fau
em

b -b
Medellin

in
Su -Ma

o
lae
t syst

Su
pa
zo

ca
bla
eo faul

cuy
ya
Bo
Ta Llanos Orientales

Co
Tra
a pala

ns
ter Tunja Basin (LLA)

em
1100 pa

st
lae

sy
o
Bituim

fau Yopal

t
ul
lt

fa
o-
le
pa
Manizales
BERRIASIAN

o
m
VALANGINIAN
ra
ica
ua
Na
G

za Bogota PALEOGEOGRAPHY
1000 re
th
pa Positive relief (absent).
Ibague lae
o- Alluvial fan and fluvial
fau
lt sandstones and mudstones
Villavicencio Coastal plain predominantly
al

sandstones.
nt r

Coastal plain predominantly


Ce

900 mudstones.
Littoral to inner shelf
- sandstones.
0 100 200 Km Shallow marine inner shelf
carbonates.
Shallow marine inner shelf
mudstones and siltstones.
La Macarena

Neiva
Serrania de

Outer shelf shales or


800 carbonates
Turbiditic sandstones.e
San Jose del Guaviare

800 900 1000 1100 1200 1300 1400 1500

Figure 2.12: Berriasian-Valanginian palaeogeography without palinspastic restoration (modified from Etayo Serna
et al, 1968, Geotec, 1992; Cooper et al., 1995).

Aptian sedimentation on the palaeo-Upper Magdalena Valley (UVM). Cretaceous


sedimentation started during the Aptian (Vergara and Prssl, 1994) in an extensional basin formed
initially in Jurassic time. Feldspathic and lithic sandstones, conglomerates and red mudstones (Yav
Fm) were deposited as alluvial fans on valley slopes, while finer sandstones and mudstones (Alpujarra
Fm) accumulated within a fluvial system flowing northward. (Florez and Carillo, 1994; Etayo-Serna,
1994; Figs. 2.16 and 2.17).
In the whole EC basin, abrupt lateral thickness changes and ubiquitous turbidite deposition
attest for a local tectonic/differential subsidence depositional conditions in Early Cretaceous time (Figs.
2.13 to 2.17). Regional correlation of Early Cretaceous relative tectono-eustatic cycles is difficult to
establish due to local active extensional tectonics. Since Aptian time these relative tectono-eustatic
level cycles become more tractable or traceable (Figs. 2.7 to 2.11).
An important transgression followed a relative sea-level rise during late Aptian time. During
this time the sea flooded all the area of the present EC even south of the Cundinamarca sub-basin.
(Etayo-Serna et al., 1976; Etayo-Serna, 1994). During late Aptian time the sea gradually flooded the
UMV and dark mudstone and limestone (El Ocal Fm) were deposited in a shallow marine environment
(Etayo-Serna, 1994; Figs. 2.16 and 2.17). Dark grey to black mudstone was deposited regionally in a
dysoxic shallow marine shelf (upper part Paja Fm, in the former Tablazo sub-basin, Fmeque Fm, in

25
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes
S e r ra n ia M ara c aibo
de B a sin

s
de
P e r ija

An
0 0
1400

i da
Cucuta
Barinas

ca s

er
0 0

M
n Lu
Basin

e Sa
nia d
0
530
1300 74 602 Bucaramanga

Se rra
112

tl
Arauca

f au
155
258 0
123 0

eo
0 900

ala
226 258 0

zp
1000 0
ra

a re
Cord ille

1000

Su

S
698
Medellin 1156
1200

lt
798 0

f au
1000
te m

eo
0

la
300
lt sys

pa
692
1000 0

a
10 0

ac
200 757 440
o fau

y
20

Bo
600 17 0
palae

em
1210Tra Tunja

st
ns 1500 Lllanos Orientales
1100 190

sy
fe 20 0 0
0 rp
Basin (LLA)

ult
a

al 1500
ae

fa
Bituim

o Yopal 0

o
fa

e
ul 0

la
t 00

pa
2135
275 0 0 0 0

o
Manizales 0

am
3335
0 0
0
0 aic 0
0
0
0
Gu

0
Bogota 0 0
0 0
1000 Tr
an
0
0
0
Ibague sf 1110
0 er
p al 0
0 ae 0
o 0 0
0 of 0
fa 0 0
ul 0
t Villavicencio 0 0
0 0
0 0
tr al

0 0 0 0
C en

0
0 0
900
0 0

0
0 BERRIASIAN
0 100 200 Km
0 VALANGINIAN
0
Neiva

0
800 San Jose del Guaviare THICKNESS
0
(meters)
S

800 900 1000 1100 1200 1300 1400 1500

Figure 2.13: Berriasian-Valanginian thickness (meters) without palinspastic restoration. Thick lines represent alaeo-
faults believed to be active during Berriasian-Valanginian time.

the former Cocuy sub-basin, Villeta Gp, upper part Socot Fm in the former Cundinamarca sub-basin,
El Ocal Fm in UMV; Figs. 2.16 and 2.17).

3.2.2. Cretaceous post-rift sedimentation

Cretaceous post-rift sedimentation is illustrated in Figures 2.7 to 2.11. Villamil (1993)


interpreted limestone-shale or chert-shale rhythmic beds as Milankovitch cycles. Using graphical high-
resolution stratigraphic correlation he showed that these distal pelagic limestone-shale cycles are coeval
to proximal parasequences. Assuming these cycles have all the same duration, and that subsidence was
constant through time, Villamil (1993) plotted the thickness of all cycles in a modified Fisher plot (a
stacking plot for cyclic rhythmic sedimentation) to obtain a curve of changes in relative
accommodation space or relative tectono-eustatic base level.
Based on facial analysis, macrofossil biostratigraphy, high-resolution event and cycle
chronostratigraphy, together with the modified Fisher plots, Villamil (1993) proposed a sequence
stratigraphic interpretation and a relative tectono-eustatic level history.

26
Chapter 2
Maracaibo
Basin s

Uribante
ue

Trough
iq h

cas
h
1400 ac g
M r ou

San Lu
T Barinas
Cucuta
Basin

ia de
Bucaramanga

Serran
1300

ra

ult
di lle

-f a
Arauca

a
eo

n
Co r

e
ala

a s al
zp

-B gd

a
t
in

vi
bla Suare

Ma

r
in
Se

as
1200

m
ult
m

zo
Medellin

-f a

ste
te

b-B
Su

o
sys

lae

sy
Ta

pa

Su
fault

lt
ca

au
ya

f
Llanos Orientales

Bo
o

cu

o
alae

lae
Tr Basin (LLA)

Co
a
p

pa
1100 Cu nsfe Tunja
ana

nd r Yopal
p
Bit u

Su ina alae

o
m
b- m a o

a
Ba r c fau

ar
sin a lt HAUTERIVIAN

aic
Manizales

Na Gu BARREMIAN
za
1000 Ibague
re
th Bogota PALEOGEOGRAPHY
pa
lae
o
fau Positive relief (absent)
lt
Villavicencio Alluvial fan and fluvial
sandstones and mudstones
lley
al

Coastal plain predominantly


ntr

Va

san d s t on es
900
Ce

Coastal plain predominantly


na

mu d ston es
ale

0 100 200 Km Litoral to inner shelf


gd

san d s t on es
Ma

Shallow marine inner shelf


Neiva carb on at es
r

Shallow marine inner shelf


pe

La Macarena

mudstones and siltstones


Serrania de
Up

800 San Jose del Guaviare Outer shelf shales or


carb on at es
Turbiditic sandstones

800 900 1000 1100 1200 1300 1400 1500

Figure 2.14: Hauterivian-Barremian palaeogeography without palinspastic restoration (modified from Etayo Serna et
al, 1968, Geotec, 1992 ; Cooper et al., 1995).

During Albian a relative base level fall favoured progradation of deltaic and littoral sands
(Caballos Fm, Florez and Carrillo, 1994; Etayo-Serna, 1994) in the area of the UMV and the eastern
border of the basin (lower part of Une Fm, Fabre, 1985a).
During middle to late Albian transition from the near-shore marine facies of the Caballos Fm to
the deepening upward lower part of Villeta Gp in the Upper MV recorded a rise in relative tectono-
eustatic level (Villamil, 1993; Etayo-Serna, 1994). This tectono-eustatic level rise also was recorded by
the upward deepening trend from the shallow water San Gil Inferior Fm to the deeper San Gil Superior
Fm, the Socot Fm to the Hilo Fm and within the Une Fm (Villamil, 1993; Figs. 2.7 to 2.11).
During late Albian-early Cenomanian a relative tectono-eustatic level fall was recorded by
progradation of the upper part of Une Fm and a generalized shallowing upward facies trend. In the
earliest Cenomanian there is a sequence boundary expressed as a forced regression (unnamed shale
overlying the cherts of Hil Fm, shallow water sandstone of Churuvita Fm over deeper shale of San Gil
Superior Fm, Villamil, 1993). In the Upper Cenomanian, Villamil (1993) interpreted the next marked
sequence boundary (first sandstone in the shales of Villeta Gp, sandstone upper part of Churuvita Fm,
uppermost sandstone Une Fm).

27
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes

Serrania
de Perija Maracaibo
Basin

ucas

s
de
1400

An
1 00
Barinas

an L
Cucuta
200 0

a
id
50 Basin

de S

er
M
50 0 0
0

10
10

0
Bucaramanga 2000

ania

20 0
0
130 20

0
460

10
Serr

200
1300

50 0
260 187

0
476
304
0
284 283 Arauca
225

20
222 581

lt

00
au
393 1200

o f
ra 306 0

lae
2122
100

d ille

lt
pa

fau
200

0
775 0

50
z
627
1200
Cor

are

o
lae
Medellin 612 1300
Su

pa
579 312
Llanos Orientales

ca

0
782

200
ya
1210 0
315 Basin (LLA)

Bo
613
0

283 265
50

160
650 227
650 Tra Tunja em
st
ns 1040 0
100

1100 11 fer
500

0 pa
160 1040 sy
lae
o ult Yopal
fau f0a 0
433 lt 1246
eo0
Manizales la 0
804
pa 0 0
318 o
m
a ra
00

263 Bogota 0 0
1000 ic 0
10

200
480 ua
Ibague G
0 0
0
100 470 0 0
0
0
0
0 Villavicencio 0
0 0
l a

0
nt r

0
900
Ce

0
0
0 HAUTERIVIAN
0 100 200 Km BARREMIAN
0

Neiva THICKNESS
San Jose del
800
ena

(meters)
de

Guaviare
ar

a
Mac

0
r

er

Si
La

800 900 1000 1100 1200 1300 1400

Figure 2.15: Hauterivian-Barremian thickness (meters) without palinspastic restoration. Thick lines represent
palaeo-faults believed to be active during Hauterivian-Barremian time.

During late Cenomanian, Turonian and Coniacian times the tectono-eustatic base level reached
its maximum Mesozoic level. The sea flooded the entire northwestern corner of South America and
dark gray shale was deposited from Venezuela to northern Peru. (Thery, 1982 in Fabre, 1985a, 1986).
Villamil (1993) recognized smaller relative tectono-eustatic level cycles during this time interval.
A relative tectono-eustatic base level rise during late Cenomanian (Villamil, 1993) induced a
slight deepening of the basin and a notorious decrease of the detrital supply to the basin. This lead to
basin starvation and the slow deposition of black laminated shale or micritic limestone pelagic facies in
the distal parts of the basin. The maximum flooding surface located at the Cenomanian-Turonian
boundary is characterized by a highly fossiliferous concretion horizon within the Frontera Fm and
lower part of San Rafael Fm (Villamil, 1993). During Turonian-Coniacian time the present day LLA
foothills were flooded (Cooper et al., 1995) but not the entire LLA area (Figs. 2.7 to 2.11).
From Middle Turonian to late Coniacian a gradual progradation and shallowing upward during
deposition of the upper part of the San Rafael Fm and the Villeta Gp in the Upper MV was related to a
relative tectono-eustatic level fall (Villamil, 1993).

28
Chapter 2

Uribante
Maracaib o

Trough
es
Basin

as
qu
hi gh

Luc

s
1400 c

de
a u
Cucuta
M Tr o Barinas

An
S an

a
Basin

id
er
de

M
nia
t

ra
aul

1300 Bucaramanga

Ser
f

aeo
ille ra

i nvi ena
Arauca

a pal

l
b- B agda
Cor d

tlt
fau

Ser
as
M
eo
ala
z p

n
zo
1200

asi
Su
Medellin

are
tem

u lt
bla

fa

b-B
Su
s ys

Ta

o
lae

Su
Llanos Orientales

pa
au lt

ca
Basin (LLA)

ya

cuy
eo f

Bo

em
Co

st
pala

sy
Tr Tunja Yopal
1100 an

ul t
sfe

fa
rp
im a

o
Cu ala

lae
nd m
eo te

pa
s
Bit u

Su inam fau sy

a
b-B ar lt ul
t

ar
asi ca fa

m
Manizales Na n

T
za o
lae

na
re pa APTIAN
th

sia
pa o
Cu
lae m
1000 Bogota ra
of ic
Ibague au ua PALEOGEOGRAPHY
lt G
Positive relief (absent)

Villavicencio Alluvial fan and fluvial


sandstones and mudstones
Coastal plain predominantly
l a

sand ston es
nt r

900
ley

Coastal plain predominantly


Ce

Val

mudstones
Littotal to inner shelf
na

0 100 200 Km sand ston es


ale

Shallow marine inner shelf


agd

carbonates
M

Neiva Shallow marine inner shelf


per

mudstones and siltstones


Up

La Ma ca rena

800
Ser rania de

Ou ter shelf sh ales or


San Jose del carb on at es
Guaviare
Turbiditic sandstones

800 900 1000 1100 1200 1300 1400 1500

Figure 2.16: Aptian palaeogeography without palinspastic restoration (modified from Etayo Serna et al, 1968,
Geotec, 1992; Cooper et al., 1995).

In the Upper MV during the late Coniacian to Santonian the transition from the uppermost
Villeta Gp, deposited in an inner shelf, to the lower chert unit of the Olini Gp, deposited in a deeper
middle shelf (Jaramillo and Yepez, 1994; Ramirez and Ramirez, 1994), points to a deepening of the
basin and relative tectono-eustatic level rise. (c.f. Etayo-Serna, 1994, his Figure 2).
During the Santonian, Campanian, Maastrichtian and Paleocene a general regression and
progradation was recorded by littoral to transitional coastal plain facies (Guadalupe Gp, Guaduas Fm).
The Guadalupe Gp sands represent two cycles of westward shoreline progradation, aggradation and
retrogradation, dominated by high energy quartz-rich shoreface sandstones derived from the Guyana
Shield (Cooper et al., 1995; Figs. 2.7 to 2.11). Regression did not occurred continuously but with minor
transgressive events recorded by fine-grained siliceous and phosphatic facies (Fllmi et al., 1992;
Plaeners Fm, Olini Gp, upper part of La Luna Fm; Figs. 2.7 to 2.11).
A sequence boundary occurs at the base of the medium shale unit (Lower-Middle Santonian
according to Villamil, 1993 or late Santonian-early Campanian according to palinostratigraphy by
Jaramillo and Yepez 1994; Etayo-Serna, 1994) of the Olini Gp and the shallow water El Cobre
sandstones of Barrio and Coffield (1992) in the UMV (Villamil, 1993). The shallow water marine
sands of the Arenisca Dura Fm represent a lower forced regression system tract (sensu Posamentier et
al., 1992; Cooper et al., 1995; Figs. 2.7 to 2.11).

29
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes

Serrania Maracaibo
de Perija Basin

Lucas
1400

s
Barinas

de
Cucuta

An
0 Basin

0
40

10
0

e Sa

ida
20

100
0
20

er
0

M
50
nia d
110

100
220
1300 185
154 211
Bucaramanga

Serra

200
175
227 0 Arauca
176

700
45

em

4
3
212 217 831

yst
il lera
0

lt s
0 190 930
100

40
200

f au
349
Co rd
465
1200

eo
0
Medellin 0 0
10
486
20 400
840

100

pala
0
310 20 113 0

mo
50
306
714

m
16 0

ara
180

te
227 0
70

ys
182 Llanos Orientales

aic
140

lt s
640 Tunja

Gu
0

au
20 Basin (LLA)

of
0 410
1100 Tr

lae
400
9
500
840 an 140 Yopal

pa
sfe 458
0
rp

a
ar
ala
0

m
eo 0
70

Ta
Manizales fau 0
1575 l 0
t

na
sia
476 0 0 0

Cu
0
7

500 0 0
74

Bogota 0
1000 40
214
0
Ibague 240

208
0
200
0
0
0 Villavicencio
575 20 0
0
l

73
a

386
0
nt r

215
0
Ce

900 0
195 0
0

APTIAN
0 100 200 Km
0
10
0

121
Neiv THICKNESS
a
La Macarena
Serrania de

800 (meters)
San Jose del
0
Guaviare

800 900 1000 1100 1200 1300 1400 1500

Figure 2.17: Aptian thickness (meters) without palinspastic restoration. Thick lines represent palaeo-faults relieved
to be active during Aptian time.

Mudstones of the upper part of the Arenisca Dura Fm and shales of the Plaeners Fm represent a
transgressive system tract (Cooper et al., 1995). During Santonian-early Campanian a maximum
flooding surface and a relative tectono-eustatic level rise from the medium shale unit of the Olini Gp to
the Upper Chert unit has been interpreted by Villamil (1993). In contrast to the EC, where the
Cretaceous maximum flooding surface occurred at the Cenomanian-Turonian boundary, the maximum
flooding in the LLA occurred during the Campanian (CS at the top of Gachet Fm, Fajardo et al., 1993;
Cooper et al., 1995 their Fig. 4).
During late Campanian the relative sea-level continued to drop and shallow marine oxygenated
environments prevailed in the EC. The Labor Fm represents a sand dominated forced regression system
tract (Cooper et al., 1995).
The regional regression and long term relative tectono-eustatic level fall was interrupted by a
small cycle of relative tectono-eustatic base level rise during late Campanian or early Maastrichtian (?)
as suggested by Fllmi et al. (1992) and Villamil (1993) during deposition of the Upper Plaeners Fm.
According to Cooper et al. (1995) the Upper Plaeners unit represents a condensed marine mudstone
deposited during a relative tectono-eustatic level rise.
During early (?) Maastrichtian time the eastern part of the basin was filled by the littoral quartz
sands of the Arenisca Tierna Fm (Fabre, 1985a). According to Cooper et al. (1995) the latter represents
the transgressive systems tract of the next sequence (Figs. 2.7 to 2.11). The gradual uplift of the

30
Chapter 2
western margin of the UMV supplied clasts of metamorphic rocks that were accumulated by fluvial
systems close to the sea in a braided delta (Cimarrona Fm, Gmez and Pedraza, 1994). Sands were
dispersed along a littoral belt (Monserate Fm, Ramrez and Ramrez, 1994), while in the more distal
areas carbonate silt (Daz, 1994a) or mud (Umir Fm) accumulated (Etayo-Serna, 1994).

4. SUBSIDENCE ANALYSIS

The stratigraphic record provides information about vertical crustal movements and/or sea-level
changes within a basin. Basin subsidence or basement subsidence is the result of both a thermo-
mechanical component called tectonic subsidence and a component due to sediment and water loading.
Tectonic subsidence is the undistorted basin subsidence that would have occurred in the absence of
sedimentation, therefore it is related to the geodynamics of the basin.
In order to quantify the tectonic component of subsidence of the studied basin the 1D-
backstripping technique was used (Steckler and Watts, 1978; Bond and Kominz, 1984). The method is
explained in Sclater and Christie (1980), Bond and Kominz (1984) and Bessis (1986). For this purpose,
tectonic subsidence has been calculated from the stratigraphic record adopting local isostasy to correct
for the effect of sediment loading. Corrections for compaction were made using porosity-depth
relationships on the basis of the observed lithologies using standard mean exponential relations and
material parameters (cf. Sclater and Christie, 1980).
Most of stratigraphic columns are from published literature; well data are from Ecopetrol. As
pointed out by Restrepo-Pace and Villamil (1998), and locally demonstrated by Restrepo-Pace (1989),
thickness measured from outcrops in the EC can be exaggerated due to structural thickening.
Cretaceous thicknesses are not well controlled (Fabre, 1986, 1987; Restrepo- Pace, 1989). In order to
eliminate or reduce that possible source of error I carefully checked thickness of each stratigraphic
column with available geological maps to avoid structural repetitions. I also checked the consistency of
thickness between neighbouring sections. Villamil (1993) described in detail four partial (Albian to
Santonian) stratigraphic columns. I checked thickness of stratigraphic columns reported in the literature
with these well-constrained sections. In the four sections described by Villamil (1993), only the Cocuy
stratigraphic column is located in a depocenter. The other sections studied by Villamil (1993) are close
to a palaeo-high (Villa de Leiva section) or in the less subsiding southern border of the basin (Mesitas,
Olini and Yaguar sections). Based on the assumption of a Late Cretaceous passive margin, Villamil
(1993) estimated clinoform profile geometry during Late Cretaceous with a slope of 0.057. This value
indicates an almost horizontal palaeo-sediment surface profile. Thus the difference in thickness without
considering this slope is very small.
The effects of palaeo-bathymetry have been taken into account, using sedimentary facies and
faunal content as interpreted in literature. Detailed sedimentological analysis of the deepest facies
concluded that palaeo-water depths never reached values greater than 200 m (Sarmiento, 1989;
Villamil, 1994). Most of the Triassic and Jurassic sedimentary record is continental, and the marine
facies are mostly Cretaceous, with prevailing water depths 0-75 m (Sarmiento, 1989; Villamil, 1994).
Errors in water depths are likely to be less than 50 m. Sea-level changes were estimated from the curve
proposed by Villamil and Arango (1998) for the Cretaceous of Colombia and Haq et al. (1987) for
other time intervals. Ages are based on the data given in literature mainly on the regional stratigraphic
cross-sections presented by Cooper at al. (1995). In order to express ages in Ma I used the geological
time scale proposed by Gradstein and Ogg (1996). Unconformities are also included with their age and
duration in Ma. Additional parameters for forward modelling are initial crust and lithosphere thickness
and densities. Table 2.2 shows additional parameters used in forward modelling, which are average
accepted values for normal continental lithosphere.
Pulses of fast tectonic subsidence of the basement have been interpreted in terms of tectonic or eustatic
processes. Only those pulses of fast tectonic subsidence correlatable to normal fault activity were
interpreted as produced by lithosphere extension and the slower subsidence, generally later, was
interpreted as produced by thermal re-equilibration of the lithosphere following the thermal anomaly
created by stretching.

31
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes

MODEL PARAMETERS VALUE


Initial lithospheric thickness 120 km
Initial crustal thickness 35 km
Asthenospheric temperature 1333 C
Thermal diffusivity 1 10-6 m2 s-1
Surface crustal density 2800 kg m-3
Surface mantle density 3400 kg m-3
Sea water density 1030 kg m-3
Thermal expansion coefficient 3.2 10-5 C

Table 2.2. Parameters used to calculate tectonic subsidence in the forward model.

Tectonic subsidence refers to the vertical downward movement of the basement top surface
underlying the sedimentary fill of the basin accumulated during the time interval being studied. As the
studied time interval is Mesozoic, for modelling purposes I considered tectonic subsidence referred to
the top surface of any pre-Mesozoic rock. Pre-Mesozoic rocks crop out only in local areas of the EC
(Santander-Floresta and Quetame-Garzn Massifs and the core of some anticlines in the flanks of the
EC) (Fig. 2.1). Although the geodynamic igneous/metamorphic basement is not such pre-Mesozoic top
surface due to the local presence of highly compacted Palaeozoic sedimentary rocks, I did not include
this Palaeozoic record due to large uncertainty in their age, thickness, facies and lateral continuity. This
implies to neglect any compaction of these rocks during Mesozoic and Cenozoic time, which probably
was very small. Since most compaction occurs early just after deposition, most compaction of these
rocks occurred during Palaeozoic time. I extended downward the Cretaceous stratigraphic columns
down to the basement, interpolating data from the nearest stratigraphic columns and using the
palaeogeographic and restored thickness maps. For areas absent of surface outcrops, I extended the
most representative Cretaceous columns down to the pre-Mesozoic top surface according to the
palaeogeographic and isopach reconstructions. For these stratigraphic columns I studied the tectonic
subsidence referred to a pre-Mesozoic basement. Additionally, to constrain the tectonic subsidence
evolution during the Cretaceous avoiding uncertainties reconstructing Triassic and Jurassic section, I
studied the Cretaceous tectonic subsidence referred to the top surface of any pre-Cretaceous rock.
In the Cundinamarca sub-basin (Figs. 2.1, 2.12 and 2.18) a relatively complete Cretaceous
stratigraphic section crops out in the eastern and western flanks of the mountain range, yet in the axial
region the deepest wells have not penetrated the complete Cretaceous section. Stratigraphic information
for the missing section in the axial area has been extrapolated from the exposures at the flanks of the
EC.
Relative sea-level and palaeo-water depth effects. One common difficulty with subsidence
analysis is the need to separate the effect due to eustasy from tectonics (e.g. Parkinson and
Summerhayes, 1985). In order to investigate the effects of sea-level changes I used the best available
data from selected stratigraphic columns (Table 2.3) to calculate the tectonic subsidence both with and
without the incorporation of sea-level changes (Fig. 2.19).
Comparison of the tectonic subsidence curves with the palaeo-water depth curves and also with
sea-level curves (Fig. 2.19) suggests the following:
(1) In general the amplitude of the tectonic subsidence curve is greater than that of the sea-level
curve (e.g. Cocuy section, Fig. 2.19). Only where tectonic subsidence is small, the magnitude of
sea-level changes is of the same order as short wave tectonic subsidence variations (e.g. Q Olini
section, Fig. 2.19). Thus calculated tectonic subsidence in stratigraphic columns with small
tectonic subsidence is more sensitive to sea-level changes. For those sections with relatively
minor tectonic subsidence (comparable in magnitude with sea-level changes), a sea-level
correction is necessary in order to separate the tectonic and sea-level signals.

32
Chapter 2

Serrania Maracaibo
de Perija 11 Basin 12

s
s

de
1400

Luca
600- Cucuta

An
7 800 Barinas

ida
Basin

an

er
de S

M
ania
1300 600-
2 1000

Serr
Bucaramanga
Arauca

1 100-
500 Llanos Orientales
1200 9 1200- Basin (LLA)
1800
Medellin 800-
3
il ler a

1700
100-
A 10 200
Co rd

Tunja

1100
Yopal
4 1100-
1800
Manizales 0 100 200Km

Bogota
1000 5 750- 600- B
Ibague 8 1400 BASIN COMPARTMENTS
1200
300-
6 850 1. CENTRAL CORDILLERA
Villavicencio 2. MIDDLE MAGDALENA VALLEY
3. MAGDLENA-TABLAZO SUB-BASIN
l a

4. CUNDINAMARCA SUB-BASIN
nt r

900
5. SOUTHERN CUNDINAMARCA SUB-BASIN
Ce

6. UPPER MAGDALENA VALLEY


10 100- 7. SANTANDER FLORESTA MASSIF
200 8. S CUNDINAMARCA AND W COCUY SUB-BASINS
Neiva Basin Cretaceous 9. COCUY SUB-BASIN
800 compartment tectonic 10. LLANOS ORIENTALES
subsidence 11. MARACAIBO (CATATUMBO) SUB-BASIN
(m)
12. MACHIQUES TROUGH

800 900 1000 1100 1200 1300 1400 1500

Figuere 2.18: Cretaceous basin compartments (sub-basins) and their tectonic subsidence in meters. Section A shows
the tectonic subsidence of the Tablazo and Cocuy sub-basins. Section B illustrates the tectonic subsidence of the
Cundinamarca sub-basins.

(2) Uncertainty in palaeo-water depth increases with palaeo-water depth (as accuracy in palaeo-
bathymetry estimation decreases with increasing water depth). However, in the cases here
analysed the absence of truly deep palaeo-depositional environments makes this uncertainty
small.
(3) Comparing the tectonic subsidence curves calculated involving sea-level changes reveals that in
many cases differences in the subsidence curves attributable to the latter are beyond the time
resolution of the database. The frequency of the sea-level changes proposed by Villamil and
Arango (1998) for the Cretaceous of Colombia and by Haq et al. (1987) are in general higher
than those changes inferred in subsidence, due to the smaller number of data points in the
subsidence curves. Clear identification of a high frequency eustatic signal would require more
accurate time resolution in the data and adds very little information regarding regional crustal
studies. Taking into account the resolution of the data, tectonic events should be limited to those
clearly different from the sea-level signal. Pulses of tectonic subsidence with amplitude and
frequency of the same order of magnitude as the sea-level variation curve are difficult to identify
without filtering the sea-level curve signal.

33
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes
In order to distinguish small tectonic events, a sea-level correction would be necessary.
However a correction becomes impractical for the following reasons:
(1) Less time resolution in most of the stratigraphic columns compared to the sea-level curve;
(2) Controversy about the precise significance of the seal-level cycles (Miall, 1992, 1993);
(3) Differences between the global sea-level curves (Haq et al., 1987) and the cycles proposed by
Villamil and Arango (1998) for the Cretaceous of Colombia, which can include regional tectonic
effects; and
(4) Uncertainty in the magnitude of sea-level variations in the sea-level curves.

For practical purposes, therefore, in order to compare tectonic subsidence patterns of most of
the stratigraphic columns, the general analysis was done without including detailed sea-level effects. It
is assumed that the identified stretching events mainly represent a tectonic signal because:

(1) Each of them occurs in specific tectonic active portions of the basin and is not uniformly and
synchronically over the whole area as would be expected for eustatic changes; and
(2) Each of them can be correlated in space and time with normal fault activity.

Late Cretaceous eustatic sea-level rise may have contributed to the observed basin-wide
subsidence (Fabre, 1987), but this could also be the result of intraplate compressional stress (Cloetingh
et al., 1985). In addition short-term sea-level fluctuations may have played an important role in the
sedimentation distribution. The occurrence of such short-term and relatively low magnitude (compared
to tectonic effects) sea-level fluctuations can not be easily detected from the subsidence data.

CRETACEOUS SUB-BASINS STRATIGRAPHIC REFERENCES


COLUMN
1. Central Cordillera San Flix Rodrguez and Rojas (1985), Etayo-Serna
(1985)
2. Middle Magdalena Valley Tablazo Ward et al. (1973), Ecopetrol et al. (1994)
3. Magdalena-Tablazo Villa de Leiva Etayo-Serna (1968), Villaml (1994)
4. Cundinamarca Villeta/Yacop Sarmiento (1989), Rubiano (1989)
5. Southern part Cundinamarca Apulo Cceres and Etayo-Serna (1969), Villaml
(1994)
6. Upper Magdalena Valley Q Olini Villaml (1994), Etayo-Serna (1994)
7. Santander-Floresta Massif Tibasosa Renzoni (1967), Renzoni et al. (1967)
8. Eastern Cundinamarca and Western Cqueza/Paz de Ro Hubach (1957), Renzoni (1965a,b)
Cocuy
9. Cocuy Cocuy/Guateque Fabre (1986), Ulloa and Rodrguez (1976),
Villamil (1994)
10. Llanos Orientales Arauca-1 Ecopetrol and Beicip (1995)

Table 2.3. List of stratigraphic columns used to study sea-level effects.

4.1. TECTONIC SUBSIDENCE DURING TRIASSIC AND JURASSIC TIME

For the subsidence analysis of the Triassic sedimentary record I assumed, following Geyer (1982), that
the poorly fossiliferous sediments (El Sudn Fm.) of the Cinaga de Morrocoyal north of the Serrana
de San Lucas, with lithology and relative stratigraphic position similar to those of the Luisa Fm. of the
Payand region, are Triassic and time-correlative. I also assumed that Triassic sediments were
accumulated in the western flank of the EC. With these hypotheses is easier to explain the Jurassic and
Cretaceous subsidence in the western part of the EC, as partially produced by thermal subsidence of an
earlier rifting event. Further discussion is based on this assumption.
The tectonic subsidence curves obtained for Triassic and Jurassic times are shown in Figure 2.20.

34
)
Chapter 2

(
A P U L O A ir loaded tectonic subside nce
AP U L O
A g e(M a) -150
140 120 100 80 60 40 20 0

h
0
-100
100
200 120 100 80 60 40 0 -50
160 140 20

te m p
300
0
400
500

d
50
600 100
700

)
800 150
900
200
1000

(
Q O L IN I A ir lo aded te ctonic subside nce
Q OLINI
A g e(M a)
-150
140 120 100 80 60 40 20 0
0
-100

h
100 120 -50
160 140 100 80 60 40 20 0
200

te m p
0

300 50
d
400 100
)

500 150

600 200
(

T I B A S O S A A ir loade d t e c t o n ic subsid ence


A g e(M a) TI B A S O S A
140 120 100 80 60 40 20 0 -100
0
100
-50
200
160 140 120 100 80 60 40 20 0
300
te m p

0
400
500
50
d

600
700
800 100
900
1000 150

P A Z D E R IO Air l o a d e d t e ctonic s u b siden ce


A g e(M a) P A Z D E R IO
160 140 120 100 80 60 40 20 0 -150
0

200 -100

400
-50
te m p

600 160 140 120 100 80 60 40 20 0

800 0
d

1000 50
1200

1400 100

Figure 2.19: Effect of sea-level changes and palaeo-water depth on tectonic subsidence curves. On the left tectonic
subsidence curves. Left panel: dashed red line: tectonic subsidence including sea-level and palaeo-water depth.
Dotted blue lines: tectonic subsidence without considering either sea-level or palaeo-water depth. Continuous black
line: tectonic subsidence considering palaeo-water depth but not sea-level change. Right panel: blue line: sea-level
curve (modified after Haq et al., 1987 and Villamil and Arango, 1998) and black line: palaeo-water depth.

35
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes

e
p
C A Q U E Z A A ir loaded te ctonic sub sidence CAQUEZA

t
-100

age(M a)
-50

h
160 140 120 100 80 60 40 20 0 160 140 120 100 80 60 40 20 0
0 0

500 50
1000

d
100
1500

e
2000
150

p
C O C U Y A ir loaded tectonic subside nce COCUY

age(M a)
-150

t
160 140 120 100 80 60 40 20 0
0 -100

500 -50
160 140 120 100 80 60 40 20 0
1000 d 0

1500 50

2000 100
e

2500 150
p

G U A T E Q U E A ir lo a d e d t e c t o n i c subsid ence GUATEQUE

age(M a) -100
160 140 120 100 80 60 40 20 0
0 -50
200
160 140 120 100 80 60 40 20 0
400
0
600
800
50
d

1000
1200
1400 100
1600
e

1800 150

A R A U C A - 1 Air loa ded tectonic su bsidence A R A UC A - 1

age(M a) -100
100 80 60 40 20 0
0
-50
200
160 140 120 100 80 60 40 20 0
400 0
600
800 50
1000
d

1200 100
1400
1600 150

Figure 2.19: Continued

36
Chapter 2
4.1.1. Basin compartments

Similarities in the shape of the subsidence curves related to their geographic position confirm
that Triassic-Jurassic sedimentation occurred in two separate basin compartments, each one with its
own subsidence history and sedimentary fill:
(1) Upper Magdalena and Cienaga de Morrocoyal (region A in Fig. 2.20). Subsidence curves of the
Upper Magdalena-Cienaga de Morrocoyal show two fast subsidence events (Fig. 2.20). The total
tectonic subsidence that occurred in this area during Triassic-Jurassic time varies between 100 m
up to a maximum of 1000 m (Fig. 2.20).
(2) Eastern Cordillera (Region B in Fig. 2.20). Subsidence curves suggest three fast subsidence
events (Fig. 2.20). Total tectonic subsidence that occurred during Triassic-Jurassic time in this
region varies from 100 m to 1500 m (Fig. 2.20).

4.1.2. Triassic-Jurassic subsidence events

Distinct tectonic subsidence phases in different basin compartments demonstrate the dominant
role of tectonic processes on subsidence. In this section I identify fast subsidence events and compare
them to the regional geological observations to see if they can be interpreted as produced by lithosphere
extension.
Because of the limited number of control points defining the subsidence curves it is not
possible to define correlative events of increased subsidence in all the curves. However, most curves
suggest that at least three events of increased subsidence occurred in various sectors of the study area
(Fig. 2.20). A simpler interpretation for these curves would be a single increased subsiding event from
the Triassic up to Middle Jurassic, for a time span of more than 70 Ma. Based on rapid lateral changes
of thickness and facies several authors (Maca et al., 1985; Bayona et al., 1994; Mojica et al., 1996,
etc.) have proposed that the Triassic-Jurassic sediments accumulated in narrow rift basins. Guillande
(1988) described from outcrops in the UMV Jurassic normal faults affecting the Saldaa Fm. Drastic
differences in the subsidence curves for the different columns clearly indicate fault control on
subsidence. Rapid subsidence in rift basins usually is the result of lithosphere stretching (e.g. Ziegler,
1994). Thus the subsidence events were produced by lithosphere stretching phases. Each stretching
phase affected more intensively specific parts of the basin as shown by the subsidence curves.
However, there are unconformities separating several stratigraphic intervals deposited during Triassic
and Jurassic times as indicated by Cooper et al. (1995, their Figs. 4 and 6). These unconformities may
represent periods of reduced subsidence.
Tables 2.4a and 2.4b show the lithospheric stretching events interpreted from the subsidence
curves. The lithosphere stretching events are (Fig. 2.20):
Triassic event (although variable in different columns, comprised between 248 to 235 Ma,
time scale of Gradstein and Ogg, 1996, see data limitations in section 7.3). This event is best
represented in the Upper MV-Cienaga de Morrocoyal. Subsidence curves (Fig. 2.20), thickness
variations and the location of the present day fault patterns (Fig. 2.28) suggest that small narrow rift
basins formed on opposite sides of the present-day Magdalena-La Salina fault system, which probably
acted as normal master faults (Fig. 2.28). Obviously a complete network of normal faults is of common
occurrence in rift systems. Abrupt lateral changes of sediment thickness in the Upper Magdalena
Valley suggest differential subsidence at different faulted blocks (Mojica et al., 1996). The relative
location of these basins on opposite sides of major fault systems suggests that they probably were
separated by accommodation transfer zones (e.g. accommodation zone between the Rufizi and Kivu
grabens in the Western Rift arm of East African Rift system, Ebinger, 1989). Most rift basins consist of
half graben geometry, the polarity of which often changes along trend across accommodation transfer
zones (Ziegler, 1994). If the relative location of these rift basins has been preserved, their present map
distribution would suggest an en-chelon pattern. The hypothesis of a depocenter in the Cienaga de
Morrocoyal area is based on the assumption that early Triassic sedimentary rocks occur in this area as
suggested by Geyer (1982). The only Early-Middle Triassic sedimentary record, in the area of this
study, is the UMV Luisa Fm rocks according to Mojica et al. (1996).

37
e
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes
A . U P PE R M A G D A L EN A V A L L E Y M O R R O C O Y AL ( W E S T E R N PA R T OF C H I BC H A TER R A NE )

Age (Ma)
25 0 24 0 23 0 22 0 21 0 20 0 19 0 18 0 17 0 16 0 15 0 14 0 13 0 12 0
0
2 S an Luc as 6 Alpu jarra 6 P rad o
6 Coe llo L uisa
6 Neiv a
6 A tac o 50 0
6 P ay and e
6 Q Oli ni
2 El S 10 00
uda n Mo 6 R Sal da a
rr o c
o ya l

T
15 00

B. W E STER N E A S TER N C O R D I L LE R A ( E A S TER N P A R T OF C H I B C H A TER R A N E )

Age (Ma)
25 0 24 0 23 0 22 0 21 0 20 0 19 0 18 0 17 0 16 0 15 0 14 0 13 0 12 0
3 Chima 0
8 Bu ena vis ta
3 O nza ga
2 S an Luc as
9 B at a 50 0
3 Med ios
2 3 La Rus ia
Ca
sc
aj 10 00
al
es

2 Cimitarra
15 00

T
Figure 2.20: Tectonic subsidence curves obtained from the Triassic-Jurassic sedimentary record. Horizontal axis
represents age in Ma. Vertical axis represents tectonic subsidence in meters obtained from backstripping analysis.
Note also the vertical bars representing the fast tectonic subsidence events. Horizontal segments in the subsidence
curves represent times of no deposition without tectonic subsidence.
A. Upper Magdalena Valley-Morrocoyal (western part of Chibcha Terrane of Toussaint, 1995, see Fig. 2.5).
B. Western flank Eastern Cordillera (eastern part of Chibcha Terrane of Toussaint, 1995, see Fig. 2.5).
Location of stratigraphic columns with Triassic and Jurassic section is shown on Figures 2.2 and 2.5.

Acceptance of this idea would imply an alternative hypothesis of a single narrow rift basin
located in the area of the UMV as suggested by the latter authors.
Latest Triassic to Middle Jurassic event (comprised between 208 185 Ma). Subsidence
curves (Fig. 2.20) stratigraphic thickness and fault distribution (Fig. 2.29) suggest that narrow isolated
rift basins were located on opposite sides of the Magdalena-La Salina fault system. If the relative
location of these rift basins would be preserved, their present map distribution would suggest an
en-suggest an en-chelon pattern. The width of these basins increased as compared to early Triassic
ones (Mojica et al., 1996). A fast subsidence favoured marine ingression. Volcanic rocks were
abundant in this period.
Middle Jurassic event (approximately 180 176 Ma). Palaeogeographic and stratigraphic
thickness distribution indicates continued widening of rift basins. However they still remained
relatively narrow. Major depocenters developed in the MV and western flank of EC probably existed
within elongated NNE rifts on opposite sides of the Magdalena-La Salina fault system (Figs. 2.20 and
2.30). Major depocenters also developed in the palaeo-MV and the current location of the western flank
of EC probably in elongated NNE rifts. The palaeo-La Salina, palaeo-Suarez and palaeo-Boyaca fault
systems possibly limited those rift basins. According to some authors, small isolated rift basins
developed at other locations in the study area: e.g. small grabens in the area of the Santander Massif,
(Boinet, 1985; Boinet at al. 1985; Geotec, 1992; Kammer, 1993b), Perij (Shagam, 1975; Maze, 1984),
Mrida Andes (Ricardi et al., 1990, in Mojica et al., 1996), LLA (Numpaque, 1986 in Cooper et al.,
1995; Geotec 1992) and Maracaibo (Shubert and Ricardi 1980, in Mojica et al., 1996) areas. Volcanic
activity decreased at this time, mainly occurring in the Mrida Andes (basalts in La Quinta Fm, Maze,
1984).

38
Chapter 2

Basin Event 1 Event 2 Event 3


Stratigraphic
compartments Stretching Stretching Stretching
column Start End Start End Start End
Sub-basins * factor factor factor
(Ma) (Ma) (Ma) (Ma) (Ma) (Ma)

A 1 El Sudn 248.2 235 1.23 1.23 208 207 1.088 1.088
A 6 Alpujarra 208 185 1.045 1.045
A 6 Ataco 240 235 1.091 1.091
A 6 Coello-Luisa 248.2 235 1.061 1.061
A 6 Neiva 240 235 1.075 1.075
A 6 Prado 208 185 1.064 1.064
A 6 Q El Cobre 248.2 235 1.165 1.165
A 6 Q Olini 248.2 235 1.168 1.168 208 207 1.042 1.042
A 6 R Saldaa 248.2 235 1.163 1.163 208 207 1.122 1.122
B 1 San Lucas 208 185 1.061 1.061
B 2 Cascajales-1 248.2 235 1.132 1.132 177 176 1.349 1.349
B 3 Arcabuco 238 235 1.046 1.046 180.1 179 1.09 1.09
B 3 Chima 159.4 149 1.008 1.008
B 3 Medios 238 235 1.044 1.044 180.1 179 1.072 1.072
B 3 Tablazo-Lebrija 248.2 235 1.114 1.114 177 176 1.387 1.387
B 7 Riolitas Onzaga 208 185 1.025 1.025
B 8 Cqueza 150 149 1.023 1.023
B 9 Santa Maria de Bat 210 185 1.111 1.111
* Triassic-Jurassic sub-basins: A Upper Magdalena and Cienaga de Morrocoyal (western part of Chibcha
Terrane); B Eastern Cordillera (eastern part of Chibcha Terrane). Numbers refer to Cretaceous sub-basins shown
in Fig.2.18

Table 2.4a Mesozoic stretching events and stretching factors. Triassic and Jurassic stretching events and
lithosphere stretching factors calculated from stratigraphic columns where the entire Mesozoic sedimentary
record is inferred to be present.

Event 4 Event 5 Event 6


Basin compart Stratigraphic Stretching Stretching
sub-basin # column Start End Start End Start End Stretching factor
factor factor
(Ma) (Ma) (Ma) (Ma) (Ma) (Ma)

1 Abejorral 144 132 2.697 1.237
1 San Felix 141 132 1.024 1.024 112.2 104.4 1.004 1.004
1 San Pablo 142 132 1.081 1.081
2 Casabe-199 144 127.8 1.522 1.146 114 109.3 1.03 1.03
2 Cascajales-1 144 127.8 1.059 1.113 114 109.3 1.607 1.205
" " " " * 144 127.5 0.987 0.987 114 109.3 1.132 1.094
2 Infantas-1613 138 127.8 1.301 1.117 114 109.3 1.328 1.113
2 Lebrija-1 144 127 1.625 1.213 114 109.3 1 1
2 Llanito-1 138 127.8 1.313 1.098 114 109.3 1.478 1.131
2 Sabalo-1K 144 127.8 1.37 1.113 114 109.3 1.039 1.039
# Numbers indicate Cretaceous sub-basins as shown in Fig. 2.18.
* Modelled using Triassic and Jurassic actual or inferred stratigraphy.

Table 2.4b Mesozoic stretching events and stretching factors. Cretaceous stretching events and stretching
lithosphere factors from stratigraphic columns where the Cretaceous sedimentary record is present.

39
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes
Event 4 Event 5 Event 6
Basin compart Stratigraphic Stretching Stretching
sub-basin # column Start End Start End Start End Stretching factor
factor factor
(Ma) (Ma) (Ma) (Ma) (Ma) (Ma)

3 Arcabuco 140 127.5 1.209 1.209 114.8 112.2 1.069 1.011
" " " " * 140 127.5 1.099 1.099
3 Chima 144 127.5 1.43 1.265 114.8 112.2 1.082 1.026
" " " " * 144 127.5 1.47 1.261 114.8 112.2 1.063 1.019
3 Cimitarrra 143 127.5 1.354 1.31 114.8 112.2 1.329 1.111
3 Los Medios 142 127.5 1.111 1.111 114.8 112.2 1.256 1.077
" " " " * 142 127.5 1.043 1.043 114.8 112.2 1.096 1.043
3 Los Santos 136.5 127.5 1.131 1.088 114.8 112.2 1.009 1.009
3 Simacota 144 127.5 1.251 1.251 114.8 112.2 1.064 1.01
3 Tablazo 144 127.5 2.702 1.348 114.8 112.2 1 1
" " " " * 144 127.5 1.052 1.052 114.8 112.2 1 1
3 Vadorreal 134.5 127.5 1.585 1.179 114.8 112.2 1.374 1.08
3 Velez 142 127.5 1.151 1.245 114.8 112.2 1.219 1.072
3 Villa de Leiva 142 127.5 2.016 1.332 114.8 112.2 1.052 1.052
4 Chitasuga-1 142 130 1.196 1.196 121 112.2 1.536 1.187
4 La Calera 142 130 1.331 1.206 121 112.2 1.58 1.178
4 Quipile 141 130 1.298 1.126 121 112.2 2.046 1.281
" " " " * 141 130 1.511 1.147 121 112.2 1.301 1.115
4 Simijaca 142 130 1.247 1.247 121 112.2 3.238 1.318
4 Suba-2 142 130 1.29 1.194 121 112.2 1.622 1.187
" " " " * 142 130 1.085 1.085 121 112.2 1.221 1.142
4 Suesca-1 142 130 1.217 1.217 121 112.2 2.476 1.304
4 Suesca Norte-1 142 130 1.214 1.214 121 112.2 1.824 1.224
4 Sutamarchan 142 130 1.234 1.234 121 112.2 1.825 1.252
4 Tabio 142 130 1.227 1.227 121 112.2 1.587 1.206
4 Villeta 141 127.5 1.853 1.226 121 112.2 3.575 1.64
" " " " * 141 127.5 1.01 1.01 121 112.2 1.52 1.492
4 Yacopi 134.2 127.5 1.665 1.665 121 112.2 10 1.406
" " " " * 134.2 127.5 1.449 1.449 121 112.2 1.77 1.299
5 Apulo 127.6 127 1.792 1.164 121 112.2 1.261 1.049
5 Fusagasuga 127.6 127 1.571 1.13 121 112.2 1.466 1.071
6 Alpujarra 115.1 102.6 1.37 1.243
" " " " * 115.1 102.6 1.411 1.222
6 Coello 121 102.6 1.122 1.113
" " " " * 121 102.6 1.11 1.066
6 Girardot 121 102.6 1.366 1.157
6 Guataqui 121 102.6 1.366 1.173
6 Itaibe 108.5 102.6 1.176 1.087
6 Melgar 121 102.6 1.18 1.121
6 Neiva 119.1 102.6 1.407 1.124
" " " " * 121 102.6 1.312 1.082
6 Ortega 121 102.6 1.658 1.284
6 Prado 121 102.6 1.513 1.145
" " " " * 121 102.6 1.422 1.115
6 Q Calambe 119.1 102.6 1.477 1.1
6 Q El Cobre 119.1 102.6 1.154 1.093
" " " " * 121 102.6 1.275 1.024
6 Q Olini 121 102.6 1.137 1.128
" " " " * 121 102.6 1.166 1.049
7 Chivata 132.8 127 3.144 1.215
7 Cormichoque-1 122.4 122.3 2.029 1.179
7 Floresta 132.6 127 1.024 1.037
7 Guaca 132 127 1.259 1.075
7 Matanza 128.8 127 1.716 1.13
7 Tibasosa 132.8 127 2.046 1.181
# Numbers indicate Cretaceous sub-basins as shown in Fig. 2.18.
* Modelled using Triassic and Jurassic actual or inferred stratigraphy.

Table 2.4 b.: Continued

40
Chapter 2
Event 4 Event 5 Event 6
Basin compart Stratigraphic Stretching Stretching Stretching
sub-basin # column Start End Start End Start End
factor factor factor
(Ma) (Ma) (Ma) (Ma) (Ma) (Ma)

7 Tunja 133.5 127 2.204 1.185
7 Tunja-1 132 127 2.576 1.213
8 Bolivar-1 Corrales-1 132 127 2.071 1.186 98 93 1.205 1.054
8 Caqueza 144 127.5 1.234 1.228 98 93 1.29 1.129
" " " " * 144 127.5 1.215 1.204 98 93 1.002 1.032
8 Nazareth 131.5 127.5 1.493 1.155 98 93 1.022 1.043
8 Paz de Rio 144 127 1.385 1.154 98 93 1.07 1.07
8 Servita 132 127.5 1.565 1.138 98 93 1.166 1.072
9 Aguazul 142 127.5 1.205 1.309 98 93 1.096 1.098
9 Chita 139 127.5 2.525 1.482 98 93 1.037 1.037
9 Cocuy 139 127.5 3.489 1.657 98 93 1 1
9 Guateque 138 127.5 1.035 1.405 98 93 1.092 1.052
9 Labateca 132 127.5 1.478 1.285 98 93 1.581 1.161
9 Medina-1 144 127.5 3.605 1.118 98 93 1.103 1.103
9 Mojicones-1 139 127.5 1.36 1.23 98 93 1.157 1.109
9 Pajarito 142 127.5 1.169 1.303 98 93 1.108 1.104
9 R Cusay 136.5 127.5 1.643 1.35 98 93 1.233 1.076
9 San Luis de Gaceno 144 127.5 0.975 1.449 98 93 1.02 1.055
" " " " * 144 127.5 1.076 1.352 98 93 1.277 1.125
9 Sogamoso 136.5 127.5 1.459 1.354 98 93 1.06 1.06
10 Apiay-4P 85.9 85.8 1 1
10 Arauca-1 98 93 1 1
10 Ariari-1 98 93 1 1
10 Arimena-1 98 93 1 1
10 Balastrera-1 98 93 1 1
10 Cachama-1 98 93 1 1
10 Camungo-1 98 93 1 1
10 Cao Barulia-1 98 93 1 1
10 Cao Bravo-1 98 93 1 1
10 Cao Cumare-1 98 93 1 1
10 Cao Duya-1 98 93 1 1
10 Cao Garza-1 98 93 1 1
10 Cao Limon-1 98 93 1 1
10 Cao Rondon-1 98 93 1 1
10 Casimena-1 98 93 1 1
10 Centauro-1 98 93 1 1
10 Centauro Norte-1 98 93 1 1
10 Chaparrito-1 98 93 1 1
10 Chichimene-1 98 93 1 1
10 Corocora-1 98 93 1 1
10 Cravo Este-1 98 93 1 1
10 Cusiana-1X-2 112.3 112.2 1.095 1.025 98 93 1 1
10 El Palmar-1 98 93 1 1
10 Entrerrios-1 98 93 1 1
10 Garibay-1 98 93 1 1
10 Guarimena-1 98 93 1 1
10 Guarrojo-1 98 93 1 1
10 Joropo-1 98 93 1 1
10 Kioskos-1 98 93 1 1
10 La Cabaa-1 132 127.5 1.478 1.285 98 93 1.581 1.161 98 93 1.581 1.161
10 La Flora-1 98 93 1 1
10 La Maria-1 98 93 1 1
10 La Punta-1 98 93 1 1
10 La Tortuga-1 98 93 1 1
# Numbers indicate Cretaceous sub-basins as shown in Fig. 2.18.
* Modelled using Triassic and Jurassic actual or inferred stratigraphy.

Table 2.4 b: Continued

41
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes
Event 4 Event 5 Event 6
Basin compart Stratigraphic Stretching Stretching Stretching
sub-basin # column Start End Start End Start End
factor factor factor
(Ma) (Ma) (Ma) (Ma) (Ma) (Ma)

10 Leticia-1 98 93 1 1
10 Los Teques-1 98 93 1 1
10 Los Trompillos-1 98 93 1 1
10 Manacacias-1 98 93 1 1
10 Manacacias-2 98 93 1 1
10 Maremare-1 98 93 1 1
10 Metica-1 98 93 1 1
10 Morichito-1 98 93 1 1
10 Negritos-1 98 93 1 1
10 Pajaropinto-1 98 93 1 1
10 Piriri-1 98 93 1 1
10 Planas-1 98 93 1 1
10 Pomarroso-1 98 93 1 1
10 Remache Sur-1 98 93 1 1
10 Rubiales-1 98 93 1 1
10 San Joaquin-1 98 93 1 1
10 San Juan-1 98 93 1 1
10 Sardinas-2 98 93 1 1
10 ST-0-04 98 93 1 1
10 ST-CN-8 98 93 1 1
10 Stella-1 98 93 1 1
10 ST-GU-15 98 93 1 1
10 ST-GU-19 98 93 1 1
10 Surimena-1 98 93 1 1
10 Surimena Norte-1 98 93 1 1
10 SV-9 98 93 1 1
10 Tierrablanca-1 98 93 1 1
10 Tocaria-1 98 93 1 1
10 Upia-1 98 93 1 1
10 Valdivia-1 98 93 1 1
10 Voragine-1 98 93 1 1
10 Yucao-1 98 93 1 1
# Numbers indicate Cretaceous sub-basins as shown in Fig. 2.18.
* Modelled using Triassic and Jurassic actual or inferred stratigraphy.

Table 2.4b; Continued

4.2. TECTONIC SUBSIDENCE DURING CRETACEOUS TIME

The location of stratigraphic columns with Cretaceous section is shown on Figure 2.2. Figure
2.21 illustrates the great thickness of the Cretaceous section in some parts of the basin. Figure 2.22
shows the tectonic subsidence curves during Cretaceous time. Curves have been grouped according
to their similar shape and relative location within the basin into several basin compartments as
indicated in Figure 2.18.

4.2.1. Basin compartments

Tectonic subsidence curves (Fig. 2.22) and restored thickness maps (Figs. 2.13, 2.15 and
2.17) indicate that several basin compartments existed (Fig. 2.18). Distinct compartments had
different subsidence histories, and being bounded by former extensional fault systems, as suggested
by local geological evidence (Figs. 2.23 and 2.24), and constrained further by lateral thickness
changes at opposite sides of faults (Figs. 2.13, 2.15 and 2.17). Most of these faults have reversed
their relative movement during the Cenozoic (Fabre, 1987; Colletta et al., 1990; Cooper et al., 1995;
Figs. 2.23 and 2.24).
In the northern part of the EC comparison of Cretaceous subsidence in a W-E transect A
(Figs. 2.13 and 2.18) reveals the following trends:

42
S
W
C r
S
9 1 1 0 8 1 1 0 0

0 .

TWT(sec)
1 .

43
2 .
Chapter 2

A
3 .

4 .

Figure 2.21: Line drawing of a seismic section in the western part of the Cundinamarca sub-basin. Available surface stratigraphic and
structural control is indicated. The Cretaceous Villeta Gp. crops out along the whole seismic line. Note the great thickness of the
Cretaceous section. Location of this section is
show in Figure 2.2
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes
(1) Two sub-basins: the Cocuy (region 9) and Tablazo (region 3) rift sub-basins separated by the
less subsiding Santander-Floresta block (region 7).
(2) A regional westward decrease in tectonic subsidence with a maximum in the Cocuy sub-basin
(region 9) and a minimum in the Middle MV (region 2), suggesting a regional half-rift
geometry for the whole basin.
A comparison of the Cretaceous subsidence along a W-E transect B (Figs. 2.13 and 2.18) in the
southern part of the Cordillera at the latitude of Bogota, reveals the following:
(1) The existence of a single rifted basin, the Cundinamarca sub-basin (region 4, Fig. 2.18).
Thickness variations of the drilled section in the axial region with respect to the sections exposed
along the flanks of the Cordillera within specific time intervals, suggests an absence of less subsiding
ridges or highs in the axial region. Gravity models calculated by Calvache and Muoz (1984) and
Kellogg and Duque (1994) support the latter statement.
(2) Tectonic subsidence during the earliest Cretaceous (Berriasian to Hauterivian) was
maximal in the eastern side of the Cundinamarca sub-basin (region 8, Fig. 2.18), indicating that a
first stretching event mainly affected the eastern Guaicramo normal fault system. Later during the
Aptian, however, subsidence was maximal in the western side of the Cundinamarca sub-basin
(region 4, Fig. 2.18) suggesting that a second stretching event mainly affected the western Bituima
fault system. The total tectonic subsidence during the whole Cretaceous was slightly greater in the
western side of the basin (region 4). A thick Cretaceous section in the western side of the
Cundinamarca sub-basin is illustrated in Figure 2.21.
In the easternmost LLA area (region 10, Fig. 2.18), sedimentation started during Late
Cretaceous and the total tectonic subsidence during the Cretaceous was small (100 to 200 m)
compared to the EC and MV. Subsidence probably was produced by flexural thermal subsidence
(Watts et al., 1982), and by water loading due to increase in palaeo-water depth.
Total tectonic subsidence in the Upper MV (region 6, Fig. 2.18), where marine sedimentation
started in Aptian time (Fig. 2.16), is significantly smaller than that of the EC and Middle MV (region
2).

4.2.2. Cretaceous fast subsidence events

Basin compartmentalisation indicates differential tectonic subsidence of fault-bounded


blocks (Figs. 2.23 and 2.24). Using the single instantaneous stretching model of McKenzie (1978),
Hbrard (1985) and Fabre (1987) proposed that a lithosphere-stretching event during early
Cretaceous was followed by thermal subsidence. I identified several rapid subsidence events (Fig.
2.22). In the following section I will attempt to test if these different rapid tectonic subsidence events
were structurally controlled by documented faults and correlate with evidence of tectonic extensional
movements, to infer if they were produced by lithosphere extension, or by other process.
Table 2.4b highlights the different rapid subsidence events as interpreted from the subsidence
curves for each stratigraphic column. Most subsidence curves allow identification of several steps of
increased subsidence. Comparison between different subsidence curves indicates that each rapid
subsidence event was not synchronous within the whole basin. The following events of rapid
subsidence have been identified for the Cretaceous (Fig. 2.22):
Berriasian to Hauterivian event (although variable in different stratigraphic columns
comprised between 144 to 127 Ma). It occurred in the area of the EC and is best represented in its
eastern flank. Subsidence curves (Fig. 2.22) and thickness maps (Figs. 2.13 and 2.15) show evidence
for a wide (>180 km wide) asymmetrical half-rift basin divided by the Santander-Floresta high.
Maximum tectonic subsidence was associated with the pre-Guaicramo normal master fault
system that was the eastern boundary of the rift (Figs. 2.31 and 2.32). A second order half graben
was located at the current location of the western flank of the EC. This minor half graben probably
was associated with a palaeo-normal fault system approximately following La Salina-Bituima fault
system that was its western border (Figs. 2.13, 2.15, 2.31 and 2.32). To the south there was only one
depocenter, limited in the south by a NW-SE vertical transfer fault (Fabre, 1987; Sarmiento, 1989;
Figs. 2.31 and 2.32). If plate-tectonic interpretations by Pindell and Erikson (1993) are accepted this
fault probably was a rolongation of a transform fault of the proto-Caribbean ridge system (Geotec,
1992).

44
Chapter 2
Geotec (1992) suggested that NW-SE trending faults formed a graben with this orientation in
the northern Central Cordillera. However, subsidence in this area is small (Figs. 2.18 and 2.22).Early
Cretaceous turbidites at both flanks (Murca Fm, Sarmiento, 1989; Moreno 1990b, 1991; Cqueza
group, Pimpirev et al., 1992; Branquet, 1999) of the extensional basin (Fig. 2.12) can in this situation
be taken as evidence of tectonic instability associated with normal fault movement. Branquet (1999)
presented outcrop and seismic evidence of Cretaceous normal faulting. Normal faults imaged on
seismic sections (Figs. 2.23 and 2.24) confirm extensional tectonic movements that attest this rapid
subsidence event was produced by lithosphere stretching. Contrary to the Triassic and Jurassic
rifting, magmatic activity was reduced during the Early Cretaceous. Evidence for Early Cretaceous
magmatism is limited to (1) small mafic igneous intrusions described by Fabre and Delaloye (1983)
and Moreno and Concha (1993; Fig. 2.31) and (2) some volcanic input within Cretaceous shales
(Rubiano 1989; Villamil, 1994).
Aptian to Early Albian event (comprised between 121 102.6 Ma). This fast subsidence
event occurred at the current location of the EC and MV. However, it is best represented in the
southern part of the western flank of the EC and the Upper MV, indicating asymmetry in the basin.
During Barremian to Aptian time the basin was extended toward the south in the Upper MV (Figs.
2.16, 2.17, 2.22). Turbiditic deposits of Aptian age (Socota Member described by Polana and
Rodrguez, 1978; Sarmiento 1989) can in this case also be taken as evidence of tectonic instability
associated with this rapid subsidence event. The isopach map (Fig. 2.17) suggests that a master
normal fault system, located approximately at the present day Bituima Magdalena fault system, was
active (Figs. 2.33 and 2.34). In the area of the Upper MV a normal palaeo-Chusma fault system was
probably also active (Figs 2.33 and 2.34). This fast subsidence event is interpreted as produced by
lithosphere stretching. Evidence of magmatic activity within the basin is limited to some small mafic
intrusions (Fabre and Delaloye, 1983), and minor volcanic input (Rubiano, 1989; Villamil and
Arango, 1998). Some zircon fission track ages (see Chapter 3) from the Mrida Andes (Kohn et al.,
1984), Sierra de Perij (Shagam et al., 1984), and the Garzn Massif (Van der Wiel, 1991) seem to
correlate with this event. Van der Wiel (1991) interpretated this ages as related to an orogenic event
that affected the whole northwestern corner of South America between 100 and 80 Ma. However this
is in contradiction with strong stratigraphic evidence of a subsiding basin at these localities. Instead
if this correlation is valid probably the zircon ages reflect local uplift of faulted blocks located at rift
margins, as demonstrated using fission track data from a number of rift basins by Van der Beek
(1995). This author explains rift margin uplift by mechanical support of rift flanks resulting from an
upward state of flexure.
Cenomanian event (98 to 93 Ma) occurred in the eastern flank of the EC. No major
geological evidence has been reported for a tectonic event at this time (Fig. 2.22) in the EC. Based on
thickness changes across the Cusiana-Tmara fault system (Fig. 2.34), Cooper et al. (1995)
suggested that normal fault movement took place during the Campanian (73 to 80 Ma) and possibly
earlier time. During this time some small mafic intrusions were emplaced in the most subsiding parts
of the basin (Fabre and Delaloye, 1983). Van der Wiel (1991) reported in the Garzn Massif zircon
fission track ages comprised between 113 to 85 Ma (including Cenomanian) and interpreted them as
an orogenic event. However, in the area of extensional basins the zircon ages probably reflect
cooling and uplift of rift shoulders (after lithosphere stretching, c.f. Van der Beek, 1995) or intrusion
of small mafic bodies. Some Late Cretaceous fission track ages from the Central Cordillera (Gomez
et al., 1999; Toro et al., 1999) indicate initial uplift of this mountain range giving support to the
hypothesis of a Late Cretaceous magmatic arc in the Central Cordillera. During the late
Cenomanian-Turonian, the global sea-level maximum (Haq et al., 1987; Villamil and Arango, 1998)
correlates with this subsidence event, suggesting that increase in subsidence was driven by water
load during maximum sea-level. However, it is difficult to rule out the possibility of some tectonic
lithosphere extension component, especially for normal fault activity in the Cusiana-Tmara Fault
between the EC and the LLA basin (Cooper, et al., 1995).

45
e
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes

c
1. C E N TR A L C O R D I L L E R A
Age (M a)
25 0 24 0 23 0 22 0 21 0 20 0 19 0 18 0 17 0 16 0 15 0 14 0 13 0 12 0 11 0 10 0 90 80 70 60 50 40
0
San Pa bl o
Abejorra l
50 0
San Fe lix

T
2. MID D LE MAGD A L E N A VALLEY
Age (M a)
25 0 24 0 23 0 22 0 21 0 20 0 19 0 18 0 17 0 16 0 15 0 14 0 13 0 12 0 11 0 10 0 90 80 70 60 50 40
0
2 S an
Luc as

e
50 0
2
Ca
sc
aj
2 El al 10 00
Suda es
n Mo
rr o c o
ya l
2 Cimitarra
15 00

Saba lo-1 k
C asa be- 199 20 00
Inf ant as- 161 3

T
Lebrij a Llani to-1
C ascajale s-1 25 00

Tabl azo
30 00

e
3. TABLA Z O S U B - BASIN
Age (M a)
25 0 24 0 23 0 22 0 21 0 20 0 19 0 18 0 17 0 16 0 15 0 14 0 13 0 12 0 11 0 10 0 90 80 70 60 50 40
0
3 O nza ga 3 Chi ma
3 M ed io Los Sa nto s
s 50 0
3 La R Vado rea l
u s ia
Sim ac ota
C him a
Ta 10 00
C im it a bl a
M edios rr a z o

Arca buco
15 00

T
Villa de L eiva
20 00

4. C U N D I N A MAR C A S U B -BASIN
Age (M a)
25 0 24 0 23 0 22 0 21 0 20 0 19 0 18 0 17 0 16 0 15 0 14 0 13 0 12 0 11 0 10 0 90 80 70 60 50 40

0
Q u ip
il e

50 0

C hita suga- 1
Suesc a N-1 10 00
Sub a-2
Sues ca - 1 Suta m arc han
Sim ijaca
La

Tabi o
15 00
Ca

Villeta
le

T
ra

Yaco pi
20 00

Figure 2.22: Tectonic subsidence curves from the whole Mesozoic sedimentary record. Horizontal axis represents
age in Ma. Vertical axis represents tectonic subsidence in meters obtained from backstripping analysis. Vertical
shaded strips represent fast subsidence events. Numbers refer to basin compartments shown in Fig. 2.18. Note
also the vertical bars representing the fast tectonic subsidence events.

46
Chapter 2

c
5. SO U TH E R N C U N D I N A MA R C A S U B-BASIN

(c
Age (Ma)

e
25 0 24 0 23 0 22 0 21 0 20 0 19 0 18 0 17 0 16 0 15 0 14 0 13 0 12 0 11 0 10 0 90 80 70 60 50 40
0

)
50 0
Fus aga sug a
Apu lo

T
10 00

6. U P PE R M A G D A L E N A VALLEY
Age (Ma)
25 0 24 0 23 0 22 0 21 0 20 0 19 0 18 0 17 0 16 0 15 0 14 0 13 0 12 0 11 0 10 0 90 80 70 60 50 40
0
Alpujarr a
Pra do
C oello Luis a Ita ibe
N eiva
Ata co

e
Girardot 50 0
Pay an de Gua taq ui
Pra do
C oello
Q Olini N eiva
Alpujarr a
R S ald aa Orte ga 10 00

c
Q C al ambe
Q El C obre
Q Olini

T
15 00

7 . S A N TA N D E R F L O R E S TA MASSIF
Age (Ma)
25 0 24 0 23 0 22 0 21 0 20 0 19 0 18 0 17 0 16 0 15 0 14 0 13 0 12 0 11 0 10 0 90 80 70 60 50 40
0

)
Floresta

Gua ca
M ata nza 50 0
Tibas osa
Tunj a

ei
C orm ic hoque
Tunja - 1
C hivat a 10 00

T
8. EASTER N C UN D I N A M A RC A A N D WESTER N C OC O U Y S U B-BASIN S
Age (Ma)
25 0 24 0 23 0 22 0 21 0 20 19 0 18 17 0 16 0 15 0 14 0 13 0 12 0 11 0 10 0 90 80 70 60 50 40
0 0 0
8 Bu ena vis ta

C orm ich oque-1 50 0


N azar eth
Servita
Paz d e Ri o
Bolivar- 10 00
c

C orrales

C aqu eza 15 00
T

9 . CO C U Y SU B - B A SI N
Age (Ma)
25 0 24 0 23 0 22 0 21 0 20 19 0 18 17 0 16 0 15 0 14 0 13 0 12 0 11 0 10 0 90 80 70 60 50 40
0 0 0
Lab
ate

9 B at a
ca

M o ji 50 0
con
es
So
ga R C usay Pa
mo jar 10 00
so i to
C hit Agua zul
San a
Lu i s
c

de
Gac
eno 15 00
T

C ocuy Gua te qu e

20 00

Figure 2.22: Continued

47
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes
10. LLAN OS O R I E N TALES

Age (M a)

25 0 24 0 23 0 22 0 21 0 20 0 19 0 18 0 17 0 16 0 15 0 14 0 13 0 12 0 11 0 10 0 90 80 70 60 50 40
La T ortu ga-1 0
Cao Limon-1
Arauc a-1

50 0

T
Figure 2.22: Continued

Maastrichtian to Paleocene event (although variable in different columns, it is comprised


between 68 and 54.8 Ma). This fast subsidence event affected the axial part of the EC, its eastern
flank and locally the westernmost part of the LLA area. This event correlates in time with
deformation and uplift in the Central Cordillera (Jaramillo, 1978, 1981; Cooper et al., 1995). Some
authors (e.g. Kluth et al., 1997) recognized or suggested normal faulting in the Llanos area
approximately at this time.

5. CORRELATION OF FAST SUBSIDENCE EVENTS WITH MAGMATIC, EUSTATIC


AND PLATE-TECTONIC EVENTS

Below I explore the timing and nature of changes in subsidence, in the context of changes in
plate-tectonic regime and eustasy.

5.1. CORRELATION BETWEEN FAST SUBSIDENCE EVENTS AND SUBDUCTION-


RELATED MAGMATIC ARCS

The inferred Mesozoic stretching events seem to correlate in time with reduced magmatic
activity in the Central Cordillera (Fig. 2.25 modified from Aspden et al., 1987 and Guillande, 1988).
If the calc-alkaline (Alvarez, 1983) plutonic belts of the Central Cordillera were developed as
subduction-related magmatic arcs during Mesozoic times, as suggested by Aspden et al., (1987), the
extensional basins behind them may be interpreted as backarc basins. Extensional backarc basins are
developed when the velocity roll-back, due to fast subduction, exceeds the oceanward convergence
velocity of the overriding plate (Dewey, 1980; Cross and Pilger, 1982; Royden, 1993a,b). If
magmatic arc activity decreases with the oceanward convergence velocity of the overriding plate,
during times of reduced magmatic arc activity a constant roll-back velocity would exceed the
oceanward convergence velocity of the overriding plate, increasing extension and subsidence in the
backarc region. According to Aspden et al. (1987) the Triassic magmatic belt was controlled along
strike-slip faults. Evidence of the latter is presented by Restrepo-Pace (1995). Jurassic calc-alkaline
plutonism along the Central Cordillera (Restrepo et al., 1991) was interpreted by Aspden et al.,
(1987) as a subduction-related magmatic arc. However, Cretaceous plutonism is sporadically
developed only in the northern part of the Central Cordillera (Restrepo et al., 1991) whereas it is very
extensive in Peru (Cobbing, 1982 in Aspden et al., 1987). Aspden et al. (1987) suggested oblique
convergence and an offset in the subduction zone along a major NE-SW transform fault to account
for the notable absence of Cretaceous plutonism in southern Colombia and Ecuador.
Triassic-Jurassic magmatism in the Santander Massif (Ward et al., 1973; Restrepo-Pace,
1995) suggests a short magmatic arc segment (Restrepo-Pace, personal communication) that
correlates with the small subsidence in this region. Arcs tend to be associated with uplift due to
crustal thickening and thermal and physical effects of rising magma, or with reduced subsidence
(Smith and Landis, 1995).

48
Chapter 2

NW Seismic Profile ME94-1460 SE


0

t
ul
Carbonera Fm.

fa
a
ar
cl
ua
Ag
1.0

Carbonera Fm.

Fm . 2.0
d or
Mir a

TW T
Mirador Fm.

K Mirador Fm.
3.0
Basement
Basement

4.0

Basement

a. 5.0
0 5 km

NW Seismic Profile ME92-1260 SE


ult
Fa

0
lia
sa
Te

ult

.
Fm Le
Fa

on o n
Le Fm
ra

.
cla

.
Fm Ca
ua

era rb
Ag

on on
rb er
Ca a
Fm 1.0
.
TWT

m. 2.0
or F
Mirad

.
Fm
ad
or K 3.0
Mir
K
K
am ent
Bas
4.0
b.
0 5 Km

Figure 2.23: Seismic sections in the Medina foothills area, along the eastern border of the Cundinamarca sub-
basin. Note normal fault evidence during Cretaceous (K) in the Guaicramo palaeo-fault system along the eastern
border of the basin and contractional inversion of Cretaceous extensional faults occurred during Palaeogene time
as evidenced by lateral changes of thickness of the Palaeogene Carbonera Fm. (see Chapter 3). Note also the
thickness changes in the Cretaceous sedimentary fill (from Linares, 1996). Location of this seismic line is shown
in Figure 2.2.

49
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes

SE Seismic line DM87-1470e SE


5 km Arrugas-1 Pea de Oro-1 San Luis-12

0.0

0.5

1.0

1.5 K

A. 2.0
TWT (sec)

N 2 km SE
Seismic line CL861393
900 800 700 600 500 400 300 200 100
0.5

1.0

1.5
TWT (sec)

2.0

2.5

B.
3.0

Figure 2.24: Seismic sections in the Middle Magdalena Valley. Note normal faulting evidence during Cretaceous
(K) in the western border of the basin A. Seismic line DM87-1470e. Bottom of Cretaceous is the light green
lowermost reflector affected by normal faults. Note the in creasing thickness of Cretaceous (K) toward the SE and
normal faults (modified from Olaya and Serrano, 1998). B. Seismic line CL861393. Bottom of Cretaceous is the
purple lowermost reflector affected by normal faults. Note the lower-middle? Eocene unconformity truncating
Cretaceous strata. (From Mantilla, 2000). Location of these seismic lines shown in figure 2.2.

50
Chapter 2

Panam a dates

78 74 5
O c a Fault
Caribbean Sea 0

15
10 Zone V

s
5
Venezuela
Pa
na 0

lt
Fau
m

Buca
a

t
15

San
O tu
8

ra man
Zone I V
10

t a M a F ault
ga
rta
5
g

t
lt 0
au

15
aF

5 4 3 2 1
10
tin

Pacif ic Zone III


les

Ocean 5
ult
Pa

Zone V
t
ul
Fa

4
Fa

Zone I V 0

N
ult

o
m

Zone III 10
ra
Fa

Zone I I
ca
ral

Zone I I
ai

5
ti a

me

Gu
Pa

Ro

Zone I
0
ca
au

Trias s ic pluto ns 5 Zone I


C

a Ecuador Major f ault zones/s utures 0


0 50 100 150 200 250
P luton ages (Ma)
0

L i t h o sp h e re
3 stretc h i n g
e v e n ts.
50

P e ri o d s o f i n t e n se
100 m a g m a tic activity
5

4
A

150
P e ri o d s o f i n t e rm e d i a t e
3
m a g m a tic activity

2
200

P e ri o d s of sl i g h t o r no
m a g m a tic activity
1
250
0 50 100
b
C u m ul ati v e num ber of ra di o m e t r ic age
de t e rm in at io ns

Figure 2.25: Event correlation between lithosphere stretching in the area of the Eastern Cordillera and magmatic
activity in the Central Cordillera. 1 Triassic event. 2 Late Triassic-Early Jurassic event. 3 Middle Jurassic event. 4
Early Cretaceous Berriasian-Hauterivian event. 5 Aptian event. This correlation should be considered preliminary
because original data are heterogeneous. 94% of data are K-Ar (biotite, hornblende, muscovite or whole rock) and
6% of data are Rb-Sr (hornblende/biotite or whole rock).
A. Left panel: Principal structural/plutonic zones of Western Colombia. Right panel: Age distribution of
Mesozoic and Cenozoic plutonic activity in Western Colombia (modified after Aspden, 1987). B Cumulative
histogram of radiometric ages of plutonic bodies in Colombia (modified after Guillande, 1988). Periods of intense
magmatic activity are characterized by a rapid increase in the cumulative number of radiometric age
determinations for a time interval (low slope of the curve).

51
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes
5.2. CORRELATION BETWEEN FAST SUBSIDENCE EVENTS, PLATE-TECTONIC
EVENTS AND EUSTATIC EVENTS

Examination of Caribbean plate-tectonic interpretations (Pindell and Dewey, 1982; Burke et


al., 1984; Duncan and Hardgraves, 1984; Ross and Scotese, 1988; Jaillard et al., 1990; Pindell and
Barret, 1990; Pindell and Erikson, 1993; Meschede and Frisch, 1998; Figs. 2.2 and 2.3) and sea-level
curves (Haq et al., 1988; Villamil and Arango, 1998) suggests the Mesozoic rapid subsidence events
correlate with:
Triassic event. Reduced magmatic activity in the Central Cordillera (Fig.2.25) was then
related to oblique subduction (Aspden et al., 1987). The alternative hypothesis of intracontinental
rifting related to separation between South and North America, is applicable to Venezuela and
probably northern Colombia.
Latest Triassic-Middle Jurassic event. Reduced magmatic activity in the Central Cordillera
(Fig. 2.25) may be interpreted as backarc extension (Fig. 2.3). According to Aspden et al. (1987) this
reflects a change in convergence of the subducting plate from NNW/SSE to NW/SE.
Middle Jurassic event. Reduced magmatic activity in the Central Cordillera (Fig. 2.25) may
be correlated with backarc extension related to high-angle subduction (Fig. 2.3, Aspden et al., 1987).
According to Pindell and Erikson (1993) and Meschede and Frisch (1998) interpretations, opening of
the proto-Caribbean started in Northern Colombia.
Berriasian-Hauterivian event. According to the Cretaceous passive margin interpretation
(Pindell and Erikson, 1993), active opening of the proto-Caribbean was occurring north of Colombia
and also west of the palaeo-Central Cordillera (Fig. 2.4). If such an interpretation is valid for the
Berriasian to Hauterivian, stretching in the study area produced a failed-rifted arm related to a major
opening of the proto-Caribbean oceanic basin. The alternative plate-tectonic reconstruction attributes
Early Cretaceous rifting of the palaeo-EC to a backarc basin contemporaneous with reduced
magmatic activity in the Central Cordillera (Fig. 2.4). Reduced magmatic activity in the Central
Cordillera may be also associated with accretion of the Amime Terrane along the Romeral Fault
west of the Central Cordillera (Fig. 2.5) at about 125-130 Ma (Feininger, 1985, 1986; Aspden and
McCourt, 1986). Widespread dynamothermic metamorphism occurs in the Central Cordillera with
emplacement of blueschists (Aspden et al., 1987). The following evidence supports the hypothesis of
a back-arc basin located behind a partially emerged, less subsident palaeo-Central Cordillera
(magmatic arc?):
(1) The presence, in the western part of the Cundinamarca sub-basin, of Lower Cretaceous
sandstones with abundant volcanic lithic fragments and feldspar derived from a western detrital
source area as indicated by palaeocurrent data (Murca Fm. and tica Sandstone; Sarmiento, 1989;
Moreno, 1990b, 1991).
(2) The presence of progressive westerly onlap terminations of the Cretaceous carbonates on
the basement, observed in seismic lines, in the western border of the Cesar Valley, in northern
Colombia (in Mesozoic times part of the EC basinal area, Fig. 1.1; Audemard, 1991).
(3) Stratigraphical and petrographical evidence suggesting that during Berriasian (?) to
Valanginian time clastic sediments near San Felix in the western flank of the Central Cordillera
(between the Romeral and the Palestina Faults, Fig. 2.5) came from erosion of uplifted areas with
metamorphic rocks and small tectonic blocks with plutonic rocks (Rodrguez and Rojas, 1985).
Rodrguez and Rojas (1985) identified west-verging thrust faults post mid-Albian/pre-Miocene (but
according to these authors fault activity probably started during Albian time).
(4) The presence of some Cretaceous volaniclastic rocks in the Central Cordillera (Rodrguez
and Rojas, 1985), made up of mixture of pyroclastic and epiclastic fragments probably derived from
a magmatic arc.
(5) Relatively high concentration of volcanogenic clay minerals in Hauterivian-Barremian (0
to 30%), middle Albian (0 and 21%), and Turonian (6 and 9 %) shales of the Villeta Gp (Rubiano,
1989) and Valanginian-Hauterivian Rosablanca Fm (Moreno, 1989 in Rubiano, 1989; Moreno,
1990a) in the Cundinamarca sub-basin. Thin beds of volcanogenic clays or bentonites within the
Cenomanian- Turonian stratigraphic interval (Villamil and Arango, 1998) and the Salada Member of
the La Luna Fm (Patterson, 1970 in Rubiano, 1989), as well as subaqueous volcanic volcanic tuffs
within La Frontera and La Luna Fms in the MV (Restrepo-Pace, personal communication).

52
Chapter 2
(6) Jurassic (185 Ma) and Cretaceous ( 77 Ma) zircon fission track ages from the Central
Cordillera (Toro et al., 1999; Gmez et al., 1999) evidencing uplift. In the Ecuadorian Andes
(Rivadeneira, 1996) and the Central Cordillera (Rodrguez and Rojas, 1985) these authors have
suggested uplift or deformation during Late Cretaceous time. Late Cretaceous (85 to 113 Ma) zircon
ages in Venezuela and the Garzn Massif suggest a tectonic-thermal event interpreted by Van der
Wiel (1991) as an orogenic event. However in the area of the Mesozoic extensional basins it is
difficult to assume orogenic uplift. Probably these data indicate local uplift of faulted blocks located
at rift margins (c.f. Van der Beek, 1995). However this hypothesis needs further confirmation.
(7) Cretaceous igneous intrusions, as the Late Cretaceous Antioquian Batholith and others of
Early Cretaceous age in the Central Cordillera (e.g. San Diego, Cambumbia and Mariquit stocks;
Restrepo et al., 1991) which define a magmatic arc. However such a magmatic arc is not well
defined.

Aptian-Early Albian event. Pindell and Erikson (1993) hypothesized that during Aptian the
western border of Colombia was a passive margin. Spreading west of Colombia ceased and the
proto-Caribbean lithosphere began to subduct westward under the Amime-Antilles arc that was
approaching the western margin of northern South America (Fig. 2.4). If this interpretation is valid,
stress changes due to the above-mentioned plate-tectonics changes could have triggered stretching in
the study area. The alternative plate-tectonic hypothesis of Meschede and Frisch (1998) also assumes
possible extinction of spreading of the proto-Caribbean lithosphere northwest of Colombia and the
beginning of subduction of the Farallon/Pacific plate under the Panama-Costa Rica arc west of
Colombia (Fig. 2.4). Such plate-tectonic reorganisations could have induced lithosphere stretching in
the palaeo-EC basin. Bourgois et al. (1982a,b; 1987) proposed that obduction of oceanic terranes
took place over the Central Cordillera.
Cenomanian event. This age is included within the 100 to 80 Ma time interval proposed by
Duncan and Hardgraves (1984) and Hill (1993) for the formation of the Caribbean plate by partial
melting within the initial plume head of the Galapagos hotspot. The plate-tectonic interpretation of
Pindell and Erikson (1993) proposes that during the Cenomanian Colombia remained as part of a
passive margin without major changes in its plate-tectonic configuration. An alternative plate-
tectonic interpretation (Nivia, 1987; Meschede and Frisch, 1998) proposes that the formation of the
basalt plateau in the Caribbean area, west and northwest of Colombia, thickened the proto-Caribbean
crust. If such an interpretation is valid, buoyancy of the thickened Caribbean crust could have
impeded its continuous subduction under the Central Cordillera located at the continental margin.
Consequently, backarc subsidence is not favoured in this scenario. Late Cenomanian-Turonian global
sea-level maximum correlates with this subsidence event, suggesting that increase in subsidence was
driven by water load. However, the event affected only the eastern flank of the Cordillera, where the
maximum thickness of Cretaceous is present. While the maximum flooding surface for the
Cretaceous sediments of the EC is the Cenomanian-Turonian boundary (Villamil and Arango, 1998),
it is Campanian in the eastern LLA (Fajardo et al., 1993; Cooper et al., 1995; Figs 2.7 to 2.11). If
subduction of the Caribbean thick and buoyant lithosphere under South America was inhibited, it
probably exerted horizontal stresses on the northwestern margin of South America. Horizontal
stresses can induce local flexural lithosphere bending, which is maximal where the lithosphere is
weakest (Cloetingh, 1988; Cloetingh and Kooi, 1992). This process probably enhanced the relative
sea-level rise, creating a maximum Cenomanian-Turonian marine flooding surface in the depocenter
of the EC, characterized by weak lithosphere due to earlier stretching. In contrast, at that time
horizontal stress produced a submarine shallow water depth bulge in the LLA, which partially
compensated the maximum eustatic signal.
Maastrichtian-Paleocene event. All plate-tectonic interpretations agree that during latest
Cretaceous and probably Paleocene, the accretion of the Western Cordillera oceanic terranes along
the Cauca-Patia Fault occurred, producing deformation and uplift of the Central Cordillera.
According to Cooper et al. (1995) loading of the Western and Central Cordilleras led to the
development of an early pre-Andean flexural foreland basin in the area of the EC. However, subtle
uplift and erosion in the Tunja-Sogamoso axial region and initial inversion of the extensional basin
have also been proposed by Fajardo-Pea (1998). Increased subsidence in the axis of the
Cundinamarca sub-basin (Sabana de Bogot) could be the result of increased horizontal

53
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes
compressional stress (Cloetingh, 1988; Cloetingh and Kooi, 1992) associated with collision of the
oceanic terranes of western Colombia and deformation and uplift of the Central Cordillera.
Development of normal faults in the Llanos area (e.g. Kluth et al. 1997) could be the result of local
tensional stresses in the flexural bulge developed.

6. FORWARD MODELLING OF BASIN EVOLUTION

Subsidence analysis has allowed the identification of seven events of rapid tectonic
subsidence followed by slower subsidence rates. The first five events of rapid tectonic subsidence,
which can be clearly correlated with tectonic activity, are interpreted as stretching events followed by
periods of slower thermal subsidence. In order to quantify horizontal extensional movements
responsible for the observed subsidence and establish a quantitative framework for the pulsating rift
evolution of the lithosphere during Mesozoic basin formation, Sarmiento and Van Wees (in prep.)
have quantified extension rates by forward modelling of tectonic subsidence. They used an
automated forward modelling technique (Van Wees et al., 1996b), which will be briefly explained
below.

6.1. NUMERICAL MODEL

The forward modelling approach is based on lithospheric stretching assumptions (McKenzie,


1978; Royden and Keen, 1980). The extension factor is used for crustal stretching and for
subcrustal stretching. For the thermal calculations, a 1D numerical finite-difference model was used,
which allows incorporation of finite and multiple stretching phases. In order to handle a large
number of wells and stretching phases in the forward model, a numerical technique is applied (Van
Wees et al., 1996b), which automatically finds best fit stretching parameters for (part of) the
subsidence data. In this procedure the timing and duration of the rift phase must be specified,
whereas best fit stretching values are found by searching for the minimum of the mean square root F
of the deviation in predicted and observed subsidence (Fig. 2.26), as a function of , :

i = num
F ( , ) = (s so ,i )
1 2
p ,i (2.1)
num i =1

where num is the number of subsidence data used in the fitting procedure and sp,i, so,i are
predicted and observed subsidence values respectively. For a rift phase, either uniform lithospheric
stretching ( = ) (McKenzie, 1978) or two-layered stretching can be used ( ) (e.g. Royden and
Keen, 1980). For uniform stretching, the solution of Eq. 1 requires that at least one observed
subsidence data point is given after the onset of rifting, whereas the two layered stretching requires at
least two data points. For polyphase stretching, the fit is accomplished in sequential order. Initially,
using an initial steady state thermal and compositional lithospheric configuration (cf. McKenzie,
1978), stretching parameters of the first phase are determined by fitting data points in the syn-rift and
post- rift time interval up to the onset of the following phase.

54
Chapter 2
A g e (M a )

2
1

2
20 0 10 0 0

t
0

20 0

40 0

60 0

p
sp 80 0
1
s o2 10 00
s o1
s p3
12 00
s p2
s o3 14 00

s
p

16 00

18 00

20 00

su b si d e nce data su b si d e nce data


f i t t e d b y p h a se 1 f i t t e d b y p h a se 2

Figure 2.26: Outline of the forward modelling technique. Explanation in the text (from Van Wees et al., 1996b).

Subsequently, using the perturbed lithosphere configuration predicted at the onset of the second rift
phase, stretching parameters of this rift phase are determined using subsidence data from its syn-rift
and post-rift time intervals up to the next rifting phase.
In the solution procedure it is assumed that values for and can both be found in an
interval from 0.2 to 10. This covers the whole spectrum of realistic parameters for crustal and
subcrustal thickening and thinning. In case of high tectonic uplift or subsidence in this post-rift
interval, it may be difficult to obtain a good fit for the subsidence, since subsidence and uplift rates
are determined by the thermal relaxation of the subcrustal lithosphere. Maximum uplift and
subsidence rates are for = 0.2 and = 10. Such predictions, implying extreme thermal
perturbations of the subcrustal lithosphere, should be interpreted with great care.
u

6.2. MODELLING PROCEDURE

In the fitting procedure initial lithospheric configuration and thermal parameters are adopted
as listed in Table 2.2. To fit the data it was assumed that each observed phase of rapid tectonic
subsidence should correspond with a stretching phase in the forward model. For these phases I
adopted a two-layered stretching model of the lithosphere ( ) in order to obtain the highest
degree of freedom in modelling subsidence data. However, I prefer to use a uniform stretching model
( = ) for those cases where uncertainty exists in estimating stretching factors due to a small number
of data points, or where there are relatively large age uncertainties, as is the case of the Triassic and
Jurassic sedimentary record.
Using the starting and finishing times previously determined for the stretching events, I
calculated the lithosphere stretching factors that would produce theoretical subsidence curves

55
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes
similar to those observed. For the forward modelling, I included for most modelled locations the
complete Mesozoic sedimentary section since the Triassic, even in those columns where the pre-
Mesozoic section is probably deep and does not crop out. In these cases I used thicknesses
interpolated from the isopach maps. In other cases, I only studied the Cretaceous subsidence. As the
lithosphere affected by Triassic-Jurassic rifting would behave differently to those unextended areas, I
applied a reduced crustal and lithosphere thickness in the previously rifted area for the forward
modelling of the Cretaceous subsidence. The reduced crustal and lithosphere thicknesses were
calculated using lithosphere stretching factors calculated from Triassic-Jurassic data.

6.3. MODEL RESULTS: STRETCHING FACTORS

Table 2.4 (a and b) shows the calculated crustal and subcrustal stretching factors for the
different stretching events as interpreted from the subsidence curves for each stratigraphic column.
The forward-modelled tectonic subsidence curves (Fig. 2.27) show a remarkably good fit with the
subsidence data, demonstrating that the minimisation technique is highly efficient. The calculated
lithosphere crustal and subcrustal stretching factors reflect the accuracy level of the database and the
assumptions of the model. Figure 2.27 also shows that better age determinations for the Cretaceous
sedimentary record are reflected in better model results as compared with the Triassic and Jurassic.
The lithosphere, crustal and subcrustal stretching factors calculated for each finite stretching
phase from forward modelling have been plotted in map view (Figs. 2.28 to 2.34).

6.3.1. Mesozoic lithosphere stretching phases

The lithosphere extension factors for each stretching phase are:


Triassic stretching phase. Uniform stretching factors = reach values up to 1.23 in the
Cinega de Morrocoyal, 1.17 in the area of the Upper MV and 1.13 in the western flank of the
palaeo-EC (Fig. 2.28). Spatial distribution of values confirms small narrow (150 km wide) rift basins
with opposite polarity located on opposite sides of the palaeo-Magdalena-La Salina fault sytems.
Triassic to Middle Jurassic stretching phase. The map distribution of stretching values (Fig.
2.29) corroborates two narrow (<150 km wide) rift basins located at the current location of Serrana
de San Lucas (Fig. 2.1) with = values up to 1.09, and inferred stretching factors = in the area
of the Upper MV with values up to 1.12. Lithosphere stretching = values in the western flank of
the palaeo-EC only reach up to 1.02. The abundance of volcanic rocks in these rift basins is
suggestive of a positive thermal anomaly that probably also weakened the lithosphere (as interpreted
in many rift basins, Ziegler, 1994). Lithosphere thermal doming could also have produced the
observed unconformities at the bottom of the syn-rift fill. Lithosphere stretching factors play a
dominant role controlling the degree of upwelling and adiabatic decompression of the asthenosphere
and lower lithosphere. Partial melting occurs when the upwelling material crosses the mantle solidus
line (McKenzie and Bickle, 1988; Wilson, 1993). Fast subsidence rates, en chelon pattern and high
level of volcanic activity are features commonly associated with oblique slip rift zones (Ziegler,
1994). The width of these basins increased compared to early Triassic, suggesting an increasing
width of the thermal weakened lithosphere.
.

56
Chapter 2

C AS C AJAL E S -1 B ETA D E L TA MED I O S B E TA D E L TA

300 250 200 150 100 50 0 300 250 200 150 100 50 0
0 0

500 500
1000 forw ard
forw ard 1000 obse rved
1500
obse rved
1500
2000
2500 2000

AR C AB U C O B E TA D E L TA TAB L AZ O B E TA D E L TA

300 250 200 150 100 50 0 300 250 200 150 100 50 0
0 0
500
500
1000
1000 forw ard forw ard
1500
obse rved obse rved
1500 2000

2000 2500

C H IMA B E TA D E L TA Q U I P I L E B E TA D E L TA

300 250 200 150 100 50 0 300 250 200 150 100 50 0
0 0
200 500
400
forw ard 1000
forw ard
600 obse rved obse rved
800 1500
1000
1200 2000

S U B A-2 B E TA D E L TA V I L L E TA B E TA D E L TA

300 250 200 150 100 50 0 300 250 200 150 100 50 0
0 0
500 500
1000 1000
forw ard 1500
forw ard
1500 obse rved obse rved
2000 2000
2500 2500
3000 3000

YAC O P I B ETA D E L TA S AN L U I S D E G AC E N O B E TA D E L TA

300 250 200 150 100 50 0 300 250 200 150 100 50 0
0 0
500 500
1000 forw ard forw ard
1500 obse rved 1000 obse rved
2000 1500
2500 2000
3000
3500 2500

C AQ U E Z A B E TA D E L TA Q O L I N I B E TA D E L TA

300 250 200 150 100 50 0 300 250 200 150 100 50 0
0 0

500 500
forw ard forw ard
1000 obse rved 1000 obse rved
1500 1500

2000 2000

Figure 2.27: Forward modelled tectonic subsidence (continuous line) and observed tectonic subsidence (dots)
curves (in meters).

57
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes

Q E L C O B R E B E TA D E L TA C O E L L O B E TA D E L TA

300 250 200 150 100 50 0 300 250 200 150 100 50 0
0 0
200 200
400
600 400
forw ard 600
forw ard
800 obse rved obse rved
1000 800
1200
1400 1000
1600 1200

N E I V A B E TA D E L TA P R AD O B E TA D E L TA

300 250 200 150 100 50 0 300 250 200 150 100 50 0
0 0
200 100
400 200
forw ard 300 forw ard
600 obse rved obse rved
800 400
1000 500
1200 600
1400 700

AL P U J AR R A B E TA D E L TA

300 250 200 150 100 50 0


0
200
400
600 forw ard
800 obse rved
1000
1200
1400

Figure 2.27: Continued

Middle Jurassic stretching phase. The distribution of stretching = factor values (Fig.
2.30) indicates rift basins located along the present day western flank of the EC, with = values up
to 1.39. Moreover, assuming that the thickness of the post-rift Cretaceous sections was large along
the western flank of the EC northwest of Bogota, part of the Jurassic-Cretaceous subsidence is likely
to be related to the evolution of this early thermal event. This consideration suggests that these basins
extended southwards into the Cundinamarca region.
Berriasian to Hauterivian stretching phase. Distribution of crustal stretching factors (Fig.
2.31) corresponds very well to the subsidence patterns (Fig. 2.18) and isopach maps (Figs. 2.13,
2.15). They confirm the presence of a (>180 km) wide asymmetrical half-rift basin divided by the
Santander-Floresta horst block. Maximum tectonic subsidence and crustal stretching up to 1.66 was
associated with the pre-Guaicramo master normal fault system at its eastern boundary (Fig. 2.31). A
minor second order half-graben was located in the current location of the western flank of the EC
with crustal stretching values up to 1.45 (Fig. 2.31). Small mafic intrusions described by Fabre and
Delaloye (1983) coincide with areas of thin crust (crustal stretching factors > 1.4) and also with
places of maximum stretching of the subcrustal lithosphere (Fig. 2.31). As a consequence of the
depth dependent lithosphere rheology assumed by the model, results suggest that more intense
stretching affected the subcrustal mantle lithosphere (Fig. 2.32). Differences between crustal and
subcrustal stretching factors suggest some decoupling occurred between the crust and the subcrustal
lithosphere, or that an increased thermal thinning affected the mantle lithosphere. The last
interpretation implies a considerable thermal anomaly produced by mantle lithosphere thinning,
which seems to be supported by the presence of magmatic mafic intrusions.

58
Chapter 2
1.23
1.1 Serrania Maracaibo
Cienaga de Perija Basin
1400 Morrocoyal
d es Barinas
A Cucuta An Basin

cas
B a
rid

Lu
e
M

lt
n

f au
Sa
1.00

lt
Bucaramanga

f au
de

eo
al a
1300 1.11

ae o
nia

ez p
r ra

p al
Se
Arauca

ca
Su a
1.13

a
Bo y
m e
1.00

yst
ra

1200 1.00
Medellin lt s
d ill e

1.05
f au

Llanos Orientales
Co r

1.04
ae o

Basin (LLA)
pa l

Eastern
im a

Cordillera Tunja
1100
B it u

1 .1

Yopal

Manizales 1.00 P a ya n d e San Lu cas


1 .1

lt Terran es (Etayo-Serna et
e fau A al. , 19 8 6). W est ern par at
Ib a gu Bogota of Ch ibch a T e r r an e
em

1000
Ibague 1.00 (Toussaint, 1995).
yst
lt s

1.06 Eastern part of Chibcha


1.16
fau

B Ter ran e ( Tou ssaint ,


Villavicencio 1995)
eo

1.17
and Guyana Shield.
ala

1.00
ap

900 1.09 1.00


usm

1.16
LITHOSPHERE
Ch

0 100 200 Km
l

STRETCHING
1 .1
a
nt r

1.08 FACTOR
Ce

Neiva TRIASSIC
La Macarena

800
Serrania de

San Jose del (248.2 - 235)


Guaviare

800 900 1000 1100 1200 1300 1400 1500


Figure 2.28: Contour map of lithosphere stretching factors (=) calculated through forward modelling for the
Triassic (248.2-235 Ma) stretching event. assuming the hypothesis that there are Triassic sediments in the
Cinaga de Morrocoyal area (Geyer, 1982) and that Triassic sediments were accumulated in the western flank of
the EC: Distribution of main early Mesozoic faults is also shown.

During rifting, stress-induced lithosphere thinning causes adiabatic decompression of the


lower lithosphere and asthenosphere, their partial melting and the diapiric rise of melts into the zone
of thinned lithosphere. Mafic melts appear to be generally derived from an incompatible element-
enriched mantle source residing primarily in the subcrustal lithosphere and/or within mantle plumes,
and from the underlying depleted asthenosphere (Wilson, 1989 in Ziegler, 1994; Wilson, 1993).
Although the 1D model can not predict regional isostatic effects, the Lower Cretaceous unconformity
on the rift margins (e.g. LLA) and locally on horst blocks (e.g. Santander-Floresta palaeo-Massif)
was probably produced by thermal uplift of rift shoulders as suggested by the subcrustal stretching
values. According to Ziegler (1988), Kusznir et al. (1991), Kusznir and Ziegler (1992) and Ziegler
(1994), unconformities on rift shoulders and intra-basinal fault blocks can be attributed to footwall
uplift in response to extensional unloading of the lithosphere. This phenomenon may be enhanced by
thermal uplift of the rift zone and a gradual strain concentration in the axial rift zone. In general
terms, the location of subcrustal and crustal stretched zones coincides, as a consequence of the 1D
model assumption of local isostasy. However, where there is some offset, this is indicative of some
asymmetry in the basin, as also indicated by the general geometry of the basin.

59
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes
Based on subsidence analysis of the Cretaceous on stratigraphic columns of the EC Hbrard (1985)
and Fabre (1987), using the instantaneous stretching model of McKenzie (1978), have calculated
uniform = stretching factors up to 2 for the whole lithosphere.

1.09 Maracaibo
Cienaga
Serrania
Basin d es
An
de Perija
1400 Morrocoyal

1.05
Cucuta a Barinas
id
A er

s
Basin

uca
B M

nL

lt
Sa

f au
de

eo
Bucaramanga
1.06

pala
nia
1300 1.00

lt
rra

f au
rez
Se

Arauca

ae o
Su a
sy st em

1.00
ra

p al
d ill e

ac a
1.00
f au lt
Cor

Bo y
1200 1.02
Medellin
Llanos Orientales
eo

1.00
a pal a

1.00 Basin (LLA)


Eastern
Bitu im

Cordillera
Tunja
1100
Yopal P a ya n d e San Lu cas
Ter ranes (Etayo-Serna et
Manizales A al., 19 86 ). W ester n parat
1.00 of Ch ibch a T e r r an e
lt (Toussaint, 1995).
e fau Bogota
gu Eastern part of Chibcha
1000 I ba 1.00
1.00 B Terr an e (Tou ssain t ,
Ibague 199 5) an d G uyan a
1.00 Shield.
5
1.0
Villavicencio
1.04
1.06
1.00
900 1.12 1.04
LITHOSPHERE
1.1

0 100 200Km STRETCHING


1.00 FACTOR EARLY
JURASSIC
al

Neiva
nt r

(208-185)
La Ma ca rena
Ce

800
Ser ra nia de

San Jose del


Guaviare

800 900 1000 1100 1200 1300 1400 1500

Figure 2.29: Contour map of lithosphere stretching factors (=) calculated through forward modelling for the
Early Jurassic (208-185 Ma) stretching event. Distribution of main early Mesozoic faults is also shown.

However these authors lumped the Cretaceous stretching events in a single instantaneous
stretching event with infinite extension rate. The higher stretching values obtained by these authors is
the logical consequence of lumping several stretching events with finite extension rates in a single
instantaneous event with an infinite extension rate.
Aptian to Early Albian stretching phase. The map view distribution of stretching values
(Figs. 2.33 and 2.34) and thickness maps (Fig. 2.17) indicates that during Barremian and Aptian time
extension took place in the south, in the area of the present day UMV. Crustal stretching factors up to
1.4 are associated with the southern segment of the Cambo fault system and up to 1.2 in the Upper
MV (Fig. 2.33). As a consequence of the depth dependent rheology assumed by the model, results
suggest that stretching affected more strongly the subcrustal mantle lithosphere. Subcrustal stretching
values reach up to 3.24 at the southern western flank of the EC and up to 1.6 at the UMV Fig. 2.34).

60
Chapter 2
Differences between crustal and subcrustal stretching values suggest some decoupling between crust
and subcrustal lithosphere or that an increased thermal thinning affected the mantle lithosphere.
These results would imply a thermal anomaly that probably is responsible of rift shoulder uplift, as
interpreted from fission track data by Van der Wiel (1991) in the UMV and Garzn Massif (Fig. 2.1).
Isostatic adjustment of the crust in response to stretching compensating the thermal uplift within the
rift basin explains the absence of thermal uplift unconformities within the basin.

1.00 Maracaibo
Serrania Basin
de Perija
1400 Cienaga
d es
An
Morrocoyal Cucuta
A B a
id
as 1.1 er Barinas
M Basin
Luc

1.00 Bucaramanga
n

lt
Sa

lt
f au
1300 1.39

f au
de

eo

aeo
nia

pala
Arauca

p al
r ra

rez
Se

ac a
Su a
tem

1.35
Bo y
1.00
s ys

1200 1.00
Medellin
ra

lt

1.09
f au
d il le

1.07
eo
Co r

?
a la

Llanos Orientales
a p

? Tunja Basin (LLA)


ui m

1100
B it

Yopal
? P a ya n d e San L uca s
Manizales 1.00 Ter ranes (Etayo-Ser na et
? A al., 19 86 ). W estern parat
lt of Ch ibch a Ter r ane
fau
gu e Bogota (Toussaint, 1995).
1000 Iba
1.00 Eastern 1.00
Eastern part of Chibcha
m

Ibague Cordillera B
e

Terr an e (Tou ssain t ,


yst

1.00
199 5) an d G uyan a
lt s

Villavicencio Shield.
1.00
u
o fa

1.00
1.00
e

900 1.00
ala

1.00
ap

0 200Km LITHOSPHERE
100
usm

STRETCHING
1.00
Ch

al

FACTOR
nt r

JURASSIC
Ce

Neiva
La Macarena
Serrania de

800 (180.1 - 176)


San Jose del
Guaviare

800 900 1000 1100 1200 1300 1400 1500

Figure 2.30: Contour map of lithosphere stretching factors (=) calculated through forward modelling for the
Jurassic (180.1-176 Ma) stretching event. Distribution of main Early Mesozoic faults is also shown.

6.3.2. Total stretching

Figure 2.35 shows a map of crustal thickness at the end of Cretaceous calculated using the
total amount of lithosphere stretching during Mesozoic time and assuming an initial crustal thickness
of 35 km. This value has been considered as representative of the undisturbed crustal initial rifting
stages, thickness in the LA area (Calvache and Muoz, 1984). No major extensional or compresional
tectonic event has affected the LLA lithosphere during Mesozoic and Cenozoic time. This map
represents the cumulative crustal stretching of the whole Mesozoic extensional history of the EC,
MV and LLA areas.

61
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes

Serrani Maracaib
a o s
de Basin de
An

as
Perija
a

Luc
1400 Cucuta Barinas
1.1 rid
Me Basin

San
1. 2
Bucaramanga

de
1.13 1.28

nia
1.21
1300

ra
1.10 1.05 Arauca

ra
Ser
1.08 1.15 1.00
lle
1.00
1.12 1.08
rd i
1.23
1.09 1.00
1.11 1.35 1.00
Co

1.14
1.66
1.25
Medellin 1.26
1200 1.31
1.48 1 1.00
2
1.00
1.18 1.15
7
8 1.25 1.04 1.35
1.00 1.00
1.24 9 1.10 1.04 1.19 3
6
14 1.33 1.00 1.18
10 to 12 15 A 1.25
1.23 1.21 Tunja 1.00
Llanos Orientales
1.30 4
13 1.22
1.31 Basin (LLA)
1100 1.02
1.18
5 Yopal
1.21 1.00
1.22
Manizales 1.01 1.23
1.20 1.40
B 1.00
1.00
1.00
1.09 1.35 1.00
1.00 1.00
1.15 1.21 1.12
16 1.00
1.00
1.16 1.00
1.00
Bogota 1.00 1.00
1000 Ibague
1.13 1.20
17 1.00 1.00
1.00 1.00
1.15
1.00 1.00 1.00
1.00
1.00 1.00 Villavicencio 1.00
1.00
1.00 1.00 1.00
l

CRUSTAL
a

1.00
1.00
nt r

1.00
900 1.00
STRETCHING
Ce

1.00
1.00 1.00
1.00
0 100 200 Km FACTOR
1.00 1.00 1.00
BERRIASIAN
Neiva Cretaceous mafic HAUTERIVIAN
igneous intrusions
800 (144-127)
1.00
Emerald mines

800 900 1000 1100 1200 1300 1400 1500

Figure 2.31: Contour map of crustal () lithosphere stretching factors calculated through forward modelling for
the Berriasian-Hauterivian (144-127 Ma, Cretaceous) stretching event. Distribution of main Early Cretaceous
faults and mafic intrusions is also shown with circles: 1. Diorite Rio Nuevo, 2. Microgabro Rodrigoque, 3.
Porfiritic basaltic lava, 4. Microgabro Rio Cravo Sur, 5. Pajarito, 6. Q. La Esperanza, 7. Q. Las Palomas, 8. Q. La
Culebra, 9. Marfil, 10. Q. Grande, 11. Q. La Chorrera, 12. La Chunchalita, 13. Q. La Fiebre, 14. Caceres, 15. La
Corona, 16. Pacho, 17. Diorite Rio Guacavia.

7. DISCUSSION

7.1. GEOMETRY OF RIFT BASINS

Triassic rift basins were narrow and increased in width and surface area during Triassic and
Jurassic times. Cretaceous rifts were wider, covered more area and were more asymmetrical than
Triassic-Jurassic rift basins. In many are as it has been observed that during reactivation of

62
Chapter 2
crustal discontinuities can lead to the subsidence of isolated grabens and half grabens that are linked
by shear zones (e.g. east African rift). With increasing strain, such grabens propagate toward each
other, coalesce and evolve into a more or less continuous rift system (Nelson et al., 1992; Ziegler,
1994). According to Ziegler (1994), propagation of established rift systems into previously
unextended areas could occur either by reactivation of pre-existing crustal discontinuities, or through
the development of new fault systems crosscutting the basement grain.

Maracaibo
Serrania
Basin s
de Perija de
An

ucas
1400
id a Barinas
Cucuta er
an L 1.1
M basin
1.2
de S

1.3
Bucaramanga 1.72
1.4

1.
ania

1.63
1300

6
1.37 1.31 1.05 1.5
Serr
a

1.52

ult
Arauca
C or di ller

1.08 1.00
1.30 1.00
eo fa

m
1.26
1.36

ste
1.2

1.06
1.13
pala

1.64

sy
1 1.57
1.
ult
1.1

3.49
5

2.0
rez

fa

3.0
1.

lt
1.3
1.00
Medellin

fau
1.25 1.00
tem

Sua

1200 1.47 1.2 2.53


l ae
sys

1.35

o
pa

lae
1.3

1.59
fault

1.4

1.38

pa
ac

1.15
eo

y
1.
1.1

1.02 1.00

mo
Bo
2

2.70
1.04
pala

1.3 .1

1.10
2.0

ra
1

2.02 2.05 2.07


1.00
ima

1.00
3.15

1.45Tr a 1.25
2.58
aic

1.23
1100 na Tunja 1.17
Bitu

Gu

1.02
sf 2.20 1.00 1.00
Llanos Orientales
3.0

er 1.21 2.0 1.21


Manizales 1.1 pa Yopal 1.00
la e Basin (LLA)
1.1

o 1.00
1.01 1.22 f a 1.03
ul 1.08
1.23 t 1.00 1.00

1.09
1.33 1.1 3.61 1.00 1.00 1.00
1.00
1.7

1000
1.6

1.79
Ibague 1.00 Bogota 1.00
1.2
1. 5

N 1.22
1 .4

1.57
1.00 1.00 az 1.00
a 1.
1

1.00 re 3 1.00
1.

t 1.00
1.00 h 1.00
pa 1.00
la 1.00
1.00
eo 1.00
1.00 1.00 fa 1.00 Villavicencio
ul
t
1.00
1.00 1.00
1.00
900
a l

1.00
1.00 SUBCRUSTAL
nt r

1.00
Ce

0 100 200Km STRETCHING


FACTOR
1.00
1.00 BERRIASIAN
La Macarena

Neiva HAUTERIVIAN
Serrania de

1.00
800 (144-127)
San Jose del
1.00
Guaviare

800 900 1000 1100 1200 1300 1400 1500

Figure 2.32: Contour map of subcrustal () lithosphere stretching factors calculated through forward modelling
for the Berriasian-Hauterivian (144-127 Ma, Cretaceous) stretching event. Distribution of main Early Cretaceous
faults is also shown.

The evolution from asymmetrical Triassic-Jurassic rifts to more asymmetrical Cretaceous


rifts may suggest a transition from a more pure shear rifting during Triassic to a simple shear rifting
during Cretaceous. However Radelli (1967, in Toussaint, 1995b) has pointed out eastward tilted
faulted blocks during Triassic and Jurassic times implying that asymmetrical basins existed. Whether
the Colombian Mesozoic extensional basins were pure shear rifts or simple shear rifts is difficult to
demonstrate. Probably both mechanisms were operating; these rift models should be viewed as end
member cases. Lithosphere extension is localized in zones where the lithosphere is weakest. Such
zones correspond to areas of thermal destabilized lithosphere such as areas of previous rifting or in
areas of crustal thickening in orogenic belts (Ziegler, 1994).

63
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes
Maracaibo
Serrania Basin
de Perija

cas
1400 d es

n Lu
Cucuta
An Barinas
a

e Sa
id Basin
er
M

nia d
Serra
1.00
1.00
1.00
1300 1.04 1.13 1.00 Bucaramanga
1.03

Servita palaeo fault


1.00
1.00 1.00
1.11
1.00 1.00 Arauca
ra 1.09 1.01

m
d il le 1.00
1.00 1.00

ste
1.00
1.00

sy
Co r

tem
1.01

lt
1200 1.02

fau
1.00
Medellin s ys 1.11
1.00

o
1.08

lae
1.00
1.00
lt

pa
1.07
f au

1.00 1.00 1.00


1.01 1.00
1.00 1.04

o
1.00
ae o

m
1.00
1.05
1.00
Llanos Orientales

aic
1.20 1.00
Tunja
pa l

1.32 1.00
1.30 1.00
Basin (LLA)

Gu
Tr 1.00
1100 1.00
Tr
an
s
1.00
i ma

1.00 1.00
an fer 1.00
1.00

s fer
1.18 fau 1.00
lt
Bi tu

1.25
pa Yopal 1.00
la e 1.02 1.00
Manizales 1.49 1.2 o 1.00
1.16 fau 1.00
1.00
1.14 lt 1.00 1.00
1.00 1.00
1.12 1.14 1.00 1.00

. 1 Na
11.05 1.00 1.00
1.00
1000 1.17 z ar Bogota 1.00 1.00
et 1.00
Ibague 1.07 1.16
1.07 h pa
1.00

1.02 1.12 lae 1.00 1.00 1.00


1.00 of 1.00
au 1.00
lt 1.00
1.00
1.00
1.00 1.00

1.28 1.00 Villavicencio 1.00


1.10 1.00 1.00
1.00
1.05 1.00
1.12 1.00
900 1.00
1.22
1.00
CRUSTAL
l a

1.00

0 100 200 Km
nt r

STRETCHING
Ce

1.08
1.00 FACTOR
Neiva APTIAN
La Macarena
Serrania de

1.00
800 (121-102.6)
San Jose del
1.09
Guaviare

800 900 1000 1100 1200 1300 1400 1500

Figure 2.33: Contour map of crustal () lithosphere stretching factors calculated through forward modelling for the
Aptian (121-102.6 Ma, Cretaceous) stretching event. Distribution of main Early Cretaceous faults is also shown.

If the orientation of pre-existing crustal discontinuities is such that they cannot be reactivated by the
stress system governing the evolution of the rift, new faults will develop and pure shear
deformation is likely to prevail (Ziegler, 1990; 1994). This mechanism may be applicable to the
Triassic-Jurassic rift system in Colombia. In contrast, if the upper crust is weakened by the presence
of pre-existing crustal discontinuities with a favourable orientation to be reactivated under the
prevailing tensional stress field, these will present zones of preferential strain concentration, even if
they are located at considerable distances to the sides of the zone of mantle lithosphere stretching
(Ziegler, 1994). This can result in simple shear deformation (Sawyer and Harry, 1991, Harry and
Sawyer, 1992; Ziegler, 1994). This mechanism may explain the development of the palaeo-EC rift
system during the Early Cretaceous. The eastern side of the rift probably developed during
Berriasian-Hauterivian time by reactivation of an older Palaeozoic rift system along the Guaicramo
palaeo-fault (c.f. Hossack et al., 1999). The western side of earlier normal fault systems developed
during Triassic-Jurassic rifting.

64
Chapter 2
The rheological properties of the lithosphere control the depth at which tensional necking
occurs and whether a rift zone is flexed upwards or downward (Braun and Beaumont, 1989; Ziegler,
1994). A deep lithosphere necking level causes upward flexure of the rift zone. Necking at shallow
crustal levels causes downward flexure of the rift zone and absence of shoulder uplifts (Kooi, 1991;
Kooi et al., 1992; Ziegler, 1994). The narrow rifting during Triassic-Jurassic as well as the presence
of unconformities suggest upward flexure and generation of rift shoulders, probably associated with
a deep level of necking. Similar deep levels of necking in the eastern side of the Early Cretaceous rift
system may have generated shoulder uplift in the LLA area during Early Cretaceous. Coarse detrital
fragments in the Lower Cretaceous Brechas de Buenavista Fm. (Pimpirev et al., 1992) and Calizas
del Guavio Fm. (Conglomerado de Miralindo, Ulla and Rodrguez, 1976a) could be derived from
this rift shoulder. In contrast, in the western margin of the Early Cretaceous rift system,
sedimentation was more continuous from Jurassic to Early Cretaceous time implying downward
flexure of the rift shoulders and thus, a shallower level of necking in this western part of the basin
during Early Cretaceous times. In the newly rifted areas (western side during Triassic-Jurassic and
eastern side during Early Cretaceous) necking level was deep, whereas in the previously rifted areas
(western side during Early Cretaceous) necking level was shallow. Therefore, necking started at a
deep level during the first stages of rifting, it then evolved to shallow-necking levels for the more
mature rifting stages. Ziegler (1994) explained that rifting, involving mechanical and thermal
thinning of the lithosphere, is accompanied by a gradual rise in lithospheric isotherms which entails
an upward shift of the lithospheric necking level as well as of the intracrustal brittle/ductile
deformation boundary. According to model experiments, the width of the rift zone depends on the
thickness of the crust and the depth at which the brittle/ductile transition zone is located at the onset
of the lithosphere extension (Allemand and Brun, 1991; Buck, 1991). As this interface rises with
rising isotherms, upper crustal strain concentrates in time on a narrower zone (Sawyer and Harry,
1991; Ziegler, 1994). The observations in the Colombian case are opposite to these modelling results.
Probably the increasing width of the Colombian rift system was the result of progressive tensional
reactivation of old upper crustal weakness zones, which was not considered in those experiments.
On a lithospheric scale the location of rift systems is controlled by the location of weakness
zones in the lithosphere, which in turn depends on its thermal state and the thickness of the crust. At
crustal scales the composition, thickness of its mechanically strong upper layer and the availability of
internal discontinuities which can tensionally be reactivated, are also important controls for the
location of rifts (Ziegler, 1994). The overall pattern of these rift basins for most of Mesozoic time
indicates several rifts NNE-SSW oriented in an en-chelon pattern compared to the more N-S
oriented Central Cordillera (e.g. Mojica et al., 1996). Some authors (Fabre 1987; Sarmiento, 1989;
Geotec, 1992; Mojica et al. 1996) have suggested also that some NW-SE faults probably represented
transfer faults. Some features, such as the Nazareth NW-SE Fault (Fig. 2.31), limiting the Early
Cretaceous basin in the south (Fabre 1987) or the NW-SE alignment connecting the two emerald
districts of the EC (Fig. 2.31, Sarmiento 1989), probably represent Mesozoic transfer faults. If a
subduction-related magmatic arc existed at the current location of the Central Cordillera during
Triassic-Jurassic and possibly (?) Cretaceous times, as has been proposed by many authors (Barrero
and Vesga, 1976 in Toussaint, 1995b; Aspden et al., 1987; Toussaint and Restrepo, 1974a, 1989,
1994), and if the orientation of that magmatic arc and the rifted basins has been preserved, the
oblique orientation of these rifts and its en-chelon pattern would suggest some oblique slip
extension with a left-lateral strike-slip component. However, with the available data it is not possible
to rule-out the hypothesis that some rift arms form acute angles to the dominant NNE-SSW trend in a
pattern similar to aborted aulacogen rifts.

65
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes
For example Mojica et al., (1996) have hypothesized the existence of NW-SE rift segments forming
oblique angles to the dominant NNE-SSW rift segments. Branching of rift systems is of common
occurrence. Most rift basins consist of half-graben depressions, the polarity of which often changes
along trend across accommodation transfer zones (Ziegler, 1994). The latter are characterized by a
complex fault geometry involving local positive and negative flower structures and folding
(Rosendahl, 1987; Morley et al., 1990; Ziegler, 1994). In plan view, master faults of half grabens
often display a curvilinear geometry. This is indicative of their listric configuration whereby their
detachment level corresponds to the crustal brittle-ductile transition zone below which deformation is
dominated by ductile shear (Gibs, 1987, 1989, in Ziegler, 1994). Deviations from such an idealized
fault geometry can be generally related to non-orthogonal extension (Avraham, 1992) or to the
reactivation of pre- existing crustal discontinuities. However, planar faults in rift zones have also
been observed. Mixed planar and listric fault geometries are typical for many rifts (Ziegler, 1994). It
is very difficult to reconstruct the geometry of the Mesozoic extensional faults in the study area. If
the inverse or thrust faults that now define the eastern and western borders of the EC originally were
normal faults that were inverted during the Cenozoic, their geometry in map view would provide
some information about Mesozoic extensional faults. Lateral changes of Mesozoic thickness suggest
that this is the case at least for the master faults that probably defined the regional rift geometry.
Adopting this hypothesis, I have suggested that the Guaicramo, La Salina, Bitima, Magdalena,
Boyac, and Chusma Faults represent original extensional faults. The oblique orientation of most of
them relative to the Mesozoic magmatic arc of the Central Cordillera would imply an oblique slip
extension during the Mesozoic. Analog model experiments of oblique extension produce a similar
map view fault pattern (e.g. Tron and Brun, 1991). However it is important to keep in mind that
during the Mesozoic, modifications in the drift pattern of plates, and consequently their interaction,
may have resulted in changes in the stress regime governing the subsidence and structural style or
development of branches of the rift system. Such interactions can result in transitions from
orthogonal to oblique extension or even to wrench deformation (Ziegler, 1994). NW-SE transfer
faults and possible NW-SE normal faults, as those interpreted by Ecopetrol et al. (1994) in the
Middle MV, were not inverted during the Cenozoic.
Delineation of basin geometry for Triassic and Jurassic times strongly depends on the
assumptions on age, in the absence of accurate age data. For example for the Early Triassic, Mojica
et al. (1996) considered that a single rift basin was located in the area of the Upper MV. This is
based on the idea that the Luisa Fm is the only early Triassic sedimentary record in the whole area. If
one accepts the hypothesis of Geyer (1982), based on facies and stratigraphic position, that
correlative early Triassic deposits exist in the Cinaga de Morrocoyal area, a second rift basin can be
postulated in that area. In general for the Triassic and Jurassic a scenario of at least three separated
rift systems (one in the Upper MV, a second in the area of the western flank of the EC, and a third in
the Serrana de San Lucas) requires testing with new data. Mojica et al. (1996) have proposed several
branches for the early Mesozoic rift system, while Geotec (1992) and Cediel et al. (1997) proposed a
large number of small rift basins. An equally acceptable hypothesis is to assume a single rift system.
Such discrepancy in interpreting the geometry of rift basins is the consequence of a limited number
of outcrops/well data for Lower Mesozoic sedimentary record. However, all the different hypotheses
on the development of early Mesozoic rifting have in common postulated rift basins with an
orientation NNE-SSW.

7.2. RELATIONSHIPS BETWEEN MESOZOIC RIFTING AND MAGMATISM

Two alternative hypotheses have been proposed for the geotectonic setting of Jurassic
volcanism in Colombia:

66
Chapter 2
Serrania Maracaibo
de Perija Basin

d es Barinas
1400
An

cas
Cucuta
Basin
a

n Lu
id
er

e Sa
M

nia d
1.00

1.2
1.00
1.00
1300 Bucaramanga

Serra
1.48 1.00
1.04

1.03

Servit palaeo fault


1.00 1.00
1.33 1.00
Arauca
ra
1.00

m
1.00

1.3

ste
d il le 1.13 1.01
1.00 1.00

lt
1.1

sy
1.00

1.1

fa u
1.00
Co r

lt
fau
1.06

eo
1.06

te m
1200 Medellin 1.33 1.00

o
p a la
1.00

lae
sys
1.37
1.2 1.00

pa
1.00
1.22

re z
1.1

o
1.00
1.5
lt

1.00

am
1.00
fau

1.00 1.07 1.10


1.7

Su a
1.00 1.00 1.00

r
1.00
1.05
Llanos Orientales

aic
eo

3.03.24 1.72
Tunja 1.00
1.00

Gu
1.77
pala

1100 1.00 1.00


1.00
1.00
Basin (LLA)
Tr Tr Yopal
an an 1.00
sfe
im a

sfe 1.00
rp r
fau
ala 1.00 1.00 1.00
Bitu

1.10
Manizales 1.52 2.0 eo lt
1.00 1.00
1.48 fa 1.00
1.42 ult 1.00
1.00
1.30 1.221.58
1.3 1.00 1.00 1.00 1.00
1.00

1.26Na Bogota 1.00


1.00
1.00
1000 1.37
za
re 1.00
1.00
1.00
1.00

Ibague 1.11 1.37 1.47


th
1.3 pa
1.27 1.18 lae 1.00 1.00 1.00 1.00
of 1.00
1.00 au 1.00
1.00
lt Villavicencio 1.00
1.00 1.00

1.66
1.00
1.48 1.00 1.00
1.00 1.00
1.17
1.42 1.00
1.00
900 1.14 1.00
1.2 4

1.00
1.1
1.

1.00
0 100 200Km
SUBCRUSTAL
l a

1.3
nt r

1.00 STRETCHING
Ce

1.31
FACTOR
La Macarena

Neiva APTIAN
Serrania de

1.00
800 San Jose del (121-102.6)
1.18
Guaviare

800 900 1000 1100 1200 1300 1400 1500

Figure 2.34: Contour map of subcrustal () lithosphere stretching factors calculated through forward modelling
for the Aptian (121-102.6 Ma, Cretaceous) stretching event. Distribution of main Early Cretaceous faults is also
shown.

(1) A subduction-related magmatic arc (Toussaint and Restrepo, 1974a; Nez, 1978; Barrero,
1979 among others). Jurassic volcanism was alkaline and related to the development of a
pericratonic arc, implying a specific coupling between the South American and Pacific plates
(Restrepo-Pace, 1995).
(2) Rifting processes related to lithosphere extension (Estrada, 1972; Maca and Mojica, 1981;
Cediel, 1983; Mojica and Maca, 1992; Maca, 1995; Mojica et al., 1996). Bayona et al. (1994)
based on geochemical data suggested back-arc volcanism.

In the study area in contrast to the alkaline Jurassic volcanism nature of Cretaceous magmatism is
different. During the Cretaceous there was still some alkaline volcanism (tuffs and bentonites,
Villamil, 1994; Restrepo-Pace, personal communication), but also intrusion of diabasic-gabroic dikes
with a toleithic affinity (Fabre and Delaloye, 1983; Moreno and Concha, 1993). The latter probably
more related to extension\transtension. The following discussion is based on the assumption that
Mesozoic volcanic rocks in the Mesozoic extensional basins would be related to rifting processes
as suggested by Mojica et al. (1996).

67
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes

Maracaibo
30
Serrania Basin
es

30
uca
1400 de Perija d Barinas
Cucuta
An

nL
Basin
a
id

Sa
er

de
M

nia
35

rr a
30,514 23,460

Se
27,344
1300
28,184
26,132
25
29,200
29,651 Bucaramanga
ille ra
32,37
7 28,153 25 Arauca
31,760 35
25,659
Cor d
19,63 30,310 24,095
7 35
30 28,690
35
30 21,123

Medellin 26,085
1200 25 25
26,889
24,048 22,774
28,780 28,398
25,15 30
20 2 24,386
35 35 35

28,294 21,625 24,179 31,818 25


24,977 29,636 27,999

20,315
23,117
35
12,164 Tunja 24,331
1100 34,044 16,869 29,536
24,352 Llanos Orientales
20 23,632
Yopal
Manizales 22,115
34,014
Basin (LLA)
23,680 35
15,853 35
24,52 29,518
8 24,909 22,439
24,636 35
30
22,508
25
28,229
35

1000 Ibague
29,838 28,664 35

27,649 28,920 25,11


29,636 35
2 35
30,756
24,418 28,504
Villavicencio
30 35
27,259
35
31,818
21,758 35
35
28,928
900 25
35
27,673
al

- 100 200 km
0
nt r

CRUSTAL
Ce

35
THICKNESS
28,419
AT THE END OF
35

Neiva 35
CRETACEOUS
La Macarena

800 30
Serrania de

San Jose del (Km)


32,051
Guaviare

800 900 1000 1100 1200 1300 1400 1500

Figure 2.35: Contour map of crustal thickness (km) at the end of Cretaceous time, calculated using the total
amount of stretching during the Mesozoic and assuming an initial crustal thickness of 35 km before stretching. 35
km is a representative value of the undisturbed crustal thickness in the Llanos Orientales area according to
Calvache and Munoz (1984).

According to McKenzie and Bickle (1988) and Wilson (1993) the amount of lithospheric
stretching plays a dominant role by controlling the degree of upwelling and adiabatic decompression
of the asthenosphere and lower lithosphere. Partial melting occurs when the upwelling material
crosses the mantle solidus line whose position on P-T space is a function of composition. However,
in the study area abundant Late Triassic-Early Jurassic volcanic rocks are associated with moderate
stretching factors ( = up to 1.12) while the Cretaceous sedimentary record is almost devoid of
volcanic rocks (only containing minor mafic intrusions), is associated to higher stretching factors (
up to 3, up to 1.66). Clearly thermal processes were more important compared to mechanical
stretching during Late Triassic-Early Jurassic rifting than during Cretaceous rifting. During the Late
Triassic-Early Jurassic, abundant volcanic rocks suggest a positive thermal anomaly in the
lithosphere but a moderate lithosphere stretching. Triassic-Jurassic unconformities could have been
produced by thermal uplift (active rifting?).
Thermal doming is the result of progressive thinning of the higher density mantle lithosphere
and its replacement by low-density asthenosphere (Bott, 1992). In contrast during Cretaceous time

68
Chapter 2
the much less abundant volcanic rocks, absence of tectonically controlled unconformities, and the
large amount of tectonic subsidence suggest absence of thermal doming. The presence of small mafic
intrusions coinciding with places of maximum crustal and mantle subcrustal stretching suggests
modest magmatism as a consequence of extension of the lithosphere (passive rifting). Subsidence
in rift basins is a consequence of isostatic adjustment of the crust to lithospheric stretching. It tends
to be counteracted by uplift induced by thermal and mechanical attenuation of the subcrustal
lithosphere (Ziegler, 1994). Subsidence patterns and the relative abundance of volcanic rocks suggest
that during Late Triassic-Early Jurassic times thermal controlled uplift counteracting subsidence was
more important than during Cretaceous times. The plutonic bodies of the Santander Massif,
radiometrically dated as Palaeozoic (Goldsmith et al., 1971; Boinet et al., 1985b; Restrepo-Pace,
1995) or Triassic-Jurassic (Goldsmith et al., 1971; Ward et al., 1973; Restrepo-Pace, 1995), could
represent Palaeozoic bodies reheated during Jurassic time, as well as early Mesozoic intrusions. The
thermal anomaly that affected them during Jurassic also may also have generated some degree of
thermal uplift that counteracted subsidence in the less subsiding Santander-Floresta Massif high
block.
Mesozoic rifting in the study area implies that tensional stresses affected the lithosphere.
However probably shear-traction exerted by the upward and outward asthenospheric flow enhanced
these stresses during Triassic-Jurassic times to the degree that crustal discontinuities could have been
tensionally reactivated. In contrast during Cretaceous times the asthenospheric effect was probably
less important and the development of a thermal anomaly and magmatism was mainly favoured by
mechanical stretching in a more passive mode.
Wrench induced pull-apart basins and oblique slip rift zones often display a relatively high
level of volcanic activity (Ziegler, 1994). Wilson and Guiraud (1992) and Ziegler (1994) suggested
that major wrench faults transect the entire lithosphere, thus providing conduits for magma migration
to the surface. If this concept is applicable to the Colombian case, the hypothesis of a single narrow
rift system can be regarded as a broad wrench zone. The hypothesis of several small narrow rift
systems with en- chelon pattern may also support this idea. Toussaint (1995b) reported some
diabase dikes that were controlled by small pull apart structures related to the left-lateral strike-slip
normal faults in the Early Jurassic Saldaa Fm of the Upper MV, which supports this idea.
Volcanic rocks associated with intra-continental rifts display a typically alkaline, mafic/felsic
bi-modal composition (Wilson 1989, in Ziegler, 1994; Wilson, 1993). The abundant Late Triassic-
Early Jurassic volcanic rocks of the study area have variable composition from felsic to mafic.
Chemical analyses of La Quinta Fm. volcanic rocks indicate calc-alkaline composition in the
diagram AFM and alkaline composition in the alkali-silica diagram (Toussaint, 1995b). Chemical
analyses of the Saldaa Fm indicate calc-alkaline composition probably generated in a backarc
environment (Bayona et al., 1994). It is generally accepted that magmatic arcs of calc-alkaline
composition located on continental borders are related to subduction (Aspden et al., 1987). The
predominance of calc-alkaline composition, however, seems to suggest a convergent-related rifting
rather than intra-continental rifting (Toussaint, 1995b). Maze (1984) based on Sr isotopic values
comprised between 0.705 and 0.715 has suggested a partial cortical source for these magmas. In
many rift systems, the mafic melts appear to be generally derived from an incompatible element-
enriched source, residing presumably in the subcrustal lithosphere and/or within the mantle plumes,
and from the underlying depleted asthenosphere (Wilson, 1989 in Ziegler, 1994; Wilson, 1993).
During the evolution of some rifts, decrease in alkalinity of the extruded mafic magmas, and an
increasing contribution of mid ocean-ridge basalt (MORB) source (depleted mantle) melts can be
recognized, both in time and generally toward the rift axis.
This can be attributed to an increasing contribution from melts from the asthenosphere, as the
lithosphere is progressively thinned (Wilson, 1993; Ziegler, 1994). These processes provide an
explanation for the change from felsic composition at bottom to intermediate at top, as reported by
Bayona et al. (1994) for the volcanic rocks of the Jurassic Saldaa Fm. Also these processes explain
the mafic calc-alkaline up to tholeitic composition (Fabre and Delaloye 1983) of the mafic
Cretaceous intrusions located in the central part of the basin where lithosphere was most thinned.
The calculation of stretching factors based on subsidence analysis does not consider flexural effects,
intra-plate stresses or possible crustal density changes. Stretching factors derived from deep crustal
seismic or measurement of crustal extension by faulting from structural sections are generally

69
Mesozoic Rifting History of the Eastern Cordillera, Colombian Andes
smaller than those calculated from subsidence analysis because the model does not take into account
these processes (Ziegler, 1994). Nevertheless this quantitative subsidence analysis contributes
substantially to the understanding of post-rift subsidence processes, providing a measure of the
thermal anomaly that was introduced during the rifting stage of the basin. A clear understanding of
the thermal regime of a basin is of primarily importance for modelling petroleum generation.

7.3. SUGGESTIONS FOR FUTURE STUDIES

It is generally accepted that extensional basins were developed in the area of the EC and MV
during the Mesozoic (e.g. Fabre, 1983a,b, 1987; Cooper et al., 1995). One of the limitations in
Colombia, especially for the continental Triassic and Jurassic poorly fossiliferous sedimentary record
is the scarcity of accurate age data. In some cases deposition age has been inferred from relative
stratigraphic positions with respect to fossiliferous units, or by simple lithological or facies
correlation assuming facies synchroneity. Another limitation results from the controversial ages
proposed by different authors for a single lithostratigraphical unit. Better data are necessary to
properly quantify the early Mesozoic of the EC and MV. Further stratigraphical studies of the early
Mesozoic subsidence history of the EC and MV are strongly recommended. Even for the Cretaceous
marine sedimentary record, which is better dated by biostratigraphy, there is a need for better data.
This Cretaceous record is locally composed of monotonous shale successions disarmonically folded,
where thickness are difficult to measure and structural repetitions difficult to detect, in the absence of
a high number of age control points (Restrepo-Pace, 1989). Better data are necessary to accurately
quantify the Mesozoic subsidence. It is extremely important to understand basin subsidence during
Mesozoic time to understand the Cenozoic tectonic evolution, because Mesozoic rifting weakened
the lithosphere and created weakness zones that strongly affected later deformation and basin
formation processes (see Chapters 3 and 4). In addition Mesozoic rifting provided heat to start
organic matter maturation and hydrocarbon generation processes, which are a key element for
hydrocarbon exploration.

8. CONCLUSIONS

The high-resolution backstripping analysis and forward modelling showed that Mesozoic
Colombian basin is marked by five rifting pulses. Periods of rift activity correlate in time with gaps
of subduction-related magmatic arc activity as suggested by Aspden et al. (1987) especially for
Jurassic time, supporting the hypothesis of backarc extension. If backarc extension continued during
the Early Cretaceous by oblique plate convergence, it probably has a strong strike-slip component, as
suggested by Aspden et al. (1987). Evidence supporting the hypothesis of a backarc basin located
behind a partially emerged less subsiding palaeo-Central Cordillera (magmatic arc?) could be
summarised as follows: (1) The presence, in the western part of the Cundinamarca sub-basin, of
Lower Cretaceous sandstones with abundant volcanic lithic fragments and feldspar derived from a
western detrital source area as indicated by palaeocurrent data (Murca Fm. And tica Sandstone;
Sarmiento, 1989; Moreno, 1990b, 1991); (2) The presence of progressive westerly onlap
terminations of the Cretaceous carbonates on the basement, observed in seismic lines, in the western
border of the Cesar Valley, in northern Colombia (Fig. 2.1; Audemard, 1991); (3) Petrographical
evidence suggesting that Berriasian (?) to Valanginian clastic sediments near San Felix in the
western flank of the Central Cordillera came from erosion of nearly uplifted areas containing
fragments of metamorphic rocks and small tectonic blocks with plutonic rocks (Rodrguez and
Rojas, 1985). (4) Cretaceous volaniclastic rocks (Rodrguez and Rojas, 1985; Rubiano, 1989;
Villamil and Arango, 1998) that were probably also derived from a magmatic arc; (5) Late
Cretaceous zircon fission track ages in the Central Cordillera (Gomez et al, 1999; Toro et al, 1999);
(6) Some Cretaceous plutonic bodies in the Central Cordillera. However, the hypotheses of a passive
margin (Pindell and Erikson, 1993) or aborted rift arms related to the break-up of Pangaea (Geotec,
1992) cannot be completely ruled out due to the absence of a well-defined Cretaceous magmatic arc.
Three stretching events are suggested during Triassic-Jurassic time. Spatial distribution of values
suggest that small narrow (<150 km wide) asymmetric rift basins were located on opposite sides of
the palaeo-Magdalena-La Salina fault system, which probably was active as a master normal fault

70
Chapter 2
system. During Berriasian-Hauterivian subsidence curves and isopach maps suggest that a (>180 km)
wide asymmetrical half-rift basin existed divided by the Santander Floresta high block. In the south
there was a single depocenter limited southward by a vertical transfer fault. Location of small mafic
intrusions (Fabre and Delaloye, 1983) coincides with areas of thin crust (crustal stretching factors >
1.4) and also with places of maximum stretching of the subcrustal lithosphere. During Aptian to early
Albian time, the basin extended towards the south in the UMV. Differences between crustal and
subcrustal stretching values suggest either that there was some lowermost crustal decoupling
between crust and subcrustal lithosphere, or that an increased thermal thinning affected the mantle
lithosphere.
Triassic rift basins were narrow and increased in width during Triassic and Jurassic times.
Cretaceous rifts were wider, and were more asymmetrical than Triassic-Jurassic rift basins. During
Berriasian-Hauterivian time the eastern side of the rift possibly was developed by reactivation of an
older Palaeozoic rift system associated to the Guaicramo fault system (c.f. Hossack, et al., 1999).
The western side probably developed by reactivation of an earlier normal fault system developed
during Triassic-Jurassic rifting. During the first stages of rifting, lithosphere necking started at a deep
level, and evolved to shallower necking levels in the latter rifting stages. Probably the increasing
width of the rift system was the result of progressive tensional reactivation of pre-existing upper
crustal weakness zones. Lateral changes of Mesozoic sediment thickness suggest that the reverse or
thrust faults that now define the eastern and western borders of the EC were originally normal faults
inverted during the Cenozoic Andean orogeny. Thus the Guaicramo, La Salina, Bitima,
Magdalena, Boyac, and Chusma were originally extensional faults. The oblique orientation of most
of them relative to the Mesozoic magmatic arc of the Central Cordillera may be the result of oblique
slip extension during Mesozoic, or alternatively can be inherited from the pre-Mesozoic structural
grain. However, not all the Mesozoic extensional faults were inverted (e.g. NW-SE in the Middle
MV), some normal faults were passively transported with short-cut basement blocks during Cenozoic
inversion (e.g. Esmeraldas Fault, ESRI and Ecopetrol, 1994, Cooper et al., 1995).
Thermal processes were more dominant than mechanical stretching during Late Triassic-
Early Jurassic phase than during Cretaceous rifting phase. During Late Triassic-Early Jurassic
abundant volcaniclastic rocks suggest a positive thermal anomaly in the lithosphere but a moderate
lithosphere stretching. Triassic-Jurassic age unconformities could have been produced by thermal
uplift (active rifting?). In contrast during the Cretaceous less abundant volcanic rocks, absence of
tectonically controlled unconformities, and the large amount of tectonic subsidence indicates absence
of thermal doming. The presence of minor mafic intrusions coinciding with places of maximum
crustal and mantle subcrustal stretching, suggest that a modest magmatism took place as a
consequence of extension of the lithosphere (passive rifting).

71
CHAPTER 3

PALAEOGENE INCIPIENT BASIN INVERSION HISTORY OF THE EASTERN


CORDILLERA, COLOMBIAN ANDES

1. INTRODUCTION

Inverted extensional basins and upthrust basement blocks are common intraplate
compressional/transpressional structures associated with continental plate margins (Ziegler et al.,
1998). Such compressional features at plate margins commonly result from collision-related orogenic
processes. Compressional/transpressional stresses that are related to collisional plate interaction are
responsible for inversion of tensional hanging-wall basins (Ziegler et al., 1995). Compression in the
regions behind a magmatic arc is associated with Andean type orogens and occurs during periods of
increased convergence rates between the subducting and overriding plates (Ziegler et al., 1998).
Inversion of tensional hanging-wall rift basins located behind a magmatic arc is the result of
acceleration of convergence rates between the colliding plates, their increased mechanical coupling and
the transmission of compressional stresses into the backarc domain of the overriding plate (Uyeda and
McCabe, 1983; Ziegler, 1993; Ziegler et al., 1998).
The Palaeogene plate tectonic history of Colombia seems to provide an appropriate tectonic
setting for inversion of Mesozoic extensional basins. Some authors (e.g. Van der Hammen, 1961;
Roeder and Chamberlain, 1995; Restrepo-Pace et al., 1999a,b; Gomez et al., 1999) have suggested the
possibility that upthrusted blocks and/or incipient inversion of Mesozoic extensional basins in the area
of the EC (Fig. 3.1) during Palaeogene created some topography. However, the most commonly
accepted view is that of a unique simple foreland basin related to the topographic load of the Central
Cordillera (e.g. Cooper et al., 1995, Fig. 1.2). Typical palaeogeographic interpretations show the area
east of the Central Cordillera as a single basin where sedimentation covered its entire surface during
early Tertiary times. Was the area east of the Central Cordillera (MV, EC and LLA) a single simple
foreland basin (Cooper et al., 1995, Fig. 1.2) or a compartmentalized basin with local
palaeotopographic emergent areas resulting from incipient inversion of Mesozoic extensional basins
(Fig. 3.1, Gomez et al., 1999) during Palaeogene time?
The aim of this Chapter is to constrain alternative possible tectonic scenarios for the
Palaeogene tectonic history of the EC. To this purpose, local detailed geological evidence is compiled
into a regional geological model and quantitative modelling of tectonic thermal and flexural subsidence
is carried out to test these different tectonic scenarios. Additionally this chapter contributes to
understanding of the Palaeogene tectonic basin forming and inversion mechanisms in terms of
geodynamic processes that govern deformation of the lithosphere. I attempt to put the Palaeogene
tectonics in the study area in a plate-tectonics framework addressing the mechanical control on the
tectonic processes, such as the role of stresses, lithosphere rheology, and pre-existing crustal
discontinuities. In this Chapter and subsequent chapters I discussed flexural subsidence of lithosphere
produced by topography generated by uplift. In order to discusse uplift I use the terminology proposed
by England and Molnar (1990) as defined in equation (3.1). Surface uplift (H) is the upward
displacement of earths surface with respect to the geoid. Uplift of rocks (U) is the upward
displacement of the rocks with respect to the geoid. Exhumation (E) is the upward displacement of
rocks with respect to the earths surface. Exhumation and denudation are both the same, but have
opposite signs; denudation is the downward movement of earths surface relative to a fixed rock
volume. These variables are related:
H = U E (3.1)
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera, Colombian Andes

JURASSIC

Palaeo-Magdalena Magdalena-Tablazo Santander-Floresta


Valley sub-basin palaeo-high

63 km

(from average Cretaceous


crustal stretching factor
1.2)
Moho

CRETACEOUS

Palaeo-Magdalena Magdalena-Tablazo Santander-Floresta Cocuy Llanos


Valley sub-basin palaeo-high sub-basin Orientales

Moho

PALAEOGENE
Palaeo-Central
Cordillera
Palaeo-Magdalena Magdalena-Tablazo Tunja Axial Cocuy Llanos
Valley sub-basin Region sub-basin Orientales

Moho
10 km
(from this Chapter )
NEOGENE

Magdalena Inverted Magdalena-Tablazo Tunja Inverted Cocuy sub-Llanos


Valley sub-basin Axial Region basin Orientales

105 km

(from Colletta et al. 1990)

0 100 km Moho

Approximate horizontal and vertical scale

Figure 3.1: Cartoon showing the suggested tectonic evolution of the Eastern Cordillera since Mesozoic time. Results
of this chapter indicate an incipient inversion of Mesozoic basins and generation of local topography within the
study area (e.g. Gmez et al., 1999). This interpretation contrasts with the hypothesis of a single foreland basin
during Palaeogene time (e.g. Cooper et al., 1995). Neogene and Cretaceous sections are the balanced and restored
sections from Colletta et al. (1990); however, sections are not balanced at depth. Cretaceous shortening estimated
from Cretaceous extension factors (Chapter 2), Palaeogene shortening estimated from kinematic modelling (section
11 of this chapter) and Neogene shortening from Colletta et al. (1990).

No subsidence studies or flexural models have been done for the Palaeogene of the study area. Several
hypotheses about tectonic vertical movements are tested through 1D-thermal subsidence modelling and
2D-flexural modelling to explain the observed tectonic subsidence. An extensive data set of more than

73
Chapter 3
100 stratigraphic columns and wells from the EC, MV and LLA areas (Fig. 3.2) from literature and
well data from Ecopetrol was used. Because lithosphere stretching generated the basin during the
Mesozoic, the first hypothesis tested was whether some remaining thermal subsidence after Mesozoic
stretching affected the Palaeogene evolution, which is a similar scenario to the Aquitanian Basin, that
was quantitatively tested by Desgaulx et al. (1991). The second hypothesis tested was lithosphere
flexure produced by a topographic load in the palaeo-Central Cordillera, as proposed in literature (e.g.
Cooper et al., 1995). The third hypothesis tested was flexural subsidence produced by the topography
generated by some local uplifted blocks within the study area. This last hypothesis was considered by
taking into account evidence of active deformation, slight inversion (Gomez et al., 1999) and surface-
uplift of hanging-wall blocks adjacent to Mesozoic extensional faults during the Palaeogene. Although
the observed subsidence could not be explained by any individual hypothesis, a combination of all
effects gave a satisfactory result. The observed local flexural component of subsidence is interpreted as
having been produced by surface-uplift of local faulted blocks. To better understand relationships
between vertical and horizontal movements, horizontal compression movements responsible for rock-
uplift of local faulted blocks were calculated through forward modelling of kinematics, structural
deformation, sedimentation and erosion assuming an incipient inversion of the Mesozoic extensional
basin during Palaeogene (Sassi et al., 1998). This model allowed an estimate to be made of the amount
of basin inversion and shortening during Palaeogene.

2. TECTONIC SETTING

2.1. PLATE-TECTONIC INTERPRETATIONS

2.1.1. Latest Cretaceous and Tertiary

During latest Cretaceous (since Campanian) and Tertiary all published plate-tectonic
interpretations (Pindell and Dewey, 1982; Burke et al., 1984; Duncan and Hardgraves, 1984; Ross and
Scotese, 1988; Jaillard et al., 1990; Pindell and Barret, 1990; Pindell, 1993; Pindell and Erikson, 1993;
Meschede and Frisch, 1998) propose that a convergent margin existed west of Colombia. During latest
Cretaceous to Oligocene the Caribbean plate was moving eastward relative to South America, while the
Farallon Plate was subducting west of Southern Colombia (Pindell and Erikson, 1993 and Pindell and
Tabut, 1995).
Late Cretaceous-Early Paleocene. There is agreement about the subduction of a south-western
portion of a Caribbean or proto-Caribbean oceanic plate west of the Central Cordillera and obduction-
accretion of oceanic terranes (Calima Terrane according to the terminology of Toussaint and Restrepo,
1989; Toussaint, 1995a,b) to form the Western Cordillera (McCourt et al., 1984). Caribbean collision
with north-western South America was diachronous, becoming younger northward and eastward:
Cenomanian-Campanian in Ecuador and Campanian-Maastrichtian in Colombia (Pindell and Erikson,
1993 and Pindell and Tabut, 1995). According to Cooper et al., (1995) deformation was restricted to
the Western and Central Cordilleras except for initial rock-uplift of the Santander Massif (Fabre, 1987).
The amount of compressional deformation generated during accretion may have been limited by the
oblique convergence of the Nazca and South American plates until 49 Ma (Pardo Casas and Molnar,
1987). According to Cooper et al. (1995) an early-Pre-Andean foreland basin was developed in the area
of the EC, MV and LLA.
Eocene. An increase in rate of Caribbean-South American plate convergence between 49 and
42 Ma (Daly, 1989) elevated the Central Cordillera and produced folding and thrusting in the Middle
MV. These structures are truncated and unconformably overlain by upper Eocene clastics (Morales and

74
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera, Colombian Andes

Serrania

as
Maracaibo

und ary
de Perija

Lu c
Basin

es
d
An
1400

S an
la te bo
Cucuta

i da
Barinas
Basin

r
e

Me
ia d
(F re e p

ran
Fig.3.15

S er
Fig 0 100 200 km
.3 . b a
1300 1 Bucaramanga

Arauca
1

Llanos Orientales
ra

1200 Medellin
ll e

Basin (LLA)
rdi
Co

2
Tunja
1100
Yopal
Manizales

3
Bogota
1000 Fi
re

Ibague g.
3. b
Su tu

14 a
Fi
g.
e ra l

3 .9
Ro m

Villavicencio

900 Fig.3.8
5
al
nt r

Neiva Guyana Shield


Ce

La Macaren a
Serrania d e

800
San Jose del Guaviare 8

1
2
800 900 1000 1100 1200 1300 1400 1500

Figure 3.2: Location of stratigraphic columns, wells used, and location of lithosphere flexure 2D model sections.
Numbers along sections refer to labelling of stratigraphic transversal sections (Figs. 2.7, 2.8 and 2.9), and also to
labeling of lithosphere flexure 2D models (Table 2.1; Figs. 3.26 to 3.33). The Romeral suture is the westernmost
boundary of continental crust. It was used as a free plate boundary for the 2D flexural models

the Colombian Petroleum Industry, 1956). The Eocene unconformity is regional in eastern Colombia.
According to Cooper et al. (1995) a late-Pre-Andean foreland basin developed in the area of the EC,
MV and LLA from Middle Eocene to Early Miocene time.
Latest Eocene and early Oligocene. The rate of plate convergence was reduced, from about 2
cm/yr to 1 cm yr. Convergence mainly was driven by South America moving westward toward a nearly
stationary Caribbean Plate, referred to a mantle reference frame. Sedimentation reassumed in eastern
Colombia possibly by relaxation of the greater compressive stresses in early and middle Eocene times
(Pindell et al., 1997).

75
Chapter 3
Late Oligocene-Early Miocene. The Farallon Plate was broken into the Cocos and Nazca Plates
during Oligocene at about 25 Ma ago (Wortel and Cloetingh, 1981; Duncan and Hardgraves, 1984).
According to Cooper et al. (1995), changes in plate-tectonic motions (Pilger, 1984; Avraham and Nur,
1987) did not cause any deformation in the EC or the LLA. Deformation of this age has been described
from the Cauca Valley (Alfonso et al., 1989) and the MV where the reactivation of the middle Eocene
structures created an upper Oligocene unconformity (Schamel, 1991).
Middle Miocene. Collision and accretion of the Cuna Terrane (Serrana de Baud in NW
Colombia) with the northwestern margin of South America occurred during the Middle Miocene
(Duque-Caro, 1990). According to Cooper et al. (1995) the collision may have contributed to loading
and may have initiated deformation in the EC. However Kerr et al. (1997), based on the age of the
Dabeiba subduction-related magmatic arc, suggested that obduction and accretion of the volcanic rocks
of the Serrana de Baud (Cuna Terrane) probably occurred during late Eocene. A similar westward
shift of the subduction zone during Eocene has been proposed by Feininger (1980, 1986) in Ecuador.

3. STRATIGRAPHY

In this section I summarise the Palaeogene stratigraphy (Figures 2.7 to 2.11). Because the
Palaeogene tectonic history of the EC is not well constrained, I also present in this section
stratigraphical (e.g. unconformities revealing local erosion), sedimentological (e.g. palaeocurrent data
suggesting location of detrital provenance areas) and petrographical (mineralogical detrital sandstone
composition suggesting detrital provenance areas) evidence, useful to constrain the Palaeogene tectonic
history.

3.1. LATE MAASTRICHTIAN-EARLY PALEOCENE

According to Cooper et al. (1995) the final accretion of the Western Cordillera (Figs. 1.1 and
1.2) resulted in a change from marine to the non-marine deposition in the pre-Andean foreland basin
(Figs. 2.7 to 2.11 and Figs. 3.3 and 3.4). During late Maastrichtian-early Paleocene, paralic coastal
plain and alluvial plain claystones containing coal beds of the Guaduas Fm were deposited (Laverde,
1979; Fabre, 1985a; Sarmiento, 1992, 1993). The lower limit of the Guaduas Fm is defined by a
generalized marine flooding surface (Sarmiento, 1993), and upwards a general regression and change
from coastal marine environments to alluvial plains is recorded. To the northward facies are more
marine (Catatumbo Fm, Fabre, 1985a; Sarmiento, 1993). Sarmiento (1993) recognized four smaller
sequences within the Guaduas Fm. Each sequence starts with a marine flooding surface and records a
change from coastal marine environment to alluvial plains. Based on thickness changes (Fig. 3.3)
Sarmiento (1993) suggested that faults controlled the sedimentation, with the most subsiding blocks
located in the western flank of the Cordillera. The Fm is limited on top by an unconformity (Sarmiento,
1993), which is interpreted as an erosional truncation increasing in effect eastward (Figs. 2.7 to 2.11).
In the LLA Foothills (Figs. 2.1), upper Maastrichtian-lower Paleocene rocks are not present (Cooper et
al., 1995). In the Upper MV (Fig. 2.1) mudstones and sandstones (Seca Fm, Guaduala Fm) were
deposited by meandering streams in a delta plain environment during the marine regression (Gmez
and Pedraza, 1994). The basin was completely full of sediment by the end of early Paleocene
(Sarmiento, 1993). In the EC the Guaduas Fm is conformable over the Guadalupe Gp or equivalent
units. However, Guillande (1988) reported that locally in the Upper MV the Guaduas Fm. rests
unconformably on rocks as old as Jurassic to Upper Cretaceous, suggesting a Late Cretaceous
deformation event in this area.

76
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera ,Colombian Andes

Serrania Maracaibo
de Basin
Perija
1400 s
de

Lucas
Cucuta An Barinas
a

de San
e rid Basin
M

ia
Serran
>550 >800
1300 .
Bucaramanga
.
942.
100
0. 0. 0. Arauca

lt
vi f au
ra 374.
il l e a
>250 0.

eo
.

Ser
a
rd

320.
la

pal

u lt
1200
Co
pa

fa
ez

eo
eo
0. Llanos Orientales

r
390.

al a
Su a

la
Medellin

pa
Basin (LLA)

op
0.

m
ca
360.

am
530.

s te
ya
>242. 0. 0.

r
Bo

sy
0 665. 575.
200

a ic
660.

ul
Tunja

Gu

fa
450. 0. 0.
1100 309.
850.
810.
817.
0 559.
>100.
540. 0. 0.
140.
1400. >345 Yopal 0.
.565.
0.
1150.
Manizales 0 760.
100 >258 130.
. 130. 0.
657. 90.
Bogota
1000 198. 458.
Ibague 510. 0. 0.
>40 >65 478.
.
>50. >107 0. 0.
. . 400. 0. 0.
>45 0.
. 0. 0. 0.
133. 0.
>144 0. Villavicencio 0.
>115 . 0.
. 131. 0.
900 >61
.
0.
0.
70. 0. 0. MAASTRICHTIAN
0. -
EARLY
l

0.
a

PALEOCENE
nt r

0.
335.
Ce

THICKNESS
800 Neiva
La Macarena

0 100 200 km
Serrania de

(metres)
30.

800 900 1000 1100 1200 1300 1400 1500

Figure 3.3: Maastrichtian-early Paleocene restored thickness (meters) without palinspastic restoration. Large bold
numbers indicate palaeotopography in meters estimated from 2D flexural modelling. Thick lines represent palaeo-
faults believed to have been active during early Paleocene time.

The cycle ended with a local unconformity due to slight exhumation of some areas, probably related to
the initial deformation of the sedimentary pile. Subtle exhumation and erosion resulted from the initial
inversion of the normal faults at the borders of the basin during early Paleocene (Fajardo-Pea, 1998).
Evidence of this deformation is the unconformity at the top of the Guaduas Fm.
Detrital source areas. During Maastrichtian time rocks of the Central Cordillera (Fig. 1.1)
began to uplift and supply detrital sediments in some places along the western margin of the basin
(quartzite, chert pebbles in the Cimarrona Fm, Gmez and Pedraza, 1994). Also the mineralogical
composition of sandstone in the Upper MV (metamorphic fragments in the sandstones of Monserrate
Fm, Waddell, 1982, and in western Venezuela, Van Andel, 1958) indicate exhumation of the Central
Cordillera. Although most of the Maastrichtian-early Paleocene Guaduas Fm sandstones are quartz-
rich, in its upper part, sandstones are texturally and compositionally immature, having also been
derived from erosion of the Central Cordillera (Sarmiento, 1993).

77
Chapter 3

Maracaibo

Sa n ta
Basin

s
s

n Luca

n de r
1400 Cucuta de
An Barinas

M as s
ia de Sa
ida Basin
er
M

if
High
Serran

nv
Bucaramanga

chira
illera
1300

Ca
Cord

lt
a
lle

f au
pal S alin
Arauca

Va
a le l e

aeo
gdrv i d d
na
l

La
entra
ta

ult
aul
i

fa

n os Se
f
o C

aeo

o
Ma
1200 Medellin

l ae
sin
b- o
Su laz

o
Ba
al

pa
na Tab
zp
Pale

io m
ca

eg a
ale d
ar e

R S og
gd rte

ya
Llanos Orientales

lls
Ma Inve

Bo
Su

thi
a
Basin (LLA)

nj

oo
Tu

sF
no
em
a Tunja st

Lla
1100 t
go sy
Bo sin ul Yopal
Manizales Ba fa
de u b- o
a S l ae
an cu
y
pa
em

sin

b
Sub onda

Sa Co o
-Ba
s ys t

v m PALEOCENE
In ra
H

Bogota a
ic
a u lt

1000 Ibague ua PALEOGEOGRAPHY


G
eo f

Positive relief.
Ba t
b- do
p ala

sin

sif
Su irar a

Alluvial fan and fluvial


as
M
ey
G
sm a

Coastal plain predominantly


am
Vall

Villavicencio
sandstones
et
Chu

Qu
a
alen

900 Coastal plain predominantly


mudstones
agd

0 100 200 km Littoral to inner shelf


M

sandstones
per

Shallow marine inner shelf


Up

Neiva carbonates
Shallow marine inner shelf
mudstones and siltstones
La Macarena

800
Serrania de

Outer shelf shales of


in
b-B a
Su eiv

San Jose del


as

carbonates
N

Guaviare Turbiditic sandstones

800 900 1000 1100 1200 1300 1400 1500


Figure 3.4: Paleocene palaeogeography without palinspastic restoration (modified from Geotec, 1992; Cooper et al.,
1995 and Villamil, 1999).

3.2. LATE PALEOCENE

During late Paleocene time in the LLA region and EC (Fig. 1.1) sedimentation started after a
period of erosion (Figs. 2.7 to 2.11, Sarmiento, 1993). Abrupt changes of thickness of the Guaduas Fm
(e.g. across the Soapaga Fault with a thinner section in the eastern side, Fig. 3.5) suggest some faults
moved during the time represented by the unconformity (Cspedes and Pea, 1995). In the LLA
foothills (Figs. 2.1 and 3.4, near Paz de Ariporo) locally the unconformity has been described as
angular (Vanegas and Arango, 1994). In the Middle MV (Figs. 2.1 and 3.4) an onlap relation of
Paleocene deposits over older rocks can be recognized on seismic lines (Ecopetrol-ICP, 1996) and
might be the expression of the unconformity. Upper Paleocene deposits form a sequence bounded by
unconformities. The upper boundary is an unconformity that represents the earlymiddle? Eocene
recognized in Eastern Colombia (Figs. 2.7 to 2.11). In the LLA (Fig. 1.1) this sequence extends farther
east, possibly due to transgression and early flexural loading of the Central and Western Cordilleras
(Cooper et al., 1995). The lower basal transgressive system tract is represented in the LLA by mature,
sandstone-rich, estuarine deposits (Barco Fm, Cooper et al., 1995).

78
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera, Colombian Andes

Serrania Maracaibo
de Perija

s
Basin

n Luca
1400
s
de

ia de Sa
Cucuta
An
Barinas
ida
er

Serran
Basin
M
500
1100
1300
Bucaramanga
100
314 0

ra
964 Arauca
100
d il le
>223 0
458 0
100 1007
834
Cor

1200 0
Medellin 951
100

7
82
24
859

0 560 0
0
100 904
666 525 >387
Llanos Orientales
708 610 560 0
0
100 70? 480 >443 Basin (LLA)
1100 0 Tunja 708 105 0
76
200 1170
50 270
54
660
Manizales 0 56? 109
Yopal 0
950 96 0
29 24
1385 95 12

100 3617
200 280
530
180 72 0 0
0
200 16
0
0 0 0
1000 155
Bogota 1500 47 0
Ibague 145
0

68 0 0 0
>173
31
0 0
0
77 0
>58 Villavicencio 0
75 60? 0
0
900
tr al

0
00
C en

0
PALEOCENE
0
415 THICKNESS
Neiva
La Macarena

0
800
Serrania de

(meters)
0 0 100 200 km

800 900 1000 1100 1200 1300 1400 1500

Figure 3.5: Paleocene restored thickness (meters) without palinspastic restoration. Large bold numbers indicate
palaeotopography in meters estimated from 2D flexural modelling. Thick lines represent palaeo-faults believed to
have been active during Paleocene time.

In the EC this lower part is represented by braided fluvial sandstones (Thanetian age, Sarmiento in
Jaramillo et al., 1993; Arenisca del Cacho and Socha Inferior Fms, Cspedes and Pea, 1995), and in
the Middle MV by a sandstone stratigraphical interval (basal part of the Lisama Fm considered by
Bueno, 1971 in Mora et al., 1996) to be lateral equivalent of the Barco Fm in the Catatumbo area).
Marine influence is strong in the LLA foothills (Cooper et al., 1995) while in the EC a deltaic coastal
plain with marine ichnofauna represents a maximum flooding surface (lower part of Socha Superior
Fm, Cspedes and Pea, 1995). The transgressive system tract is represented in the LLA by the
transition from estuarine deposits and the highstand system tract is represented by an upward transition
to coastal (lower Los Cuervos Fm) and alluvial plain mud-rich deposits (Cooper et al., 1995).
Sandstone deposition ended when a relative tectono-eustatic level highstand was established (Cooper et
al., 1995). Coarse clastics appear to have bypassed the LLA foothills and EC, where a regressive mud

79
Chapter 3
dominated coastal (Cooper et al., 1995) to alluvial plain with high sinuosity streams was established
(Bogota Fm, Hoorn, 1988; upper part Socha Superior Fm, Cspedes and Pea, 1995). In the Middle
MV shallow marine to transitional variegated mudstone, minor sandstone and thin beds of coal in
coarsening upward successions are interpreted as produced by decrease of the accumulation
space/sediment supply ratio (Ecopetrol-ICP, 1996). Northward in the Sierra Nevada del Cocuy the
environment was more marine (Barco Fm with littoral environment, Fabre, 1986).
For the LLA area, Fajardo et al. (1993) have proposed that this sequence bounded by
unconformities represents incised valley fills deposited during a base-level rise trend formed by four
smaller base-level cycles. The lower part consists of fluvial sandstones (Barco Fm) overlain by
transitional deltaic(?) varying to fluvial deposits eastward (Los Cuervos Fm).
Thickness variations (Fig. 3.5) within this sequence are interpreted by Cooper et al. (1995) to
result from extension on the Cusiana-Tmara fault system in the LLA foothills. In the EC also changes
in thickness have been recognized (Cspedes and Pea, 1995).
Detrital source areas. Upper Paleocene sandstones (Arenisca del Morro, Barco, Socha Inferior
and Arenisca del Cacho Fms) are quartzarenites (Jaramillo et al., 1993; Cooper et al., 1995,),
litharenites, sublitharenites (Jaramillo et al., 1993, Barrientos and Torres, 1994) and subarkoses
(Jaramillo et al., 1993). Sandstones contain quartz similar to that of Upper Cretaceous Guadalupe Gp
(Barrientos and Torres, 1994), chert fragments (Notestein et al., 1944; Aalto, 1971; Vsquez, 1983;
Hathon and Espejo, 1997; some of them derived from La Luna Fm, Fabre, 1986) and low-grade
metamorphic clasts (Hathon and Espejo, 1997). Palaeocurent data from these sandstones are: (1)
Northward to northwestward (Jaramillo et al., 1993, Barrientos and Torres, 1994) in the LLA foothills
(Fig. 3.4); (2) Southeastward in the northwestern axial region of the EC (Fig. 3.4, Jaramillo et al.,
1993); (3) Northeastward in the Sabana de Bogot area (Fig. 3.4, Jaramillo et al., 1993); (4) Northward
in the Tunja Paz de Ro area (Fig. 3.4, Cspedes and Pea, 1995); (5) Toward the north in the Tunja-
Paz de Ro area east of the Soapaga Fault (Cspedes and Pea, 1995), (6) Toward northwest in the
Tunja-Paz de Ro area west of Soapaga Fault (Cspedes and Pea, 1995). Based on detrital composition
and palaeocurrents data the following detrital source areas have been proposed: (1) Santander-Floresta
Massiff (Fig. 3.4, Fabre, 1986; Jaramillo et al., 1993); (2) Quetame Massif (Fig. 3.4. Hoorn, 1988;
Jaramillo et al., 1993; Barrientos and Torres, 1994); (3) The Guyana Shield (Fig. 3.4, Rincn 1982 in
Barrientos and Torres, 1994; Jaramillo et al., 1993, Cooper et al., 1995); (4) The Central Cordillera:
proposed only for the sandstones of the Middle Magdalena area (Fig. 3.4, Lisama Fm, Hathon and
Espejo, 1997). Detrital composition and palaeocurrent data suggest that local erosion of the Upper
Cretaceous sedimentary rocks of the EC supplied quartz and fragments of chert and mudstones and
generated local exhumated areas that controlled deposition (Fabre, 1986; Jaramillo et al., 1993;
Barrientos and Torres, 1994; Hathon and Espejo, 1997). Evidence presented here indicates that some
areas of the EC started to exhumate and to supply sediment indicating that the early pre-Andean
foreland basin proposed by Cooper et al. (1995) was segmented by local exumated areas.

3.3. PALAEOGENE (EOCENE TO EARLY MIOCENE)

According to Cooper et al., (1995) the Eocene to Early Miocene sedimentary record was
deposited in a late pre-Andean Foreland basin. Because most Tertiary sediments of the MV and some
of the EC and LLA are dominantly continental and generally poorly fossiliferous, there are no reliable
ages for Palaeogene rocks in the whole MV (Jordan and Gomez, 1996) and correlation with the EC or
LLA is not well established. Thus the summary will proceed separately for different areas.

80
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera, Colombian Andes
3.3.1. LLanos and Eastern Cordillera

Late Eocene

These rocks are bounded at the bottom by the regional early to middle? Eocene unconformity
(Figs. 3.6 to 3.9). The unconformity in some places has been described as a paraconformity, but in
others places, at least locally, as an angular unconformity (e.g. locally in the LLA foothills (Aguaclara
area) the lower contact of the Mirador Fm is an angular unconformity over the Cuervos Fm., Barrientos
and Torres, 1994). Fabre (1986) has reported a local angular unconformity at the bottom of the Middle
(?) to late Eocene Picacho Fm in the Paz de Ro area.

Serrania Maracaibo

Buc
ucas de Perija Basin

aram
1400 Cucuta Barinas

ang

Chit
San L

(?) Basin

aF

aga
F
Romeral F

ia de

(?)
Serran

1300 Bucaramanga
ra
ille

Arauca
rd
Co
e

S
F

Medellin
z

1200
are
Su

(?)
Llanos Orientales
aF

Basin (LLA)
lin

Tunja
Sa
La

1100
em Yopal
st
F

1 sy
Cambras

Manizales fa
ult 0 100 200 km
o
(?) am
ar
u aic
Bogota G EARLY EOCENE
1000 Ibague UNCONFORMITY
Beheaded accretionary wedge
F l
era
Rom

Villavicencio
Highly variable angular
unconformity
(?)
l

900
a
nt r

Truncated homocline / broad low


Ce

relief folds
(?)
Neiva Uplifted blocks with no deformation
or very modest deformation
La Macarena
Serrania de

800
F

F
a

ira
m
us

m
ta
Ch

(?) Paraconformity
Al

800 900 1000 1100 1200 1300


Evidence eroded 1400 1500

Figure 3.6: Nature of the early Eocene unconformity. Without palinspastic restoration (modified from George et al.,
1997).

In the LLA area (Fig. 1.1) initial deposition consisted of marine-influenced, sandstone-rich,
fluvial and estuarine valley fill deposits, and continued in muddier coastal plain sediments (Figs. 2.7 to
2.11, 3.10, Pulhamn, 1994; Cooper et al., 1995; Fajardo, 1995). Coarse and often pebbly, fluvial

81
Chapter 3
and alluvial fan sandstones are the dominant sediments deposited over a wide area of the LLA area
(Mirador Fm). Continued transgression eventually submerged the alluvial plain and established a
shallow marine shelf across the Cusiana area (Cooper et al., 1995). The upper part comprises heavily
bioturbated estuarine parasequences punctuated by sandstone-rich, estuarine valley fill deposits
(Cooper et al., 1995). In the Cusiana oil field major flooding ended sand deposition (Fajardo et al.,
1993; Fajardo, 1995; Cooper et al., 1995). In the axial region of the EC a braided fluvial fining upward
succession with conglomerate at the bottom followed by coarse to medium grained sandstone
(Regadera Fm, Julivert, 1970; Picacho Fm, Giraldo et al., 1993) recorded a base-level rise. A maximum
flooding surface possibly is represented by marine iron-rich ooltic facies in the lower part of the
Concentracin Fm. (Cspedes and Pea, 1995; late Eocene according to Cazier et al., 1996). This
sequence is only present in the western side of the Soapaga Fault. Significant lateral changes of
thickness may be the result of fault control (Cooper et al., 1995).

4 Km
NW SE

700 600 500 400 300 200 100


0.0

0.5

1.0

1.5

2.0

2.5

3.0 K
3.5

4.0
TWT (Sec)

4.5

Figure 3.7: Seismic line BR8609-SC through the Cchira High in the Middle Magdalena Valley. Note the lower-
middle? Eocene angular unconformity truncating earlier structures (from Mantilla, 2000). Location of this line is
indicated in figure 3.2.

Detrital source areas. In the LLA area sandstones are extremely mature quartz arenites,
although locally fine-grained litharenites occur (Fajardo et al., 1993; Fajardo, 1995; Cooper et al.,
1995). In the EC (Tunja-Sogamoso area) conglomerates contain clasts of black chert, some with
Cretaceous foraminifera, sedimentary and low grade metamorphic rocks, evidencing erosion of some
areas of the EC (Figs. 3.10 and 3.11), as also suggested by palaeocurrent data (toward SW, Cspedes
and Pea, 1995). Northward in the Cocuy area the unit contains conglomerate and lithic to sublithic
sandstones with fragments of Cretaceous chert, mudstones and polycrystalline quartz (Fabre, 1986).
Fabre (1986) proposed the uplift of the Santander Massif (Figs. 3.10 and 3.16) as detrital source.

82
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera, Colombian Andes
According to this author, the relative decrease in chert content compared to late Paleocene units
suggests that the erosion reached deeper levels than the Late Cretaceous chert eroded during late
Paleocene. Based on detrital mineralogical composition of sandstones, in the Tunja-Sogamoso area
Meja and Giraldo (1993) concluded that the detrital source area was located in the Floresta Massif and
probably in the Santander Massif (Figs. 3.10 and 3.11). Based on sandstone composition and
palaeocurrent data (toward NNE) Barrientos and Torres (1994) proposed the Guyana Shield and the
Upper Cretaceous sandstones of the EC as detrital sources for the LLA foothills (Aguaclara area).
Evidence presented here indicates that some areas of the EC started to became exhumated and to
supply sediment. Once again local exhumated areas segmented the late pre-Andean foreland basin
proposed by Cooper et al. (1995).

Figure 3.8: Outcrop view of the lower-middle? Eocene angular unconformity truncating Cretaceous beds. On top the
Palaeogene Chicoral Fm (part of the Gualanday Gp.) Drawing from a photograph looking at the South, taken in the
Loma de Los Caballos, Upper Magdalena Valley (from Amzquita and Montes, 1994). Location of this outcrop is
indicated in figure 3.2.

Oligocene to Early Miocene?

In the LLA area and foothills four cycles of marine influenced, lower coastal plain deposition
accumulated (Figs, 2.7 to 2.11 and 3.12, Carbonera Fm). These sequences recorded eastern migration
of foreland basin subsidence, which culminated with the onset of EC deformation (Cooper et al., 1995).
Eastward these sequences are sandier and more continental approaching the Guyana Shield. Sequences
thicken westward (Fig. 3.13). Cooper et al. (1995) suggested that continued episodic normal
displacement on the Cusiana fault system took place during this time. Each sequence consists of a thin,
forced regression systems tract, a sand prone transgressive systems tract, a maximum flooding surface
and a mud-dominated highstand system tract (Cooper et al., 1995). In the axial region of the EC
correlative deposits are yellow to grey claystone with thin beds of medium grained sandstone
(Concentracin Fm, Reyes and Reyes, 1976, Oligocene-early Miocene according to Germerard, 1968;
Usme Fm, Julivert, 1970). Northward in the Cocuy area, Fabre (1986) described the lower part of the
Concentracin Fm, as containing dark grey to black locally carbonaceous mudstone with thin beds of

83
Chapter 3
sandstone, one of them with iron oolites, which he interpreted as having been deposited in a littoral to
shallow marine environment. Cooper et al. (1995) stated that sediments prograded westward and onlap
eastward. Julivert (1970) described local unconformities in anticlines but continuous sedimentation in
the Usme Syncline in the Regadera and Usme Fms of the Sabana de Bogot area and interpreted this as
the result of simultaneous folding and sedimentation.
WN W ESE
2 3 4
1 5 6
6

Figure 3.9: Section constructed from field data through the Carmen de Apical Syncline in the Upper Magdalena
Valley. 1. Cretaceous Villeta Fm. 2. Upper Cretaceous Guadalupe Gp. 3. Palaeogene middle part of Gualanday Gp.
4. Palaeogene Upper part of Gualanday Gp. 5. Palaeogene? Barzaloza Fm. 6. Neogene Honda Gp. Note
unconformities and lateral changes of thickness (from Guillande, 1988). Location of this section is indicated in
figure 3.2.

Detrital source areas. According to Cooper et al. (1995) the main source of sediments was the
Guyana Shield (Fig. 3.12). However, Cardona and Gutierrez (1995) studied the mineralogical detrital
composition of the sandstones (sublitharenites to muddy sublitharenites and quartzarenites at the
bottom) of the Carbonera Fm in the LLA foothills (NW of Yopal). They found metamorphic rock
fragments (phylite quartzite), plagioclase, microcline, chert, polycrystalline quartz, zircon and
tourmaline, feldspar and locally igneous rock fragments. Metamorphic fragments generally are phylite
indicating a relative near detrital source area. Well preserved, angular up to medium sand size grains
also indicate a near detrital source area. Palaeocurrent data dominantly are eastward (SSE and ENE).
They conclude that at least two detrital source areas existed which were located at the west, one of
them near the deposit. In the detrital source areas both metamorphic-igneous and sedimentary rocks
were eroded. The nearby area could be located at the Floresta, Santander or Quetame Massifs (Fig.
3.12) and probably includes some erosion from Late Cretaceous sedimentary rocks from the EC
(Guadalupe Gp), as suggested by the presence of chert and polycrystalline quartz fragments. The far
detrital source area could be the Central Cordillera (Fig. 3.12, Cardona and Gutierrez, 1995). Moreno
and Velzquez (1993) described an upward increase in sand grain size and interpreted it as reflecting a
slight exhumation of the detrital source area. The presence of Cretaceous chert clastics with
foraminifera (Fajardo-Pea, 1998) corroborates that some parts of the EC started to exhumate and
supply sediment, thus the late pre-Andean foreland basin proposed by Cooper et al., (1995) was
segmented by local exhumated areas. Seismic evidence (Fig. 3.14) suggests that contractional inversion
of Mesozoic faults occurred during Palaeogene.

3.3.2. Magdalena Valley (MV)

Tertiary sediments of the MV (Fig. 1.1) are dominantly continental and poorly fossiliferous
(Figs. 2.7 to 2.11 and Figs. 3.10 to 3.13). There are no reliable ages for Palaeogene rocks in the whole
MV.

84
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera, Colombian Andes

EOCENE Serrania

ucas
Maracaibo
PALEOGEOGRAPHY de Perija
Cucuta Basin

San
an L
1400 Positive relief.

Inv Sa High
tan

Floresta
Alluvial fan and fluvial Barinas

de S

erd
Coastal plain predominantly Basin

ntand
Mas
sandstones

nia
Coastal plain predominantly

sif

er
mudstones

Serra

y
1300 Littoral to inner shelf

lle
Bucaramanga Arauca

Va
sandstones
Shallow marine inner shelf

na

na
carbonates

ale

ale
Shallow marine inner shelf

gd

as agd
mudstones and siltstones

Ma

M
Outer shelf shales of

in

thills
le
carbonates

Su lazo
1200 Turbiditic sandstones dd

b-B
Mi

Foo
ab
Llanos Orientales

T
Basin (LLA)
a

tem

so
ed

nos
ult
iller

Re gamo
ert
fa
sys
ord

Lla
Inv
eo Tunja

n
la

em
l C

gio
So
lt

pa
fau

st
ntra

sy
z

nja
1100
a re

ult
eo

Yopal
Ce

Su

fa
T
pala

ta e
Manizales

fa laeo
eo

go d
Bo ana
Pala

eo pa
b- da

ult
ima

sin

b
Su on

Sa

pa mo
ba
Bitu

ra
la

ya aic

1000 Bogota
Bo Gu

Ibague
ca
ba t o
sin
Su rard

Villavicencio
y
lle
Va
Gi
b-

a
len
da

900
ag
r M
pe
sys eo

Up
te m
la

Guyana Shield
fault a pa

Neiva
sm

San Jose del


La Macarena
Serrania de
as in

-
0 100 200 Km
b- b a
Chu

800 Guaviare
i v
Ne
Su

800 900 1000 1100 1200 1300 1400 1500

Figure 3.10: Eocene palaeogeography without palinspastic restoration (modified from Geotec, 1992; Cooper et al.,
1995 and Villamil, 1999).

These dominant continental rocks are characterized by important lateral changes of facies and
thickness.
In the Middle MV (Fig. 2.1) from bottom to top the units are the Chorros Gp (La Paz,
Esmeraldas Fms) and the Chuspas Gp (Mugrosa and Colorado Fms). In the Upper MV (Fig. 2.1) the
major units from bottom to top are the Gualanday Gp and the Barzaloza Fm. In the Honda sub-basin
(Fig. 2.1), the Hoyn and San Juan de Rioseco Fms have a similar relative stratigraphic position as the
Gualanday Gp. However, in the Honda sub-basin, Porta (1965, 1966) showed that the conglomeratic
Hoyn Fm pinches out and disappears totally toward the south and east of the Guaduas syncline and
thus, it is not a northward lateral continuation of the Gualanday Gp of the Upper MV. In the Honda
sub-basin the Santa Teresa Fm, which lithologically overlies these units, is chronostratigraphically
similar to the uppermost part of the Chuspas Gp of the Middle Madgalena Valley (La Cira fossiliferous
shale horizon, Nuttal, 1990, Figs. 2.7 to 2.11 and Figs. 3.10 to 3.13).

85
Chapter 3

Maracaibo

ca s
Serrania

es
Basin
de Perija

d
n Lu

An
1400 EOCENE

a
Cucuta

id
e Sa

er
M
THICKNESS

nia d
(meters) Barinas
200 Basin

Serra
1300 935 Bucaramanga
67
0

lt
39 68

fau
Arauca
ra 200

eo
78
d ille
100 80

ala
220

zp
284?

are
Co r

1200

Su
t
ul
1150 45
fa 284
Llanos Orientales
o
lae 380
32
pa Basin (LLA) 0
ca 250
ya 53
Bo
0 186
80
Medellin
Tunja 50
27
489
1100 0 40?
160
Yopal 75 43
79
345
33

40? 86?99? 0
230
110 48 45
136 29 28
946 62
340 38 28
160 105 61 71
680 300 132 76 48 23
Manizales 100 51 53
76
38 43 77
27
1000 824 27
38
35?
660
Bogota
64 43 61 2
20
Villavicencio 173 0
73 0
7
490 92 24
294 36 0 0
45
49 0
900
1150 0?
l

28
n tr a

Neiva 0? 9
Ce

746
0 Guyana Shield

0
La Macarena

800 0 100 200 Km


Serrania de

San Jose del


Guaviare

800 900 1000 1100 1200 1300 1400 1500

Figure 3.11: Eocene restored thickness (meters) without palinspastic restoration. Thick lines represent palaeo-faults
believed to have been active during Eocene time.

Middle Magdalena Valley

The Tertiary record (Figs. 2.7 to 2.11) is composed of cross-bedded sandstone, locally
conglomeratic (fluvial channel), variegated to red mottled siltstone and mudstone (alluvial plain, lakes,
and palaeosoils), dominantly of fluvial to lacustrine origin (Figs. 3.10 to 3.13). Three major units with
different sand /mudstone ratios are recognized, each one limited at the bottom by an unconformity.
Each unit was deposited during a continuous increase of the accommodation space/sediment supply
ratio that ended with lacustrine sedimentation (Ecopetrol-ICP, 1996). Lacustrine facies are fossiliferous
green to black shales with fresh water to brackish mollusks and locally glauconite (e.g. top Mugrosa
Fm, Morales et al., 1956). Three of these fosiliferous shales lie at the uppermost part of each unit: (1)
The Chorro Group (sand-rich La Paz Fm, and Esmeraldas Fms); (2) The Mugrosa Fm (mud-rich); and

86
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera, Colombian Andes
OLIGOCENE Serrania Maracaibo
EARLY de Perija Basin

San

s
MIOCENE

de
1400 Cucuta

tand

An
PALEOGEOGRAPHY

Lucas

a
er

id
Positive relief.

er
de San

Mas

M
Alluvial fan and fluvial

sif
Coastal plain predominantly
sandstones Bucaramanga

Serrania
1300

Inv Santand
Floresta
Coastal plain predominantly

y
lle
mudstones

Va
Littoral to inner shelf Arauca

na
sandstones

ale

High
Shallow marine inner shelf

gd
carbonates

er
Ma
Shallow marine inner shelf
1200 mudstones and siltstones
le
dd

as y
Outer shelf shales of

b-B cu
Mi

in
carbonates

Su Co
Llanos Orientales

b-B len zo
Turbiditic sandstones

Su gda bla

Inv
Basin (LLA)

as a
a
Ma d T

in
Tunja
tral

e
ert

1100
Cen

Inv
lera

Yopal
b- da
sin
eo

rdil

Su on

ills
de
ba

Manizales
Pala

Co

th
H

na

o
ba ota

Fo
Sa og
if
ss

B
s
Ma

Bogota no
Lla
1000
e

Ibague
m
eta
ba ot
sin
Su irard

Qu
y
alle

Villavicencio
G
b-

a V
len

900
gda
Ma
per

Guyana Shield
Up

Neiva
La Macarena

San Jose del


Serrania de
as in

-
800 0 100 200 Km
Su e i v a

Guaviare
b- b
N

800 900 1000 1100 1200 1300 1400 1500

Figure 3.12: Oligocene-Early Miocene palaeogeography without palinspastic restoration (modified from Geotec,
1992; Cooper et al., 1995 and Villamil, 1999).

(3) the Colorado Fm (mud-rich, deposited in a mandering fluvial to lacustrine environment, Ecopetrol-
ICP, 1996). Sand-rich units were deposited predominantly by braided fluvial systems and mud-rich
intervals predominantly in meandering fluvial systems (Jordan and Gomez, 1996). Mud-rich facies
were developed especially during maximal base-level. Important lateral changes of thickness or texture
occur (Morales et al., 1956). Maximum thickness occurs near the EC foothills (Morales et al., 1956)
and decrease considerably or reduce to zero on palaeo-highs that were episodically active during
sedimentation (e.g. Chucur-Casabe, Cchira, Figs. 3.7, 3.11 and 3.13) separating sub-basins (e.g.
Cantagallo-Yarigu; Cross et al., 1996a,b). Several smaller sequences have been recognized by
Ecopetrol et al. (1994), Cross et al. (1996a,b) and Ecopetrol-ICP (1997). Ecopetrol et al. (1994)
recognized 6 sequences (one approximately equivalent to La Paz and Esmeraldas Fms, two
approximately equivalent to Mugrosa Fm, and three approximately equivalent to Colorado Fm) which
are richer in sand westward. According to Ecopetrol et al. (1994), 5 sequences display similar
truncation and onlap patterns (Fig. 3.15). Each sequence thins westward.

87
Chapter 3

Serrania Maracaibo
OLIGOCENE de Perija Basin

s
de
1400 EARLY

An
Cucuta
Barinas

Lucas

a
MIOCENE

id
Basin

er
M
THICKNESS

de San
(meters) -
0 100 200 Km
1300 100 1117
950 2130 Bucaramanga

Serrania
1034
>689 952 464
Arauca
100 270
>64
1350 358 285
140
315

Medellin
1200 100 1475 410
ra

315
d i ll e

405
1200
1500 460 299
637 Llanos Orientales
Co r

0 100
150
100 Tunja 478
Basin (LLA)
380
441
600
1100 0 50
1050
768
719
309

Yopal 451
200
Manizales 0 80
1478 995 407
100 650 >301 991 665 358 304
829 1051 671
780 533 387 336
1010 1079 582 381
1040 650 534 415 381
2615 430
710 671
Bogota 698 427
721 250 325
1079 701 334
1000 Ibague 250 520
305

887
512 256 247
597
1621 55
1026 232
912 309
425
49
382
1000 561
Villavicencio 726 271 238
564
426
900 850 748 377
374
797 378

268
al

807
ntr

La Macarena

Neiva 227 Guyana Shield


Serrania de
Ce

800 San Jose del


Guaviare

800 900 1000 1100 1200 1300 1400 1500

Figure 3.13: Oligocene-Early Miocene restored thickness (meters) without palinspastic restoration. Large bold
numbers indicate palaeotopography in meters estimated from 2D flexural modelling. Thick lines represent palaeo-
faults believed to be active during Oligocene-Early Miocene time.

The basal beds of each sequence onlap the underlying lower sequence boundary from east to west
while the tops of each sequence are truncated by the overlying upper sequence boundary. These
relations are the effect of regional tilting and accommodation with sediment having greater preservation
potential away from the Central Cordillera. All sequences onlap the western flank of the basin in the
southern Middle MV, they are fault-bounded against the Cantagallo Fault, and are exhumated and
truncated in the northern part of the basin (Ecopetrol et al., 1994). Cross et al., (1996a,b) and
Ecopetrol-ICP (1996) also recognized 6 stratigraphic intervals delimited by time correlation surfaces
defined through recognition of base-level or accommodation space/sediment supply ratio cycles.
Similar to the Ecopetrol et al.s sequences, one interval is approximately equivalent to La Paz and
Esmeraldas Fms, two approximately equivalent to the Mugrosa Fm, and three approximately equivalent
to the Colorado Fm.
The regional Eocene unconformity is the base of Palaeogene sedimentary record (Figs. 2.7 to
2.11 and 3.6). In the Middle MV it is a dramatic angular unconformity (Fig. 3.7). Paleocene-Cretaceous
and older Fms below it are strongly folded and faulted. This unconformity post-dates

88
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera, Colombian Andes
most but not all exhumation and folding (Cross et al., 1996a,b). According to Cross et al. (1996a,b)
several unconformities and multiple times of tectonism are developed in the Cretaceous to Lower
Tertiary interval, resulting in a single unconformity (Fig. 3.15). During deposition of Chorro Gp (La
Paz and Esmeraldas Fms) there was fault activity (Ecopetrol et al., 1994). Topographic relief on the
unconformity is up to 920 m near the persistent Cchira (N-S oriented, Fig. 3.7) and Chucur-Casabe
(SW-NE oriented) palaeo-highs. The palaeo-highs separate two sub-basins to the northwest
(Cantagallo-Yar area). During deposition of the Chorro Gp. the irregular basin began to fill. The
southwest end of the Chucur-Casabe palaeo-high was an erosional highland while the Cchira
palaeohigh was overlapped (Cross et al., 1996a,b). The first phase of basin filling is rich in sandstone.
Multiple episodes of sub-basin filling are revealed in the basal onlap pattern. The top of Chorro Gp
(approximately Orange horizon of Cross et al., 1996a,b) is an onlap surface for the overlying unit. The
upper boundary of the interval is characterized by minor erosion (Cross et al., 1996a,b).

NW Seismic Profile ME94-1400 SE

.
Fm
. 0
Fm

ra
on

cla
Le

ua
. Leo
Fm

Ag
a n Fm.
ner
arbo
C Car
bon 1.0
e ra
Fm
.

2.0

TWT (sec)
Fm.
Mirador Mirado
K r Fm.

3.0
K

4.0

0 5 km

Figure 3.14: Seismic profile in the Medina area, eastern flank of the EC. Note contractional inversion of Cretaceous
extensional faults occurred during Palaeogene time (from Linares, 1996). Location of this profile is indicated in
figure 3.2.

89
Chapter 3
During deposition of the Mugrosa Fm, continued onlap occurred onto the Chucur-Casabe
palaeo-high, which remained as a palaeo-high during deposition of the lower interval
(Orange-purple of Cross et al., 1996a,b). However, it almost disappeared at the end (a more
evenly thick isopach of the upper interval) during a gradual elimination of topographic relief.
Each interval represents a basal fill episode, limited by an erosion surface and bi-directional
onlap on that surface. Minor episodes of tectonic movement were recognized on some basal
onlap patterns. Accommodation space/sediment supply ratio increased within each interval
and between different intervals (Cross et al., 1996a,b).

Oligocene. C. Early Oligocene to Eocene. D. Paleocene-latest Cretaceous. E. Late Cretaceous. BT. Backthrust. Note lateral
Figure 3.15: Seismic profiles through the Paya Anticline in the Middle Magdalene Valley. A. Quaternary to Miocene. B. Late

changes of thickness of Palaeogene sequences and angular unconformities between them evidencing synchronous deformation
1 Km

E
D

F
C
B
A

LT
FAU
OA
PAY
Payoa -21

and sedimentation (from Kovas et al., 1982). Location of these profiles is indicated in figure 3.2.
Payoa -4

ULT
D

A FA
Aguas Claras -4

ALIN
BT

LA S
E
D
B

D
C
Aguas Claras -1

C
NW

C
1.0

2.0

3.0
0

TWT (sec)

90
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera, Colombian Andes

1 Km

D
A

E
Payoa -26

A T
AYO UL
P FA

Figure 3.15: Continued


T
FAUL
C
B

ALINA
LA S
BT
D
C
NW

0.0

1.0

2.0

3.0

TWT (sec)

91
Chapter 3
During deposition of the Colorado Fm local tectonic reactivation of the Cchira palaeo-high occurred
and remained prominent. The lower boundary is a very well marked erosion surface that cuts into the
underlying interval and is characterized by a pronounced facies offset. Cross et al., (1996a,b)
recognized three intervals, each one characterized by highly aggradational vertically stacked mud-rich
facies associations culminating in shales intervals deposited during periods of maximum base-level.
Accommodation space/sediment supply ratio increased within each interval and between different
intervals (Cross et al., 1996a,b).
Detrital source areas. Sandstone of the La Paz, Esmeraldas, Mugrosa and Colorado Fms are
characterized by quartz, sedimentary chert and K-feldspar as principal detrital components, with more
K-feldspar than plagioclase and accessory aplitic lithic fragments, micas and zircon. An upward trend
of increasing mineralogical maturity is observed (Hathon and Espejo, 1997). These sandstones fit into
the craton/continental block provenance field on quartz, feldspar and lithic fragments plots (Dickinson,
1985). Hathon and Espejo (1997) interpreted an exhumated plutonic/high-grade metamorphic basement
as source for the sediments, probably located in the area of the Central Cordillera (Figs. 3.10 and 3.12).
However, southeasterly and easterly local sources from the area of the EC are suggested by
palaeocurrent measurements and grain size trends in the Opn, Malpaso and Ro Sogamoso areas
(Hathon and Espejo, 1997). According to Geotec (1994) dramatic lateral changes of thickness and
palaeo-current with a westward dominant direction during deposition of the Esmeraldas Fm suggest a
local detrital source area located probably in the Santander Massif (Figs. 3.10 and 3.12). Evidence
presented here indicates that some parts of the EC started to be exhumated and to supply sediment, thus
demonstrating that the late pre-Andean foreland basin proposed by Cooper et al. (1995) was segmented
by local exhumated areas.

Upper Magdalena Valley and Honda sub-basin

The Gualanday Gp of the Upper MV and similar coarse detrital units of the Honda sub-basin,
i.e. the Hoyn and San Juan de Rioseco Fms. All these rock units are characterized by the presence of
conglomerate and conglomeratic sandstone.
The base of the Gualanday Gp varies spatially from conformable to angular unconformable
(Figs. 3.8 and 3.9). This geometric variability, along with strong thickness and facies variations (Figs.
3.10 to 3.13), suggests both palaeogeographic and tectonic controls on its accumulation. In the
Gualanday Gp in the Neiva sub-basin (southern part of the Upper MV, Fig 2.1), Anderson (1970, 1972)
described three conglomerate intervals separated by two mottled red mudstones and lithic sandstone
intervals. Conglomerates are predominantly close-packed made mainly of chert pebbles.
Conglomerates and sandstones show an upward increase in quartz through the section. The lower
contact of the Gualanday Gp in the Neiva sub-basin has been reported to be paraconformable, at least
locally. However, in the Girardot sub-basin (northern part of the Upper MV, Fig 2.1) various levels of
the Gualanday Gp overlap tilted and deformed Cretaceous rocks or older rocks (Figs. 3.8 and 3.9,
Raasvelt, 1956; Raasvelt et al., 1957; Guillande, 1988; Amaya and Santamara, 1994; Etayo-Serna,
1994; Ojeda and Pea, 1994). Guillande (1988) also reported local unconformities at the top of
Potrerillo and Doima Fms of the Gulandaly Gp. The conglomeratic levels were deposited in alluvial
fans and braided streams, while the red mudstone or siltstone levels suggest deposition on flood plains
between channels. Coarse detrital fluvial/alluvial fan facies are overlain by lacustrine (?) mudrich
facies (Santa Teresa Fm. in the Honda sub-basin, and Barzaloza Fm in the Upper MV) suggesting a
relative base-level rise.

92
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera, Colombian Andes
Detrital source areas. Palaeo-currents and clast composition indicate that the gravels were
derived form Cretaceous strata, mainly due to episodic exhumation of the Central Cordillera (Figs. 3.10
and 3.12, Beltrn and Gallo, 1968; Van Houten and Travis, 1968; Anderson, 1970, 1972). Chert clasts
were derived from erosion of Upper Cretaceous rocks while detrital kaolinite-rich mudstones probably
were derived from Cretaceous shales (Anderson, 1970, 1972). With increasing depth of erosion through
time the quartzose lower part of the Cretaceous supplied an increasing amount of quartz to the basin.
Van Houten and Travis (1968) postulate two pulses of exhumation of the Central Cordillera, the first
represented by the lower Gualanday Gp (Chicoral Fm), while the second episode corresponds to the
uppermost part of the Gp (Doima Fm). Anderson (1970, 1972) postulated three pulses of exhumation,
one for each of the three conglomeratic levels described by him in the Gualanday Gp. Active thrusting
during Eocene-Oligocene times has been proposed in the Upper MV (Butler and Shammel, 1988;
Guillande, 1988; Caicedo and Roncancio, 1994; Amezquita and Montes, 1994; Amaya and Santamara,
1994). Locally in the western part of the Girardot sub-basin the Gulanday Gp has been related to
exhumation along the Chaparrral thrust fault (Caicedo and Roncancio, 1994), close to the western
border with the Central Cordillera. They demonstrated that deposition of the Gualanday Gp occurred
simultaneously with east-vergent thrusting and that sediments were derived from erosion of
progressively deeper stratigraphic levels, including a Cretaceous sedimentary cover, Jurassic volcanic
rocks and Palaeozoic igneous and metamorphic basement.

4. SUBSIDENCE ANALYSIS OF THE PALAEOGENE SEDIMENTARY RECORD

The study of the sedimentary fill of a basin and its associated tectonic subsidence signal gives
important information about basin formation mechanisms (Cloetingh et al., 1993). In order to quantify
the tectonic component of subsidence of the Palaeogene basin of the study area, a 1D-backstripping
technique was used (Steckler and Watts, 1978; Bond and Kominz, 1984). Details of the backstripping
subsidence analysis, the procedure and data are described in Chapter 2.
Figure 3.2 shows the location of the columns and wells used. The Palaeogene sedimentary
record of the study area only has been preserved in the LLA and MV areas, and partially along the axial
zone of the EC (Fig. 2.1, Sabana de Bogot, Tunja and Sogamoso regions). In these regions only a very
limited number of stratigraphic sections can be used for the subsidence analysis. The original data from
the stratigraphic columns have been interpolated or extrapolated assuming the interpretations drawn on
the palaeogeographic and thickness maps were valid (Figs. 3.3 to 3.5 and Figs. 3.10 to 3.13). Where
possible I used also the restored original thicknesses estimated from thermal organic matter maturity
modelling calibrated by vitrinite reflectance data from the available literature (Keal, 1985; Hbrard et
al., 1987a,b; Bachu et al., 1995; Cazier et al., 1995; Mayorga and Vargas, 1995; Ecopetrol and Beicip,
1995; Garca-Gonzalez et al., 1997; Mora, 1997, Mora et al., 1997).

4.1. RESULTS

Figure 3.16 shows some examples of the tectonic subsidence curves during early Tertiary times.

Magdalena Valley. Where the sedimentary record has not been partially eroded, or it is
incomplete, there are important differences in the shape of the subsidence curves for different wells.
(1) Middle Magdalena Valley Detailed thickness maps of several stratigraphical intervals in the
Middle MV (Ecopetrol-ICP, 1996) show dramatic lateral changes in thickness, which together with
evidence for active tectonics (Fig. 3.15) suggest piggy-back sedimentation. In the Guaduas Syncline

93
Chapter 3
(Honda sub-basin, Fig. 2.1) Gomez (1999) have shown seismic evidence of Palaeogene sedimentation
simultaneous with folding. Maximum tectonic subsidence occurred in the eastern border of the MV
(e.g. Tablazo section, Palaeogene tectonic subsidence 500 m) close to the faults limiting the western
border of the EC. The convex upward shape of the tectonic subsidence curves indicating increasing
subsidence rate through time during Palaeogene (Fig. 3.16) is similar to the foreland basin subsidence
pattern (Allen and Allen, 1990).
(2) Upper Magdalena Valley. Amount of Palaeogene tectonic subsidence in this area is variable,
but less than 500 m. Some authors (Amezquita and Montes, 1994; Amaya and Santamara, 1994;
Caicedo and Roncancio, 1994) have demonstrated Palaeogene sedimentation coeval with both east- and
west-vergent thrusting (Etayo-Serna and Florez, 1994). Detailed thickness maps and local evidence
(Figs. 3.8 and 3.9) also suggest piggy-back sedimentation.
Llanos Orientales. The convex upward shape of tectonic subsidence curves (Fig. 3.16) clearly
shows acceleration of tectonic subsidence in a typical foreland pattern (Allen and Allen, 1990).
Maximum Palaeogene tectonic subsidence is about 250-m in the Arauca1 well (Fig. 3.16b).
Sabana de Bogot. A limited number of stratigraphic columns represent an incomplete
sedimentary record. The most complete stratigraphic section in the Usme Syncline (La Calera and
Cqueza curve with about 750 m of tectonic subsidence, suggests a relative rapid Paleocene subsidence
during deposition of the Guaduas and Bogot Fms. Julivert (1970) has presented evidence of
Palaeogene folding simultaneous with sedimentation in the Usme Syncline.
Tunja-Sogamoso region. Limited data suggest two periods of relative rapid subsidence during
the Paleocene (Guaduas and Socha Fms) and the Oligocene (Concentracin Fm). Maximum tectonic
subsidence is about 400 m.

4.1.1. Events of tectonic subsidence

Although there are important differences in the subsidence pattern of different stratigraphic
columns the following subsidence events were identified:
(1) Late Maastrichtian-Paleocene accelerated subsidence, especially in the axial zone of the palaeo-EC
(Sabana de Bogot region).
(2) The Eocene unconformity is represented as a period of no subsidence.
(3) Oligocene increased subsidence is regionally recognized.
The following is a preliminary attempt to correlate events observed in the subsidence curves with plate
tectonic events as proposed in the literature:

During Maastrichtian and Paleocene several authors (McCourt et al., 1984; Pindell and Tabbutt, 1995)
have interpreted accretion of the oceanic terranes of the Western Cordillera, with maximum tectonic
activity during the Paleocene. This accretion produced exhumation in the palaeo-Central Cordillera,
development of east-vergent thrusts in the Middle MV, and initial pre-Andean deformation and
differential rock-uplift/subsidence of blocks in the area of the EC. As mentioned in Chapter 2 increased
subsidence in the axis of the Cundinamarca sub-basin (Sabana de Bogot) could be the result of
increased horizontal compressional stress (Cloetingh, 1988; Cloetingh and Kooi, 1992) associated with
collision of the oceanic terranes of western Colombia and deformation and uplift of the Central
Cordillera. Development of normal faults in the Llanos area (e.g. Kluth et al. 1997) could be the result
of local tensional stresses in the flexural bulge developed.

94
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera, Colombian Andes

MIDDLE MAGDALENA VALLEY


200 150 100 50 0
0

E Miocene
Oligocene
Age (Ma) W Lebrija

Paleo
200

Ma as
Eocene

Subsidence(m)
EL
400 Llanito-1
600
800 Sabalo-1k
1000
Casabe-199
1200

Tectonic
1400 Infantas-1613
1600
Cascajales-1
1800
2000

EASTERN CORDILLERA
250 200 150 100 50

Tectonic Subsidence Uplif of rock (m)


-4000

E Miocene
Oligocene
Paleo
Maas
Eocene
-3000 Arcabuco
E L

-2000 Los Medios

Age (Ma)
-1000 Pajarito

0 Tibasosa

1000
Velez

2000 Lebrija

3000

LLANOS ORIENTALES
100 80 60 40 20 0
0
Subsidence(m)

200
Age (Ma)
400 La Maria-1

600
Arauca-1
800

1000 Cao Limon-1


Tectonic

1200 La Tortuga-1
Oligocene

Miocene
Maas

Paleocene
Early

Eocene 1400
E L
1600

Figure 3.16: Curves of tectonic subsidence of some wells (see Fig. 3.2 for location). Horizontal axis represents age
in Ma. Vertical axis represents tectonic subsidence in meters obtained from backstripping analysis.

95
Chapter 3
The Eocene unconformity correlates with an increase in convergence rate between Caribbean
and South American plates. During this time a maximum peak of deformation occurred (first event of
pre-Andean deformation).
During Oligocene time a second event of pre-Andean deformation occurred in the EC. It
correlates with thin-skin obduction of Panam (Choc block) onto Western Cordillera-San Jacinto
deformed belt (Pindell and Tabbutt, 1995), and with the break-up of Farallon plate into the Cocos and
Nazca Plates (Wortel and Cloetingh, 1981; Duncan and Hardgraves, 1984).

4.1.2. Maps of tectonic subsidence

Figures 3.17 to 3.20 show maps of tectonic subsidence for selected time intervals during the
Palaeogene. Maximum tectonic subsidence during the Paleocene occurred in the Sabana de Bogot area
(La Calera, Cqueza) as indicated in the thickness map. The calculated tectonic subsidence for the
Eocene in the EC is very small and the only significant tectonic subsidence occurred in the MV, as
suggested by the Eocene thickness map. The Oligocene to Early Miocene subsidence was maximum in
the eastern and western foothills of the EC.

Serrania Maracaibo
de Perija Basin

s
1400

de
Cucuta

An
Lucas

a
Barinas

id
er
Basin
de San

34.00 44.00 M
1300
Bucaramanga 0 100
8.00
200 Km
Serrania

0.00 124.00
Servita palaeo fault

0.00
0.00 0.00
1.00
20.00
0.00 Arauca
52.00
36.00 157.00
39.00 0.00
82.00
45.00
ra

Medellin 24.00
d i ll e

1200 29.00
o ca

50
t

0.00
ul

62.00
lae ya
fa

57.00
pa Bo
Cor

128.00 0.00
tem

em

90.00
110.00 0.00
t

95.00 Llanos Orientales


ys
s ys

82.00
0.00 79.00
ts

103.00
78.00 0.00
Basin (LLA)
ul

123.00 0.00
u lt

fa

144.00 82.00
0.00
o fa

147.00 Tunja 129.00


e

8.00
la

1100 98.00 169.00 111.00


pa
alae

0.00 119.00 0.00


11.00 0.00
o

0.00 Yopal
m

105.00
ap

ra

Manizales
Ic

233.00
im

69.00 0.00
86.00
ua

129.00 0.00
Bit u

0.00
G

84.00 7.00 0.00


33.00 0.00
109.00 108.00
0.00 0.00
339.00 0.00
1000 Ibague 36.00 106.00
Bogota 0.00
18.00 14.00 106.00 100.00
0.00 0.00
19.00 21.00 0.00
9.00 198.00 0.00
0.00

0.00 Villavicencio
26.00 10.00 0.00

49.00 0.00 0.00


900 BACKSTRIPPED
2.00

0.00 0.00 LATE


0.00
al

0.00
MAASTRICHTIAN
ntr

0.00 EARLY
Ce

81.00
0.00 PALEOCENE
Neiva
La Macarena

TECTONIC
Serrania de

800
0.00
San Jose del SUBSIDENCE
0.00 Guaviare Guyana Shield

800 900 1000 1100 1200 1300 1400

Figure 3.17: Observed late Maastrichtian-early Paleocene tectonic subsidence, in meters, calculated from
backstripping analysis.

96
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera, Colombian Andes

Serrania Maracaibo
de Perija Basin

s
1400 Cucuta

de
Lucas

An
Barinas

a
Basin

de San

id
er
M
105.00 -
-1.00 22.00
0 100 200 Km

Serrania
1300 117.00
0.00 Bucaramanga
41.00 88.00 0.00
0.00 Arauca
114.00

lt
0.00 0.00

f au
90.00 0.00

m
130.00

aeo

ste
102.00

lt
lae ca
au
19.00

sy
pal
Medellin

pa oy a
32.00
1200

of

lt
ra

au
ez
109.00

B
97.00 110.00

f
ar
d i ll e

eo
16.00

Su
155.00 7.00 0.00

a la
tem

39.00

op
39.00 0.00
Llanos Orientales
Cor

s ys

0.00 11.00
-1.00 123.00

m
21.00
Basin (LLA)

ra
49.00 Tunja 42.00
ult

49.00

i c
0.00
1.00 66.00 85.00
o fa

4.00 53.00

a
23.00
1100

Gu
0.00 19.00
68.00 Yopal
alae

17.00 52.00
57.00 14.00
25.00
ap

101.00
Manizales 40.00 38.00
im

73.00 0.00 0.00


99.00
107.00
Bit u

6.00 32.00 0.00


102.00 66.00 0.00
31.00 5.00 0.00
20.00 140.00 385.00
62.00 10.00 0.00
27.00 19.00 Bogota 142.00 0.00
1000 Ibague
26.00 18.00 0.00

17.00 20.00 243.00 0.00 0.00


0.00
29.00 23.00 30.00 0.00
0.00 0.00
11.00 0.00
0.00
8.00 Villavicencio
0.00
0.00 10.00
14.00
73.00
900
al

tem

0.00
n tr

BACKSTRIPPED
s
sy
Ce

0.00
LATE
lt
au

0.00
0.00
of

PALEOCENE
e

70.00 0.00
al a

0.00 TECTONIC
ap

Neiva 0.00 San Jose del


La Macarena

800 SUBSIDENCE
sm

Serrania de

Guaviare
u
Ch

0.00

800 900 1000 1100 1200 1300 1400 1500

Figure 3.18: Observed late Paleocene tectonic subsidence, in meters, calculated from backstripping analysis.

5. MODELLING OF REMAINING THERMAL SUBSIDENCE AFTER MESOZOIC RIFTING

Because lithosphere stretching generated an extensional basin during the Mesozoic, the first
hypothesis tested to explain the Palaeogene observed tectonic subsidence, was whether some remaining
thermal subsidence affected Palaeogene evolution (c.f. Desgaulx et al., 1991). Assuming that the basin
was not deformed and that the only basin subsidence mechanism during the Palaeogene was remaining
thermal subsidence following Mesozoic rifting, I calculated for several time intervals the remaining
thermal subsidence due to thermal re-equilibration of the lithosphere during Palaeogene. A description
of the forward modelling technique and procedures developed by Van Wees et al. (1998) is given in
Chapter 2. I used the lithosphere stretching factors calculated for Mesozoic extensional phases and all
the modelling parameters used in the Mesozoic rifting episodes (see tables 2.4a,b in Chapter 2). Finally
I compared the Palaeogene thermal subsidence with the observed tectonic subsidence, previously
calculated using backstriping techniques, in order to test the hypothesis of tectonic quiescence within
the basin during the early Tertiary.

97
Chapter 3

Serrania Maracaibo
de Perija Basin

s
1400 Cucuta

de
An
Barinas

Lucas

a
Basin

id
er
de San

M
8.00 -
1300 66.00
0.00
0 100 200 Km

Servita palaeo fault


0.00 0.00 65.00 Bucaramanga

Serrania
4.00
0.00 11.00
0.00 22.00 Arauca

lt
0.00

f au
10.00
98.00 2.00 28.00
-1.00

laeo
18.00
3.00

z pa
1200 Medellin 10.00

re
17.00
ra

o ca
t
Sua
106.00

ul
15.00

lae ya
fa
d i ll e

11.00

pa Bo
9.00
21.00
9.00
3.00 24.00 18.00 Llanos Orientales
Cor

0.00 2.00 -1.00


syste o

3.00 24.00
Basin (LLA)
fau lt a palae

Tunja
m

-2.00
22.00 17.00
4.00 5.00 14.00 11.00

1100 0.00 4.00 -1.00


5.00
24.00 15.00
im

17.00
32.00 Yopal 11.00
Bit u

21.00 12.00
Manizales 22.00
10.00 28.00 0.00
27.00 17.00
-2.00 33.00 19.00
0.00 15.00 8.00
15.00 25.00 19.00
14.00 16.00 48.00 20.00
36.00 0.00 16.00 21.00 10.00
52.00
21.00
29.00
14.00 16.00 20.00 23.00
1000 132.00 72.00 Bogota 12.00
Ibague 16.00
14.00
76.00 -2.00
30.00 24.00
-1.00 0.00 34.00 35.00 17.00 28.00 1.00
5.00
11.00
47.00
Villavicencio 39.00 0.00
3.00
55.00 29.00 10.00
41.00 0.00 0.00
25.00 21.00
79.00 12.00
20.00
900 3.00
al

191.00 0.00
nt r

4.00 1.00
0.00 BACKSTRIPPED
Ce

EOCENE
106.00 0.00
Neiva
TECTONIC
Guayana
0.00
SUBSIDENCE
Sheld
La Macarena

800
Serrania de

San Jose del


Guaviare

800 900 1000 1100 1200 1300 1400 1500

Figure 3.19: Observed Eocene tectonic subsidence, in meters, calculated from backstripping analysis.

5.1. RESULTS

Figures 3.21 to 3.24 show the theoretical remaining thermal subsidence in a scenario of tectonic
quiescence. A comparison of these theoretical remaining thermal subsidence maps (Figs. 3.21 to 3.24)
with the observed tectonic subsidence (Figs. 3.17 to 3.20), clearly shows profound differences implying
that the hypothesis of tectonic quiescence is not applicable to the Palaeogene history of the EC.
In general in the place of the former extensional basin (palaeo-EC) the observed tectonic
subsidence is smaller that the predicted thermal subsidence. Thus if a component of remaining thermal
subsidence occurred after the Mesozoic rifting, such subsidence was partially compensated by a
moderate relative rock-uplift produced by a different tectonic mechanism. Conversely in the MV and
LLA foothills areas, just out the margins of the former extensional basin, the observed tectonic
subsidence is larger than the theoretical remaining thermal subsidence, implying an additional
subsidence produced by some tectonic mechanism (Figs 3.17 to 3.24). These trends can be explained at
least qualitatively by assuming that an early episode of inversion of the former extensional basin
occurred during Palaeogene time.

98
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera, Colombian Andes

Serrania Maracaibo
de Perija Basin

Lucas

s
Cucuta

de
1400

An
Barinas

de San

a
Basin

id
er
M
Serrania
40. -
0.
Bucaramanga 0 100 200 Km

ra
169.
1300

Co rdille
158.
186. 187.
103.
303. Arauca
0. 42.
199. 201. 145. 119.
23.
49.

Medellin 0.
164.
1200 50.
0. 212.
172. 119.
0. 231.
104. 179.
0. 0. 22. Llanos Orientales
26.
0.
6. 107.
177.
Basin (LLA)
187.
43. 0. 177.
Tunja 87. 285.
180.
0. 127.
1100 25. 175. 7. 166. Yopal 182.
186. 265.
181. 264. 146. 178.
Manizales 25. 6. 154. 156.
97.
288. 210. 135.
37. 132.
148. 359. 163.
259. 156.
116. 98. 240.
33. 245. 170.
110. Bogota 148.
Ibague 374. 160. 145.
1000 34. 68.
260. 207.
139.
356. 261.
131. 165.
114. 218. 129.
18. 442. 256.
93. 123. 36.
374. 119.
45 Villavicencio 261.
7. 162.
269. 297. 124. 112.
113. 217.
184.
32.
900 139. 310. 179.
178.
338.
177. BACKSTRIPPED
OLIGOCENE
114.
137. EARLY MIOCENE
al

TECTONIC
ntr

Neiva 111.
La Macarena
Serrania de

800 SUBSIDENCE
Ce

San Jose del Guayana


11. Guaviare
Sheld

800 900 1000 1100 1200 1300 1400 1500

Figure 3.20: Observed Oligocene-Early Miocene tectonic subsidence, in meters, calculated from backstripping
analysis. Negative values mean tectonic uplift.

6. FLEXURAL BEHAVIOUR OF THE LITHOSPHERE

During processes of mountain building, the lithosphere bends due to the topographic loading
(Price, 1973). This bending demonstrates that the lithosphere is capable of supporting localized loads
such as mountain ranges by deforming over a much broader area. As a result, depressions are formed at
both sides of a mountain chain in which sediments can accumulate. The origin of these foreland basins
is, therefore, controlled mainly by the flexural behavior of the lithosphere (Beaumont, 1981). Flexural
studies that treat the lithosphere as an elastic sheet floating on a fluid asthenosphere simulates the
observations (Beaumont, 1981). Flexural studies of the lithosphere are constrained by the geometry of
the basement subsiding asymmetrically under the load of the mountain belt (e.g. Zoetemeijer, 1993).
Not only surface mass loads, but also intraplate forces can play a significant role in the development of
foreland basins (Cloetingh, 1988). In some flexural studies the topographic load has proved not
sufficient to explain the observed deflexion of the foreland (e.g. Royden and Karner, 1984).

99
Chapter 3
Serrania Maracaibo
de Perija Basin

s
Lucas

de
1400

An
Cucuta

a
de San
Barinas

id
er
50 Basin

M
60

Serrania
7 -
1300 16
29
2 0 100 200 Km
6 24 12 Bucaramanga
18
- 0 0
16
22 20
18 39
Arauca
27 15 51 0 0
41
53
32
1200 Medellin 24
14 0
ra

20 48
d i ll e

26 0
27 40
11
15 35 0
40 15 0
Cor

- 16
35 42 32 51
26
19 Tunja 0
Llanos Orientales
37 47 40 0
38
1100
74
30 32 Basin (LLA)
- 50 34 0 0 0
8 30
31
33 0
Manizales
43 Yopal 0
30 8 0 0
22
58 0 0 00
50 23 0
32 17 0 0 0
27 0 0
43 39 41 53 0 0
0
40 0 0 0 0
Bogota 00 0
1000 Ibague
23 35 0
0
0
14 41 40 44 30
27
20 19 0 0
31 0 0
30 0
20 Villavicencio 0
0 0 0
0 0
30 0 0
25 0 0 0
17 31 0
0
900
al

32
30 0
0
nt r

PREDICTED LATE
Ce

MAASTRICHTIAN
22
EARLY PALEOCENE
Neiva
0 THERMAL
800 20
La Macarena
Serrania de

San Jose del SUBSIDENCE (m)


1
7 Guaviare

800 900 1000 1100 1200 1300 1400 1500

Figure 3.21: Predicted late Maastrichtian-early Paleocene thermal subsidence after Mesozoic rifting, in meters,
calculated from forward modelling using lithosphere-stretching factors discussed in Chapter 2. Negative values
mean tectonic rock-uplift.

In these cases the introduction of plate boundary forces produces better model results but the physical
meaning of these forces is still a matter of debate (Zoetemeijer, 1993).
The strength of the lithosphere is an important factor in the development of foreland basins.
The distribution of strength in the lithosphere varies vertically as a result of its thermal and
compositional layered structure. Application of a depth-dependent rheological model to the continental
lithosphere implies stacking of two strong elastic layers, embedded in weak layers which yield by
brittle or ductile deformation (Burov and Diament, 1995; Cloetingh and Burov, 1996). Such a model
represents more accurately the behaviour of continental lithosphere, which is necessary for the
adequate analysis of strength and stress distribution (Zoetemeijer, 1993). However, an appropriate first
order description of the flexural behaviour of the lithosphere can be formulated assuming that it
behaves as a thin elastic plate (e.g. Turcotte and Schubert, 1982). An important independent constraint
in flexural studies is the gravity field (Karner and Watts, 1983; Zoetemeijer, 1993).

100
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera Colombian Andes

s
Serrania

de
Maracaibo

An
de Perija Basin
1400

a
Cucuta

id
er
Barinas

Lucas

M
Basin

de San
23 54

1300 4 13 10
11
50

Serrania
12 Bucaramanga
-19 0
8 0
14 29 Arauca
20 11 0
3
29 9 0
39
23
ra

1200 Medellin 17 20
d i ll e

7
36
19
19 0
40
Cor

20 7
1 25 0 0
11
-30 1 1 Llanos
30 2 25 3
7
16 13 Tunja
30 0
0
Orientales
22 32
1100 -
60 35 30 23 0
Basin (LLA)
26 0
7 4022 26 23
Yopal 0
16 33
Manizales 5 0 0
16
43
30 13
0 0 0
25
21 0 0 0
34 31 30 40 0 0
0 0

1000 Ibague 19 28 Bogota 30 0


0 0
0 0
0
0
11 22 33 34
17 14 3023 0 0 0 0
Villavicencio 0
0 0
0 0 0
25 20 0 0
0 100 200 Km
16 0 0 0
25 0
14 0
900
al

24
30 0
0
nt r

0
Ce

PREDICTED LATE
20 0
19 PALEOCENE
THERMAL
Neiva 0
800 SUBSIDENCE (m)
La Macarena
Serrania de

14
San Jose del
Guaviare

800 900 1000 1100 1200 1300 1400 1500

Figure 3.22: Predicted Late Paleocene thermal subsidence after Mesozoic rifting, in meters, calculated from forward
modelling using lithosphere-stretching factors discussed in Chapter 2. Negative values mean tectonic rock-uplift.

6.1. MODEL DESCRIPTION

The flexural response of the lithosphere to applied loads is a function of the vertical load due to
a mass distribution, the horizontal force and the effective elastic thickness (EET) of the assumed thin
elastic plate, which can change laterally. The parameter that characterizes the apparent strength of the
lithosphere is the flexural rigidity D, which is commonly expressed through the EET of the lithosphere
(Zoetemeijer, 1993). A useful parameter to judge how intense the flexural deformation of the
lithosphere is and whether the assumption of elastic deformation is still valid, is the flexural bending
stress xx or fiber stress (Ranalli, 1995), which for a thin elastic plate is given by
Ey d 2 w
xx =
1 2 dx 2 (3.2)

101
Chapter 3
Where y represents the vertical axis, x represents the horizontal axis along the profile, E represents the
Youngs modulus, represents the Poissons ratio and w the deflection. A first step in extending the
model from pure elastic deformation (Figure 3.25a) to depth-dependent multilayered decoupled
rheology (Figure 3.25b) is setting an upper bound for the applied elastic deformation (following
Turcotte and Schubert, 1982).
Assuming an elastic-plastic rheology for the bending of the plate (Figure 3.25c), the stress
increases linearly with distance y (in depth) from the center of the plate (y = 0). The plate bends
elastically until the stress at the surface becomes sufficiently large that yielding occurs. The yield stress
may be simplified to * = xx.. From these relations and using the expression for the bending moment
M(x) = -D(x)d2w/dx2 , one can determine the bending moment corresponding to the onset of plasticity
Mon for y = h (h=EET) and Mc the critical bending moment for the situation of entire yielding for
y=0: Mon = -* h2/6 and Mc= -* h2/4 (Zoetemeijer et al., 1999; for details see Turcotte and Schubert,
1982, section 7.11). All flexural models in this chapter gave fiber stress values well below the onset of
plastic yielding considered as 1GPa. following Zoetemeijer et al. (1999).

Serrania Maracaibo
de Perija

s
Basin

de
1400 Cucuta

An
Barinas

a
Lucas

id
er
Basin

M
100
a de San

59
120 139
1300 47 23
21
13
Bucaramanga
Serrani

34
- 0
28 0 Arauca
65
34
75
51 30
10 0 0

4 77
102
0

57
1200 Medellin 44 0 0
ra

15
94
d i ll e

48 0
40 12
52
67
80
60 28 0 Llanos 0
Cor

- 29 31
103 62 99
80 35 29
Tunja
79 0
Orientales
0
44 75 91
80 Basin (LLA)
1100 -
100
144
40 66
60
0 0 0
23 35
60
Yopal
54
0
Manizales 68
0
15 0
113 39
34 0 0
0 00
57 32 0 0 0 0
54 0
76 74 103 0 0 0
86
0
0 0
1000 Ibague 41 67 Bogota 80 0
0 0 0
0 0
0
80 86

41
27 51
80 60 0
0 100 200 Km
34 0
60
0
60 0
0
53 40 0
Villavicencio 0
0

46 0 0 0
35 0
61
900 0
al

71 0
n tr

0 0
0
Ce

60
0
PREDICTED
43
EOCENE THERMAL
Neiva 0 SUBSIDENCE
800
La Macarena
Serrania de

20 San Jose del (m)


35 Guaviare

800 900 1000 1100 1200 1300 1400 1500

Figure 3.23: Predicted Eocene thermal subsidence after Mesozoic rifting, in meters, calculated from forward
modelling using lithosphere-stretching factors discussed in Chapter 2. Negative values mean tectonic rock-uplift.

102
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera Colombian Andes

Serrania Maracaibo s
de Perija Basin de
30 An

Lucas
1400
ida
Cucuta

40 er Barinas
M

de San
Basin

Serrania
32 76 0 100 200 Km
-153
1300 16 -229
Bucaramanga
7

-54
16
50
8
18
41
0 0
Arauca 0
2 1
54 0 0
7 6
42
55
31
Medellin
1200 24
ra

-195 0
50
30
dill e

25 0
28
4
36 0
Cor

15 0
-85 15
17
34 53
Llanos Orientales
43
Basin (LLA)
-2 0

13 12 Tunja 0
15 0
-526
1100 -19
40 49 33
0 0
10 36
23
32 Yopal
0
Manizales 29
8 0 0
61 20
0 0 0 0
13 0
27 17 0
29 0 0 0 0
41 37 5 0 0 0
-287
6 0
0 0
00 0
0
1000 Ibague -194 -123 Bogota 0 0
0
-225 47
15 24
- 18 0
0 0 0
265 32
0

-152 0 Villavicencio
0
0
0
0
0 0

-55 10 0
0
0
0
-125 0
900 -343
al

-157 0
nt r

0 0 PREDICTED
0
Ce

OLIGOCENE
26
0 0 EARLY MIOCENE
THERMAL
Neiva
SUBSIDENCE (m)
La Macarena

800 San Jose del


Serrania de

0 Guaviare Guayana
18

Shield

800 900 1000 1100 1200 1300 1400 1500


Figure 3.24: Predicted Oligocene-Early Miocene thermal subsidence after Mesozoic rifting, in meters, calculated
from forward modelling using lithosphere-stretching factors discussed in Chapter 2. Negative values mean tectonic
rock-uplift

7. MODELLING OF FLEXURAL SUBSIDENCE PRODUCED BY THE TOPOGRAPHIC


LOAD OF THE PALAEO-CENTRAL CORDILLERA DURING PALAEOGENE TIME

In this section, I test the currently accepted hypothesis proposed in literature (e.g. Cooper et al.,
1995) that the area of the MV, palaeo-EC and LLA was a simple foreland basin related to lithosphere
flexure, produced by a topographic load represented by the palaeo-Central Cordillera. I tested this
hypothesis through modelling lithosphere flexure, assuming the lithosphere behaves as an elastic plate
with lateral variations in thickness (EET) and applying a load representing the topography of the
Palaeogene Central Cordillera. I assumed a broken plate with a free boundary located at the end of the
continental plate during Palaeogene along the Romeral suture just west of the Central Cordillera
(Figure 3.2). I modelled lithosphere flexure along four 2D regional sections and also in 3D.

7.1. 2D FLEXURAL MODELLING

I modelled the flexural lithosphere deflection in 2D sections using the program Cobra
developed by Zoetemeijer (1998). The program calculates 2D deflections for distributed loads on a

103
Chapter 3
th
plate with a variable thickness. The 4 order differential equation of the flexural behaviour is solved
using a finite difference technique (Bodine, 1981). Input data for the model are: the length of the plate,
the size of the finite difference cell, the horizontal co-ordinate of the plate boundary in case of a broken
plate, the horizontal co-ordinate of the boundary between the load and the foreland basin, densities of
the crust, mantle, mountain load and sediment filling the foreland basin (Table 3.1). Several parameters
can vary laterally: Elastic thickness of the plate, topographic load, additional surface or subsurface
loads, initial and actual water depth of the foreland basin, and initial crustal thickness. As an
independent source of information, gravity anomaly calculations help to obtain a more reliable model.
For that reason the possibility to calculate the contribution to the gravity field is included in the
program.
PERFECTLY ELASTIC RHEOLOGY

Tension Stress 0 Compression Stress


BRITTLE ELASTIC AND DUCTILE RHEOLOGY

Tension Stress 0 Compression Stress


depth

ELASTIC PERFECTLY PLASTIC RHEOLOGY


Moho
Tension Stress 0 Compression Stress

depth

*
depth

B
C

Figure 3.25: Cartoon showing different rheological models of continental lithosphere. (a) Perfectly elastic
behaviour. (b) Depth-dependent brittle-elastic-ductile rheology. (c) Elastic-perfectly-plastic rheology of simplified
continental lithosphere. (from Zoetemeijer et al., 1999).

Outputs of the program are calculated flexural deflection, gravity, and fiber stress produced by
bending of the lithosphere plate, which can be displayed as profiles along the section. Table 3.1 shows
assumptions and parameters I applied to model flexural subsidence due to topographic load of the
Central Cordillera during Palaeogene time. Those unknown values (such as topographic elevation of
the Central Cordillera during Palaeogene and elastic thickness) were modified several times by a trial
and error procedure, until reaching the optimum fit between the calculated flexural deflection of the
lithosphere and the observed subsidence during the Palaeogene.

ASSUMPTIONS

-Broken plate with a free boundary located westward on the Romeral suture zone (Fig. 3).
-No water depth (Palaeogene depositional environments were dominantly continental to transitional.
-No horizontal intraplate force.
-No vertical shear force at the free end of the plate.
-No bending moment at the free end of the plate.
-Sediment density applied as a load filling the basin when sediment thickness was used to compare to the
calculated flexural deflection.
-No sediment fill load applied when observed tectonic subsidence was used to compare to the calculated
flexural deflection.
-No subsurface loads, except for simulating thermal subsidence as explained in text.
Table 3.1a. Assumptions used in lithosphere flexural models.

104
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera Colombian Andes

PARAMETERS

-Length of the plate: 1000 km.


-Size of the finite difference grid cell: 5 km.
-Densities (kg/m3)
-Sediments: 2250
-Topographic load: 2550 to 2650
-Crust: 2770
-Mantle: 3300
-Effective elastic thickness and topographic load variable as shown in Figs. 10 to 16 and Table 2.
Table 3.1b. Parameters used in lithosphere flexural models.

7.2. 3D FLEXURAL MODELLING

I modelled the flexural lithosphere deflection in 3D using data displayed as maps with the
program Flex3D developed by Van Wees (1993). The program Flex3D calculates 3D deflections for
distributed loads on a plate with lateral changes in elastic thickness. The program was developed to
forward model the vertical flexural effects of lithosphere stretching and shortening in three dimensions,
focussing on flexural effects related to these processes. The forward model for stretching combines
main features of two dimensional basin models developed by Kooi (1991) and Kusznir et al. (1991).
Necking depths and variations in effective elastic thickness can be adopted. The differential equation
for flexural behaviour is solved in 3D applying a finite difference formulation, which is solved by a
direct matrix method. To incorporate the effect of spatial variations in rigidity, additional terms for the
bi-harmonic 3D-flexure equation have been derived from a variational displacement formulation as
used in finite-element methods. Additionally, planar faults are treated as discontinuities (Kusznir,
1991). The algorithm is described in Van Wees and Cloetingh (1994). Input data for the program are
dimensions of the modelled plate, finite difference grid size. Variables varying spatially are given to the
program as files include for example: effective elastic thickness, topography, palaeo-water depth and
crustal thickness. Intraplate stresses can also be included in the program. I used the same parameters
applied to 2D flexural models (Table 3.1).

7.3. RESULTS

7.3.1. 2D models

Initially I tested the present day topography of the Central Cordillera as if it had remained
unchanged since Palaeogene times. Figure 3.26 shows the results of these models. Clearly the predicted
subsidence is many times smaller than the observed subsidence. Thus if the subsidence was produced
by the Central Cordillera, a much higher topography than present day would be necessary. Additionally
the shape of the calculated deflected profile is different to the observed profile of subsidence.
In a second trial I assumed only a topographic load in the palaeo-Central Cordillera and I
modified it and the effective elastic thickness until it produced the best fit with the observed
subsidence. Figure 3.27 shows the results of these models. In all the cases the shape of the calculated
subsidence profile is different to the shape of the observed subsidence profile. Only for the most eastern
LLA area both profiles are similar. To produce the observed subsidence in the LLA area by flexural
deflection, it would be necessary to apply 5000 m of topography for the palaeo-Central Cordillera and
an elastic thickness of 200 km. These values are completely unrealistic: an elastic thickness of 200 km
would imply a lithosphere several times stronger than stable cratonic lithosphere (e.g. Cloetingh and
Burov, 1996). These unrealistic results indicate that the observed Palaeogene subsidence in the palaeo-
EC and neighbouring MV and LLA areas cannot be explained only by flexural subsidence of the
lithosphere due to the Palaeogene topography of the Central Cordillera. The observed Palaeogene
subsidence in the Magdalena and LLA areas increase to a maximum close to the borders of the former
extensional basin. Only a load located in the Central Cordillera could never produce this.

105
Chapter 3

e
The asymmetric profile of subsidence, with larger subsidence in the western Magdalena area
e
than in the eastern LLA area, probably was due to the load of the palaeo-Central Cordillera. However,
in order to reproduce a deflection as the observed, it is necessary to adopt some additional load in the
area of the former extensional basin. One possible load in such an area would be remaining thermal
subsidence of the former extensional basin. However, we have already tested such a hypothesis (section
5 of this Chapter) and we found that thermal subsidence alone cannot explain the observed
backstripped tectonic subsidence. An alternative hypothesis is to consider that the original extensional
basin was partially inverted during Palaeogene, generating some amount of local topography that
would produce flexural deflection in the lithosphere, with maximum values close to the borders of the
partially inverted extensional basin. I also tested this last hypothesis through modelling.

e
SECTION 1 SECTION 5

FLE XU R E FLE XU R E
5000
l ec i n m

5000

0
0

e
-5000
-5000
d

d
-10000
-10000
0 200 400 600 800 1000 1200
o

0 200 400 600 800 1000 1200

EFF E C T IVE EL AST IC T H I CKN ESS 80 EFFE C T IV E EL AST IC T H IC K NES S


80
e

60
60
40
40
20
) e

20
0
0
0 200 400 600 800 1000 1200 0 200 400 600 800 1000 1200
dist ance (km ) distan ce (km )

SECTION 3 SECTION 1 2
FLE XU R E FLE XU R E
5000
500

0
0

-5000
-500

-10000
-1000
0. 200. 400. 600. 800. 1000. 1200. 0. 200 400. 600. 800. 1000. 1200.
.
EFFE C T I VE ELAS TIC TH ICK NESS 80 EFFECT IVE E LASTI C T H I CKNESS
80
60
60
40
40
20
20
e

0
0
0. 200. 400. 600. 800. 1000. 1200. 0. 200. 400. 600. 800. 1000. 1200.
distan ce (km ) distan ce (km )
Figure 3.26: 2D lithosphere flexural models only assuming the present day topography load of the Central Cordillera
as if it were remained unchanged since Palaeogene times. Top panel: basement deflection and topography in meters.
Dots represent the observed thickness of Oligocene-Early Miocene deposits in meters. Bottom panel: effective
elastic thickness in km. In order to try to fit the observed deflection and taking into account the post-rifting thermal
weakening of the EC area reduced elastic thickness was assumed in this area. However in these models it was not
possible to fit the observed deflection. Figure 3.2 shows location of these sections.

106
t
(

t
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera Colombian Andes

SECTIO N 1 SECTIO N 5

FL E XU R E FL E XU R E
m)

5000 5000
d e f le c tei o n

0 0

-5000 -5000

0 500 1000 1500 0 500 1000 1500


EF FE C TI V E E L AS TI C TH I C K N ES S EFFECT IVE E LASTI C T H I CKNES S
200 200
e t ( ( m)

150 150

e
100 100

50 50

e
0 0
0 500 1000 1500 0 500 1000 1500
distance ( k m ) dist ance (km )

SECTIO N 3 SECTIO N 1 2

FL E XU R E FL E XU R E
m)

5000 4000

2000
d e f le c t i o n

0
0

-5000 2000

0 500 1000 1500 0 500 1000 1500


EFF EC TI VE EL AST IC TH IC KN ESS EF FE C T IV E E LA S TI C T H I C K N E SS
200 200
150 150

100 100

50 50
e

0 0
0 500 1000 1500 0 500 1000 1500
distan ce (km ) dist ance (km )

Figure 3.27: 2D lithosphere flexural models assuming only a topographic load in the palaeo-Central Cordillera that
best fits the observed thickness of Oligocene-Early Miocene deposits. Top panel: basement deflection and
topography in meters. Dots represent the observed thickness of Oligocene-Early Miocene deposits in meters. Bottom
panel: effective elastic thickness in km. Figure 3.2 shows location of these sections.

7.3.2. 3D model

As an additional test I modelled in 3D the flexural deflection of the lithosphere produced by a


topographic load in the Central Cordillera. Figure 3.28 shows the calculated flexural deflection, which
monotonously increases westward producing a deflection comparable to the asymmetric regional long
wave component of observed Palaeogene subsidence. However the model did not predict the observed
local maximum subsidence in the former extensional basin and the LLA area, nor the short wave length
component of the observed Palaeogene subsidence (e.g. compare Fig. 3.28 with Fig. 3.20).

107
Chapter 3

Serrania Maracaibo
de Perija Basin

cas
1400

s
Cucuta

de
n Lu

An
Barinas

e Sa

a
Basin

id
er
nia d

M
Serra
1300 Bucaramanga

Arauca

1200 Medellin
ra
d i ll e
Cor

Llanos
Tunja Orientales
Basin (LLA)
1100
Yopal

Manizales

1000 Bogota 0 100 200 Km


Ibague

Villavicencio

900 LITHOSPHERE
al
n tr

FLEXURE PRODUCED
Ce

BY A UNIFORM LOAD
ON THE CENTRAL
Neiva
CORDILLERA
La Macarena
Serrania de

800 (m)
San Jose del
Guaviare

800 900 1000 1100 1200 1300 1400 1500

Figure 3.28: 3D lithosphere flexural model. 3D flexural deflection, in meters, of the lithosphere produced only by a
topographic load located in the palaeo-Central Cordillera. See these results with thickness maps of Oligocene-Early
Miocene deposits for a comparison (Fig. 3.13).

8. MODELLING OF FLEXURAL SUBSIDENCE PRODUCED BY THE TOPOGRAPHIC


LOAD OF THE PARTIALLY INVERTED EXTENSIONAL BASIN DURING
PALAEOGENE TIME

In this section I test the hypothesis of flexural subsidence produced by some local surface-
uplifted blocks (incipient basin inversion, Gomez et al., 1999) within the former extensional basin, as
suggested by exhumation of some areas indicated by local sedimentological (e.g. Anderson, 1970;
1972; Laverde, 1989; Ecopetrol-ICP, 1996; Gmez et al., 1999), petrographical (e.g. Porta, 1966;
Anderson, 1970, 1972), mineralogical (Cheilletz et al., 1993, 1997; Branquet et al., 1996), structural
(e.g. Julivert, 1970; Butler and Schammel, 1988; Casero et al., 1995, 1997; Restrepo-Pace, 1999a,b;
Gmez et al., 1999) and fission track data (e.g. Shagam et al., 1984; Toro, 1990; Van der Wiel, 1991;
Hossack et al., 1999; Gmez et al., 1999).
I model flexural subsidence in 2D, along four regional sections, produced by only the
topographic load of the partially inverted extensional basin during Palaeogene time using the same
methods, parameters and software.

108
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera Colombian Andes
8.1. RESULTS

Figure 3.29 shows that to generate a flexural deflection similar to the observed subsidence
would require a topographic load in the area of the former extensional basin. Such topographic load
would be approximately 2000 m high, with effective elastic thickness of 60 km in the LLA areas,
decreasing westward to values of approximately 20 to 25 km in the palaeo-EC and MV area. Although
these values are theoretically possible (e.g. Cloetingh and Burov, 1996), a topography of 2000 m for
the palaeo-EC during Palaeogene time is in conflict with the presence of a Palaeogene sedimentary
record in the axial zone of the Cordillera (Sabana de Bogot, Tunja, and Sogamoso regions). The local
evidence of marine sedimentation in the Oligocene Concentration Fm also contradicts this hypothesis.
If such topography of 2000 m had existed, there would be abundant evidence of it in the sedimentary
record, in the palaeocurrents pattern and detrital mineralogical composition (in the neighbouring LLA
and Magdalena basins), that is not the case.
Additionally many authors (Van der Hammen et al., 1973; Helmens, 1990; Hooghiemstra,
1984, 1989) have reported palynological evidence of the surface-uplift of the Sabana de Bogot area in
the last 5 Ma, which implies that the area was low during Palaeogene time. Scarce fission track data do
not support the hypothesis of 2000 m of topography of the palaeo-EC during Palaeogene. If some early
inversion of the former extensional basin occurred during Palaeogene, it would have generated a much
lower local topography, probably limited to the borders of the former extensional basin without
disturbing so much the sedimentary and palaeocurrents pattern in the LLA, MV and axial region of the
palaeo-EC. A lower basin inversion topography of the former extensional basin alone is not enough to
produce the observed Palaeogene subsidence. Thus additional components including topographic load
of the Central Cordillera as well as remaining thermal subsidence after Mesozoic-rifting events can be
proposed to explain the observed Palaeogene subsidence.

9. MODELLING OF SUBSIDENCE PRODUCED BY THE COMBINED EFFECT OF (a)


REMAINING THERMAL SUBSIDENCE AFTER MESOZOIC RIFTING, (b) FLEXURAL
SUBSIDENCE PRODUCED BY THE TOPOGRAPHIC LOAD OF THE PALAEO-
CENTRAL CORDILLERA AND (c) FLEXURAL SUBSIDENCE PRODUCED BY LOCAL
TOPOGRAPHY DUE TO PARTIAL INVERSION OF THE MESOZOIC EXTENSIONAL
BASIN

As none of the hypotheses alone is adequate to explain the observed tectonic subsidence during
Palaeogene times, I also tested whether a combination of all hypotheses to see if the combination of
them could explain the observed tectonic subsidence. I modelled subsidence produced by the combined
effect of: (a) remaining thermal subsidence after Mesozoic rifting, (b) flexural subsidence produced by
topographic load of the palaeo-Central Cordillera and (c) flexural subsidence produced by local
topography within the former extensional basin. To this aim I performed 2D and 3D modelling
experiments.

9.1. 2D MODELS

Along 4 regional cross-sections (sections 2, 5, 8 and 12, see location in Figure 3.2) I applied the
following procedure: (a) From a series of stratigraphic columns or wells close to each section (Figure
3.2), I calculated the observed tectonic subsidence using backstriping techniques, for six different time
intervals from late Maastrichtian to Early Miocene; (b) I calculated also the remaining thermal
subsidence after Mesozoic rifting events using forward modelling and all the parameters obtained from
the study of Mesozoic rifting (Chapter 2); (c) Assuming local isostasy for each stratigraphic column or
well and for each time interval, I calculated a subsurface load that would produce the same tectonic
subsidence as the remaining thermal subsidence alone; (d) I modelled in 2D the flexural deflection of
the lithosphere.
For each regional cross-section and for each time interval, I modelled the effect of three
different loads: (1) A subsurface load that would produce a subsidence effect equal to the remaining

109
Chapter 3
thermal subsidence; (2) A surface topographic load representing Palaeogene topography of the palaeo-
Central Cordillera; and (3) Surface loads representing Palaeogene topography in some local areas of the
former extensional basin. In this way I include in the flexural model the combined effects of remaining
thermal subsidence after Mesozoic rifting, and flexural subsidence produced by loads in the palaeo-
Central Cordillera and by local surface-uplifted regions related to basin inversion processes. I used the
same values used in previous models for densities, crustal thickness, etc. (see Table 3.1). I
systematically changed the magnitude of the topographic load of the palaeo-Central Cordillera and both
the location and magnitude of the topographic load of the former extensional basin as well as the values
of effective elastic thickness until I reached the best fit between the observed and predicted Palaeogene
subsidence.
SECTION 5
FLEXURE FLEXURE

4000. 4000.
deflection (m)

deflection (m)
2000. 2000.
0. 0.
-2000. -2000.
-4000.
-4000.
-6000.
0. 500. 1000. 1500. 0. 500. 1000. 1500.

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


80. 80.
eet (km)

eet (km)

60. 60.
40. 40.
20. 20.
0. 0.
0. 500. 1000. 1500. 0. 500. 1000. 1500.
distance (km) distance (km)

SECTION 3 SECTION 12
FLEXURE FLEXURE
4000.
deflection (m)

deflection (m)

2000.
2000.

0. 0.

-2000.
-2000.
-4000.
-4000.
0. 500. 1000. 1500. 0. 500. 1000. 1500.

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


80. 80.
eet (km)

eet (km)

60. 60.
40. 40.
20. 20.
0. 0.
0. 500. 1000. 1500. 0. 500. 1000. 1500.

distance (km) distance (km)

Figure 3.29: 2D lithosphere flexural models assuming a positive topography restricted to the former extensional
basin. The assumed topography fits best the observed thickness of Oligocene-Early Miocene deposits. Top panel:
basement deflection and topography in meters. Dots represent the observed thickness of Oligocene-Early Miocene
deposits in meters. Bottom panel: effective elastic thickness in km. Figure 3.2 shows the location of these sections.

110
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera Colombian Andes
9.1.1. Results

The best results are presented in Figure 3.30 to 3.33. Although the fit is not perfect I
systematically changed each variable and exhaustively tested many possibilities until I became
convinced I had arrived to the best possible fit. The differences between the observed and predicted
subsidences are short wave length effects that probably were produced by local relative rock-uplift or
subsidence in fault-limited blocks. As such effects are not considered in the theoretical model, the
model can obviously not predict them. The subsidence pattern of a partially inverted basin is
characterized by short wave local subsidence variations produced by relative movement of faulted
blocks. Table 3.2 shows the values and location of topographic loads as well as of the spatially variable
effective elastic thickness necessary to produce the best fit between the observed and predicted
subsidence. In order to approach the short wave length pattern of observed subsidence, it was necessary
in many cases to apply two topographic loads close to the border faults limiting the former extensional
basin. However, the number and location of loads is limited in the program. This pattern suggests that
this topography could have been generated by early partial inversion of the border faults limiting the
former extensional basin. I also calculated the bending stress in the adopted thin elastic plate (Figs 3.30
to 3.33). Bending stresses in all cases are within reasonable values under normal rock strength
assumptions.

Distance Maastrichtian Oligocene


from Early Paleocene Late Paleocene Eocene Early Miocene
Sub-basin 60.9 Ma 54.8 Ma 33.7 Ma 16.4 Ma
Romeral
(km)
EET * Topo EET Topo EET Topo EET Topo
(km) (m) (km) (m) (km) (m) (km) (m)
SECTION 2

Central Cordillera 0 1 100 10 100 10 0 10 100


Middle Magdalena Valley 270 1 100 10 100 10 0 10 100
" " 271 1 0 1 0 5 0 5 0
Tablazo sub-basin 350 1 0 1 0 5 0 5 0
" " 351 1 0 1 100 5 0 5 200
" " 355 1 0 1 100 5 0 5
" " 380 0 100 5 0 5 0
Santander Massif 416 5 0 5 100 5 0 5 0
" " 417 5 0 5 100 5 0 5 0
" " 424 5 0 5 100 5 0 5 0
Cocuy sub-basin 437 5 0 5 0 5 0 5 0
" " 438 5 0 5 0 5 0 5 0
" " 509 5 0 5 0 5 0 5 0
Llanos foothills 523 5 0 5 0 5 0 5 200
" " 524 1 0 5 0 50 0 60 200
Llanos Orientales 555 5 0 5 0 50 0 60 200
" " 556 50 0 60 0 50 0 60 0
" " 800 50 0 60 0 60 0 60 0
SECTION 5

Central Cordillera 0 1 0 10 100 10 0 10 100


Middle Magdalena Valley 257 1 0 10 100 10 0 10 100
" " 258 5 0 5 0 5 0 5 0
* EET: effective elastic thickness, Topo: palaeotopography
Table 3.2: Effective elastic thickness and palaeotopography loads that best fit the observed subsidence
during four different Palaeogene time intervals

111
Chapter 3

Distance Maastrichtian Oligocene


from Early Paleocene Late Paleocene Eocene Early Miocene
Sub-basin 60.9 Ma 54.8 Ma 33.7 Ma 16.4 Ma
Romeral
(km)
EET * Topo EET Topo EET Topo EET Topo
(km) (m) (km) (m) (km) (m) (km) (m)
Tablazo sub-basin 300 5 0 5 0 5 0 5 0
" " 301 5 0 5 100 5 0 5 100
Cocuy sub-basin 367 1 0 5 100 5 0 5 100
" " 368 1 0 5 5 0 5 0
" " 402 1 0 1 1 0 1 0
" " 405 1 0 1 1 0 1 0
" " 417 1 0 50 0 0
" " 445 1 0 50 0 0
" " 446 1 0 50 0 500
Llanos foothills 452 1 0 5 50 5 0 5 500
" " 453 1 0 5 0 5 0 5 500
" " 460 0 5 0 5 0 5 500
" " 480 10 0 0 5 0 5 500
" " 481 50 0 0 5 0 5 500
Llanos Orientales 485 50 0 50 0 5 0 5 500
" " 486 50 0 50 0 50 0 60 0
" " 800 50 0 50 0 50 0 60 0
SECTION 8

Central Cordillera 0 10 200 10 100 10 0 10 100


Middle Magdalena Valley 228 10 200 10 100 10 0 10 100
" " 229 1 200 5 0 5 0 5 0
Tablazo sub-basin 260 1 0 5 0 5 0 5 0
" " 262 1 0 5 0 5 0 5 0
" " 263 1 0 5 200 5 0 5 0
Cundinamarca sub-basin 293 1 0 5 200 5 0 5 0
" " 294 1 0 5 200 5 0 5 0
" " 310 1 0 5 200 5 0 5 0
" " 311 1 0 5 0 5 0 5 0
" " 326 1 0 5 0 5 0 5 0
" " 330 0 5 0 5 0 5 0
" " 367 0 5 0 5 0 5
" " 369 0 5 200 5 0 5
" " 394 0 5 200 5 0 5
" " 395 0 5 0 5 0 5
" " 407 0 5 0 5 0 5 1000
" " 408 0 5 0 5 0 5 0
Llanos Orientales 428 5 0 5 0 5 0 5 0
" " 429 50 0 50 0 50 0 50 0
" " 800 50 0 50 0 50 0 50 0
SECTION 12

Central Cordillera 0 5 100 10 100 10 0 10 100


Southern Cundinamarca 200 5 100 10 100 10 0 10 100
and Cocuy sub-basins 201 5 0 5 0 5 0 5 0
" " 231 5 0 5 0 5 0 1 0
" " 265 5 0 3 0 5 0 1 0
" " 303 5 0 3 0 5 0 0
" " 304 5 0 3 300 5 0 0
* EET: effective elastic thickness, Topo: palaeotopography.

Table 3.2: Continued

112
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera Colombian Andes

Distance Maastrichtian Oligocene


from Early Paleocene Late Paleocene Eocene Early Miocene
Sub-basin 60.9 Ma 54.8 Ma 33.7 Ma 16.4 Ma
Romeral
(km)
EET * Topo EET Topo EET Topo EET Topo
(km) (m) (km) (m) (km) (m) (km) (m)
Quetame Massif 318 5 0 3 5 0 5 0
" " 319 5 0 3 50 0 5 700
" " 345 5 0 3 0 50 0 5 700
Llanos Orientales 350 5 0 3 0 50 0 5 700
" " 351 50 0 3 0 50 0 60 0
" " 354 50 0 3 0 50 0 60 0
" " 360 50 0 50 0 50 0 60 0
" " 800 50 0 50 0 50 0 60 0
* EET: effective elastic thickness, Topo: palaeotopography.

Table 3.2: Continued

Late Maastrichtian-early Paleocene: (Fig. 3.30). To fit the total tectonic subsidence that
occurred during this time interval, it was necessary to apply topographic loads only in the area of the
Central Cordillera. This result suggests that during this time the former extensional basin was not yet
inverted, but it was still part of a foreland basin comprising the MV, palaeo-EC and LLA areas. The
maximum amount of topography needed to fit observed data is 200 m, indicating a low topography, but
it varies from north to south between 0 and 200 m (see Figures 3.30 and 3.3). Values for effective
elastic thickness are very small (1 to 5 km) in the area of the former extensional basin, suggesting a
weak lithosphere probably as a result of the Mesozoic rifting events. Also the inferred effective elastic
thickness in the area of the Palaeo-Central Cordillera suggests a weak lithosphere. Only in the eastern
LLA area values of effective elastic thickness of 50 km indicate the presence of strong lithosphere.
These results can explain the fact that further deformation affected weak palaeo-Andean lithosphere but
not the stronger LLA lithosphere. Horizontal stresses may also have had a significant role in the
flexural history of the EC and on the lithosphere strength (Burov and Diament, 1995). However, in the
actual situation, models do not require plate boundary forces or subsurface loads, and for simplicity I
neglected their effect.
Late Paleocene: (Fig. 3.31) Results suggest low topography: up to 100 m in the area of the
Central Cordillera and also locally along the borders of the former extensional basin, with values up to
200 m. (Figures 3.31 and 3.5). Early slight inversion of the border faults limiting the former extensional
basin probably generated this topography, which did not significantly disturb the sedimentary and
palaeocurrent pattern in the basin. The occurrence of local unconformities as reported by Sarmiento
(1992, 1993) in the Guaduas Fm, might be related to this topography. Effective elastic thickness in the
area of the Central Cordillera increased to 10 km, probably as a result of earlier deformation. The
former extensional basin lithosphere remained weak (effective elastic thickness between 1 and 5 km),
and prone to further deformation, whereas the eastern LLA lithosphere remained relatively strong
(Effective elastic thickness of 50 km).
Eocene: (Fig 3.32) Models do not require topography at this time interval. If some surface-
uplift occurred, it was probably rapidly eroded (as indicated by evidences of erosion truncation of
earlier structures in the Middle MV). The map of the early Eocene unconformity prepared by George et
al. (1997), based on of geological maps and subsurface data (Fig. 3.6), shows: (a) a highly variable

113
Chapter 3

MAASTRICHTIAN EARLY PALEOCENE

SECTION 2 SECTION 8
FLEXURE FLEXURE
500.
deflection (m)

deflection (m)
100. 0.

0. -500.

-100. -1000.

-1500.
0. 200. 400. 600. 800. 1000. 0. 200. 400. 600. 800. 1000.

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


80. 80.
eet (km)

eet (km)
60. 60.
40. 40.
20. 20.
0. 0.
0. 200. 400. 600. 800. 1000. 0. 200. 400. 600. 800. 1000.
distance (km) distance (km)

SECTION 5 SECTION 12
FLEXURE FLEXURE
400. 400.
deflection (m)

deflection (m)

200. 200.

0. 0.

-200. -200.
-400.
-400. -600.

0. 500. 1000. 0. 200. 400. 600. 800. 1000.

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


80. 80.
eet (km)

eet (km)

60. 60.
40. 40.
20. 20.
0. 0.
0. 500. 1000. 0. 200. 400. 600. 800. 1000.
distance (km) distance (km)

Figure 3.30: Late Maastrichtian-early Paleocene 2D lithosphere flexural models assuming the combined effect of:
(1) remaining thermal subsidence after Mesozoic rifting; (2) topographic load of the palaeo-Central Cordillera; and
(3) local palaeotopography due to partial inversion of the Mesozoic extensional basin. Top panel: basement
deflection and topography in meters. Dots represent the observed thickness of late Maastrichtianearly Paleocene
deposits in meters. Bottom panel: effective elastic thickness in km. Figure 3.2 shows location of these sections.

angular unconformity in the area of the Middle MV; (b) Truncated homoclines or broad relief folds in
the area of the Upper MV and possibly (?) in the Santander Massif and borders of the original
extensional basin; (c) Exhumated blocks with no deformation or very modest deformation in the former
extensional basin; and (d) A paraconformity in the eastern LLA area. Values of effective elastic
thickness required to fit the models (Fig. 3.32), suggest the weakest lithosphere was located in the area
of the former extensional basin (up to 5 km), with relatively weak lithosphere in the Central Cordillera
area (10-km), and strong lithosphere in the eastern LLA area (50-km).

114
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera Colombian Andes
LATE PALEOCENE

SECTION 2 SECTION 8
FLEXURE FLEXURE
400.
200.
deflection (m)

deflection (m)
200.
0.
0.

-200. -200.

-400. -400.

-600. -600.
0. 500. 1000. 0. 200. 400. 600.

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


80. 80.
eet (km)

eet (km)
60. 60.
40. 40.
20. 20.
0. 0.
0. 500. 1000. 0. 200. 400. 600.
distance (km) distance (km)

SECTION 5 SECTION 12
FLEXURE FLEXURE
400.
200.
deflection (m)

deflection (m)

200.
0.
0.
-200. -200.
-400.
-400.
-600.
-600.
0. 500. 1000. 0. 500. 1000.

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


80. 80.
eet (km)

eet (km)

60. 60.
40. 40.
20. 20.
0. 0.
0. 500. 1000. 0. 500. 1000.
distance (km) distance (km)

Figure 3.31: Late Paleocene 2D lithosphere flexural models assuming the combined effect of: (1) remaining thermal
subsidence after Mesozoic rifting; (2) topographic load of the palaeo-Central Cordillera; and (3) local
palaeotopography due to partial inversion of the Mesozoic extensional basin. Top panel: basement deflection and
topography in meters. Dots represent the observed thickness of late Paleocene deposits in meters. Bottom panel:
effective elastic thickness in km. Figure 3.2 shows location of these sections.

Oligocene - early Miocene: (Fig. 3.33) in order to fit the data, the models require low
topography in the area of the Central Cordillera (100 m) and the borders of the former extensional basin
(Figs. 3.33 and 3.13). The required topography would be about 100 m in the western border of the
basin, and 200 m in the eastern border of the basin. Locally in the Quetame Massif area 1000 m of
topography would be necessary. If such topography existed, it was probably generated by inversion of
the former extensional faults limiting the Mesozoic basin. However, such topography probably was not
sufficiently continuous to significantly disturb the sedimentary and palaeocurrent patterns. Effective

115
Chapter 3
elastic thicknesses suggest weakest lithosphere (1 to 5 km), prone to further deformation was located in
the area of the former extensional basin, with weak lithosphere in the Central Cordillera area (10 km)
and strong lithosphere (50 to 60 km) in the eastern LLA area.
EOCENE

SECTION 2 SECTION 8
FLEXURE FLEXURE

200.
deflection (m)

deflection (m)
200.
100.
0. 0.

-100.
-200.
-200.

0. 500. 1000. 0. 500. 1000.

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


80. 80.
eet (km)

eet (km)
60. 60.
40. 40.
20. 20.
0. 0.
0. 500. 1000. 0. 500. 1000.
distance (km) distance (km)

SECTION 5 SECTION 12
FLEXURE FLEXURE
150.
deflection (m)

deflection (m)

200. 100.
50.
0. 0.
-50.
-200. -100.
-150.

0. 500. 1000. 0. 500. 1000.

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


80. 80.
eet (km)

eet (km)

60. 60.
40. 40.
20. 20.
0. 0.
0. 500. 1000. 0. 500. 1000.
distance (km) distance (km)
Figure 3.32: Eocene 2D lithosphere flexural models assuming the combined effect of: (1) remaining thermal
subsidence after Mesozoic rifting; (2) topographic load of the palaeo-Central Cordillera; and (3) local
palaeotopography due to partial inversion of the Mesozoic extensional basin. Top panel: basement deflection and
topography in meters. Dots represent the observed thickness of Eocene deposits in meters. Bottom panel: effective
elastic thickness in km. Figure 3.2 shows location of these sections.

116
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera Colombian Andes

OLIGOCENE EARLY MIOCENE

SECTION 2 SECTION 8
FLEXURE FLEXURE
400.
1000.
deflection (m)

deflection (m)
200.
0.
0.

-200. -1000.

-400. -2000.
-600.
0. 500. 1000. 0. 500. 1000.

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


80. 80.
eet (km)

eet (km)
60. 60.
40. 40.
20. 20.
0. 0.
0.
0. 500. 1000. 500. 1000.
distance (km) distance (km)

SECTION 5 SECTION 12
FLEXURE FLEXURE
1000.
deflection (m)

deflection (m)

500. 500.

0.
0.

-500.
-500.
-1000.
-1000.

0. 500. 1000. 0. 500. 1000.

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


80. 80.
eet (km)

eet (km)

60. 60.
40. 40.
20. 20.
0. 0.
0. 500. 1000. 0. 500. 1000.
distance (km) distance (km)

Figure 3.33: Oligocene-Early Miocene 2D lithosphere flexural models assuming the combined effect of: (1)
remaining thermal subsidence after Mesozoic rifting; (2) topographic load of the palaeo-Central Cordillera; and (3)
local palaeotopography due to partial inversion of the Mesozoic extensional basin. Top panel: basement deflection
and topography in meters. Dots represent the observed thickness of Oligocene-Early Miocene deposits in meters.
Bottom panel: effective elastic thickness in km. Figure 3.2 shows location of these sections.

9.2. 3D MODEL

Using the results of the 2D models, I carried out a 3D model experiment for the Oligocene
Early Miocene time interval applying topographic loads in the palaeo-Central Cordillera and the
borders of the original extensional basin, as suggested by the 2D results, to calculate the flexural
deflection of the lithosphere. Figure 3.8b shows the calculated deflection. Comparison of this result
(Figure 3.34) with the observed subsidence pattern (Fig. 3.13) shows a better fit than the result obtained

117
Chapter 3
in a similar 3D model without applying topography loads to the borders of the former extensional basin
(Fig. 3.28).

Serrania Maracaibo
de Perija Basin

s
1400 Cucuta

de
An
Barinas

Lucas

a
Basin

id
er
de San

M
1300 Serrania Bucaramanga

Arauca

1200 Medellin
ra
d i ll e

Llanos
Cor

Tunja Orientales
Basin (LLA)
1100 100
Yopal
0
Manizales

1000 Ibague Bogota


-
0 100 200 Km

Villavicencio 300

900 LITHOSPHERE
al

200 FLEXURE PRODUCED


ntr

BY A UNIFORM LOAD
Ce

ON THE CENTRAL
CORDILLERE, TABLAZO
Neiva 100 AND COCUY (FOOTHILLS)
La Macarena

800
Serrania de

San Jose del INVERTED SUB-BASINS


Guaviare (m)

800 900 1000 1100 1200 1300 1400 1500

Figure 3.34: 3D lithosphere flexural models. 3D flexural defection of the lithosphere produced by a combined load
represented by topography in the palaeo-Central Cordillera and also in the inverted Tablazo-Magdalena and Cocuy
sub-basins of the former extensional domain. See these results with thickness maps of Oligocene-Early Miocene
deposits for a comparison (Fig. 3.13).

As a conclusion, this result confirms the 2D models, also suggesting development of some local
topography along the borders of the former extensional basin, probably as a result of inversion of the
original extensional faults limiting the basin. If some extensional faults were reactivated it was
probably because their strength was lower than normal rock strength. Calculated bending stresses
assuming a thin elastic plate are lower than normal rock strength values. Such topography was low and
did not significantly disturb the sedimentary and palaeocurrent pattern in the Palaeogene basin.
Modelling results require very low values of effective elastic thickness in the area of the former
extensional basin. The Mesozoic rifting events reduced significantly the strength of the lithosphere,
making it very prone to Palaeogene deformation and to further Andean deformation. Low values of
effective elastic thickness are usually associated with fault-controlled upper crustal flexure patterns
(Van Wees and Cloetingh, 1994). These authors have shown that superposition of low EET values with
high EET flexure effects, supports a multilayered rheological control on continental rifting. Probably
similar effects also are applicable to the early basin inversion processes in the palaeo-EC of Colombia.

118
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera Colombian Andes
The differences between the observed and predicted subsidence are short-wavelength effects
probably produced by local relative surface-uplift of fault limited blocks.

10. COMPARISON OF MODEL RESULTS WITH FISSION TRACK DATA AND OTHER
EVIDENCE OF UPLIFT AND DEFORMATION

Flexural models suggest local palaeotopography (surface-uplift), while erosion-related


unconformities and petrographic evidence of erosion of some areas indicate local exhumation. If
During Palaeogene time in the study area, local uplift of the earths surface occurred and also
denudation/exhumation occurred, the amount of rock-uplift must be greater than the surface-uplift
estimated from flexural models (see equation 3.1, England and Molnar, 1990). Are the model results
supported by additional evidence of exhumation? To answer this question I summarise from literature
fission track data and other evidence of exhumation and deformation. Finally I compare model results
with these data.

10.1. FISSION TRACK DATA AND OTHER EVIDENCE OF EXHUMATION AND


DEFORMATION

Although fission track data represent cooling and denudation, they can be related to
denudation/exhumation in combination with knowledge of the geothermal gradient (e.g. Andriessen,
1995; Rohrman, 1995). Figure 3.35 shows the distribution of fission track data and
deformation/exhumation evidence in northwestern South America, compiled from literature. Numbers
illustrate the localities on Figure 3.35. In most of the mentioned cases fission track data must be
interpreted with great care because track lengths have not been measured (e.g. Andriessen, 1995).
1. Mrida Andes (Venezuela). Kohn et al. (1984) studied apatite, zircon and sphene fission
tracks from 45 samples of the Merida Andes in Venezuela. Zircon ages range from 60 Ma to 172 Ma
with a strong group in the range of 81 to 113 Ma. These authors interpreted zircon ages older than 61
Ma as mixed ages. Two sphene ages at 139 and 159 Ma also were interpreted as mixed. However, Van
der Wiel (1991), Van der Wiel and Van der Bergh (1992a,b) and Van der Wiel and Andriessen (1991)
reinterpreted some of the 81 to 113 Ma ages as related to a tectonic event that produced a small amount
of exhumation.
Zircon ages of 60 and 61 Ma are interpreted as exhumation ages suggesting initial rock-uplift
during early Paleocene. According to these authors this rock-uplift was sufficient to make the Mrida
Andes positive but not enough to bring the basement rocks through the closure temperature for apatite
(Kohn et al., 1984). However this is against the petroleum occurrences in the Barinas Basin, which are
assumed to derive from the west of the Mrida Andes (Roure, personal communication).
Apatite ages range from 24 Ma to 1.4 Ma. Kohn et al. (1984) interpreted them as cooling ages
related to exhumation and erosion of cover rocks. Based on the apatite ages, they also suggested that
exhumation of the Mrida Andes started in the leading (northwestern, location 1a see Fig. 3.35) margin
during the Oligocene to Miocene, followed by exhumation of the trailing (southeastern, location 1b see
Fig. 3.35) margin in the Late Miocene, and later by rapid exhumation of the central axial Andes
(location 1c see Fig. 3.35) during Pliocene-Pleistocene time. A plot of apatite ages versus sample
elevation approximates a straight line suggesting an exhumation of approximately 2300-m during
Pliocene. These authors assume updoming of isotherms, because the estimated exhumation rate of 800
m/Ma exceeds accepted rates of thermal diffusion in rocks. Also they suggested that the time interval of
exhumation is likely to be greater than indicated by the range of ages.
2. Sierra de Perij. Shagam et al. (1984) studied apatite, zircon and sphene fission tracks from
9 samples from this area. Zircon ages range from 69 to 127 Ma. One sample gives the crystallisation
age of felsic volcanics (120- 122 Ma), whereas 2 or 3 samples are interpreted as indicative of
exhumation at the end of Cretaceous-Paleocene time and the remaining are interpreted as mixed ages.
Apatite ages range between 2.7 Ma and 27.2 Ma being interpreted as exhumation during late Oligocene
(27-22 Ma) in the southeast piedmont (location 2a see Fig. 3.35) followed by exhumation during early
to middle Miocene (19-14 Ma) in the western piedmont (location 2b see Fig. 3.35) and Middle Pliocene

119
Chapter 3
(3 Ma) in the whole Sierra de Perij (Shagam et al., 1984). One sphene age of 113 Ma was interpreted
as a mixed age.

Toas Islands
Zircon: 120 to 93 Ma.
Apatite: 13 to 12 Ma
Caribbean Sea Shagam et al. (1984)

Sierra Nevada
de Santa
Marta
2b
2c
2a Maracaibo
Sierra
de Perija Basin
Pa Merida Caribbean Mountains
na Andes 1a Zircon: 49+6, 42+5.3, 19.7+2.1 Ma.
m 1c Apatite: 6.1+1.3 M/a.
a

V)
3a 1b Kohn et al. (1984)
y (M Barinas
3b Basin Venezuela
alle
ra
dille

na V

EC)
3c
Cor

ra (
dale
ordillera
Pacific Ocean

dille

8
Mag

)
LA
4
Cor

(L
6 5 les
W estern C

tern

nta
Eas

rie
sO

7
tral

no
Lla
Cen

9a Guyana
9b
9c Shield

Putumayo
Basin

0 100 200 km
Ecuador

Napo and Cutucu Mountains


Cooling and uplift events: 40 and 12 Ma. Brazil
Marksteiner and Aleman (1997)
Cordillera Real and Occidental
Zircon ages: Cretaceous to Miocene
Apatite ages: Paleocene to Miocene
Steinmann et al (1996)

Figure 3.35: Location of fission track data and other evidence of exhumation. Numbers refer to location discussed in
the text.

3. Santander Massif. Shagam et al. (1984) also studied apatite and zircon fission tracks from 12
samples of this massif. Zircon ages range between 61 and 109 Ma and were interpreted to indicate
exhumation of the massif during the end of Cretaceous-Paleocene (with the older ages being mixed).
Apatite ages range between 18.9 and 3.8 Ma, indicative for exhumation during middle Miocene (19-14
Ma) in the central part of the massif (location 3b Fig. 3.35) and late Miocene to early Pliocene (7-4 Ma)
in the central (location 3b Fig. 3.35) and northern (location 3a Fig. 3.35) parts of the massif. Toro
(1990) reported an apatite fission track age of 30.85.8 Ma from a granite sample taken along the

120
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera Colombian Andes
Soat-Onzaga road in the southernmost Santander Massif (location 3c Fig. 3.35). Crowllie, 1989 (in
Toro, 1990), who performed the analysis, reported that the confined track lengths are shorted than
expected and interpreted the age as a mixed one. Forward modelling of the measured ages and lengths
according to the geological history suggest burial heating between 180 and 30 Ma followed by
exhumation and cooling between 30 Ma and present time (Crowllie, 1989 in Toro, 1990).
4. Floresta Massif. Toro (1990) reported an apatite fission track age of 22.34 Ma from the
Beln granite. Crowllie, 1989 (in Toro, 1990), who performed the analysis, reported that the confined
track lengths are shorted than expected and interpreted the age as a mixed one. Forward modelling of
the measured ages and lengths according to the geological history suggest burial heating between 180
and 30 Ma followed by exhumation and cooling between 30 Ma and present (Crowllie, 1989 in Toro,
1990).
5. Eastern flank of the Eastern Cordillera. According to Hossack et al. (1999), based on apatite
fission track data, the Chmeza thrust, which is the eastern margin of the EC, began its initial
exhumation at 25 Ma. However, the main deformation in the eastern foothills did not begin until 15
Ma, continued to 3 Ma, and may still be active at present. According to these authors the only obvious
synorogenic sediments that were deposited in the eastern foothills around growing thrust structures are
the youngest post 3 Ma Pliocene-Pleistocene deposits. Apatite fission track data for the more internal
parts are younger and define a general piggy-back development of the foothills from 15 Ma to 3 Ma
(Hossack et al., 1999). In the LLA foothills Rathke and Coral (1997) showed evidence of Oligocene
early development of the Cupiagua oil field structure. Corredor (1997) has reported similar evidence of
structural deformation in the eastern flank of the EC. Figures 2.23 and 3.14 show that incipient
inversion of Cretaceous normal faults occurred during Oligocene time. Confidential data from
Petroleum Companies suggest that since late Miocene (10-12 Ma) and during the Pliocene, massive
regional denudation occurred. Probably rock-uplift, exhumation and deformation started in the
marginal thrust fault systems and progressively moved toward the inner axial zone of the Cordillera.
6. Western flank of the Eastern Cordillera. Branquet et al. (1996,1999a,b) and Branquet (1999)
have shown that on the western flank of the Cordillera emerald deposits are genetically linked with tear
faults and associated thrusts developed during a compressive phase that occurred at the Eocene-
Oligocene time boundary (38-32 Ma). Such an age is based on 40Ar/ 39Ar and K/Ar dating of
syngenetic green muscovite crystallized on emerald-bearing vein wallrocks (Cheilletz et al., 1993,
1997). However, Romero et al. (2000) based on Rb/Sr analysis questioned the 38-32 Ma age and
proposed an age close to the Cretaceous-Paleocene boundary (67 Ma for the western emerald belt and
61 Ma for the eastern emerald belt) According to Cheilletz et al., (1993, 1997) emerald precipitation
occurred at temperatures of 300 C at a burial depth comprised between 4000 and 5000 m (Branquet, et
al., 1999a). Emeralds occur within Valanginian limestone and Hauterivian black shale (Branquet,
1999). If the stratigraphic thickness in this region between the Valanginian-Hauterivian boundary and
the top of the Cretaceous is approximately 7000 m, as reported by Rubiano (1989) then between 2000
and 3000 m of Cretaceous sediments would had been eroded during Eocene/Oligocene time, if
Cheilletz et al. age is accepted, or close to the Cretaceous-Paleocene boundary if Romeros age is
accepted.
In the Middle MV and western EC foothills AFTA, vitrinite reflectance data and eastward
thinning of Oligocene units (Mugrosa and Colorado Fms) constrain the age of the first uplift event of
the foothills region to the time between 34 to 30 Ma. At this time Cretaceous rocks began to cool from
a maximum temperature of 180 . Younger Upper Oligocene-Lower Miocene rocks preserve a
syntectonic stratal record of instantaneous limb rotation, associated with layer parallel shearing. Several
oil producing anticlines e.g. Provincia and Lisama belong to this generation of folds (Gomez et al.,
1999, Fig. 3.15). Miocene sediments were originally deposited across the foothills area of the present
EC as indicated by AFTA calculations of the eroded sedimentary column and balanced structural
sections (op cit). Diachroneity of westward breakthrough of the EC is indicated by the age of a
second thermal event. Cooling from maximum temperatures in the range of 110 to 120 C started
between 15 and 5 Ma in the southern part of the basin and between 5 to 0 Ma in the northern Middle
MV according to AFTA and vitrinite reflectance analyses (op cit). The main uplift of this mountain
range and complete conversion of the Middle MV into an intermontane basin occurred after 6.2 Ma.
(op cit).

121
Chapter 3
Based on outcrop, chronological data from volcanic ashes and palaeontology, fission track and
subsurface data Gomez et al. (1999) interpreted that during Eocene-Early Miocene communication
between the Middle MV and LLA basin to the east was partially interrupted by low discontinuous hills,
which resulted from incipient inversion of Mesozoic grabens in the area of the EC. Growth stratal
relations and time transgressive unconformities document this (op cit). In the Southern Middle MV
progressive forelimb rotation of west-verging anticlines, indicate that inclined shear was the likely
mechanism of Palaeogene deformation of the EC. Northward directed palaeocurrents attest to the
Eocene-Oligocene confinement of the MV between the Central Cordillera and EC. Upper Pliocene-
Lower Miocene strata finally overlapped the EC uplift as the sedimentation sourced by the Central
Cordillera overcame uplift rates of the eastern folds. Eocene-Lower Miocene basin fill also onlapped
over the Paleocene alluvial fans and the pediment surface that resulted from erosional retreat of the
Central Cordillera front (op cit). For the northern Middle MV Gomez et al. (1999) interpreted that
Paleocene and older rocks were highly deformed during Eocene as the central Cordillera front
propagated into the basin (Fig. 3.7). Net westward retreat of the Central Cordillera front has induced an
eastward-dipping unconformity, which is the base on the onlapping Eocene to Neogene sediments.
Chronology and kinematics of EC deformation events in the northern MV are not the same as in the
south. Palaeocurrent and provenance analyses indicate that central areas of the EC were uplifted during
sedimentation of the lower part of the Eocene La Paz Fm (op cit). An event of westward thrust
faulting of probable Eocene age has been identified in the central part of the Middle MV by Restrepo-
Pace et al. (1999a,b). Westward migration of EC deformation continued during Oligocene-Early
Miocene, as a series of fault propagation faults (Gomez et al., 1999).
7. Sabana de Bogot (axial zone of the Eastern Cordillera). Palynological and fission track
data on zircons obtained from ash layers indicate that the fluvial-lacustrine sediment record of the
Sabana de Bogot area registers major tectonic surface-uplift for the period between 5 and 3 Ma
(Andriessen et al., 1993). This will be discussed in detail in Chapter 4.
8. Central Cordillera. Jaramillo (1978, 1981) reported fission track ages of 10.51 Ma (Late
Miocene) on apatite and 586 and 62.43.6 Ma (Paleocene) on zircon, from the Manizales Pluton at
Fresno in this area. Based on fission track data and outcrop studies Gomez et al. (1999) proposed that
the Central Cordillera formed by transpressional strike-slip deformation beginning in the Campanian.
Fission track cooling ages of the Mariquita stock, a granitoid pluton in the eastern flank of the
mountain range, indicate slow cooling at rates of 2.8 C/Ma between the Campanian (77.67 Ma,
zircon) and the Oligocene (32.06.2 Ma, apatite), and 2.5 C/Ma since the Oligocene (op cit).
According to these authors during Paleocene eastward migrating deformation of the Central Cordillera
tilted the mountain front alluvial fans (Hoyn Fm).
In the northern part of the Central Cordillera Toro et al. (1999) reported fission track ages from
55 samples. 19 samples from the metamorphic and igneous (Antioquia Batholith) basement give zircon
ages between 75 and 35 Ma, while 16 samples correspond to Paleocene. These apparent ages indicate
that the basement cooled below 300 C during Paleocene. According to these authors an older age
obtained of 1855.5 Ma could indicate that the entire basement could not have been thermally affected
above 300 C during all the Cretaceous thermal events. Ages from volcanic zircons are comprised
between 0.350.05 Ma and 6.190.23 Ma with a peak between 1 and 4 Ma.
Maastrichtian unroofing of the Central Cordillera sourced quartzite and chert pebbles for the
Cimarrona Fm (Gmez and Pedraza, 1994). According to Campbell (1974) and Anderson (1970, 1972)
conglomeratic Cenozoic deposits in the MV recorded exhumation pulses of the Central Cordillera
during Eocene and Oligocene. Vaning exhumation of the Cordillera during early to middle Miocene
induced widespread muddy sedimentation (La Cira Colorado Fm) that recorded the beginning of
explosive volcanism in the magmatic arc of the Central Cordillera (Van Houten, 1976). Episodes of
active exhumation and volcanism of the Central Cordillera were recorded during 13.5 to 11.5 Ma
(Honda Gp), and approximately 1.4 Ma (Neiva Fm) in the Upper MV (La Venta area, Guerrero, 1993).
From the northern part of the Central Cordillera near Medelln, extensive remnants of at least
three uplifted planation surfaces have been recognized by Page and James (1981), the Pre-Central
Cordillera Erosion surface (Pre-S-1) over 3000 m, the Cordillera Central erosion surface (S-1) around
3000-2500 m, and the Ro Negro surface (S-II) around 2200 m. The tilted S-1 surface is found to
underlie the Miocene Honda Fm. By combining limited palaeomagnetic evidence on terrace deposits

122
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera Colombian Andes
and dissection depth of major rivers, Page and James (1981) calculated ages 22-18 Ma for the
development of the S-1 surface, which is in agreement with the 14-16 Ma age found for volcaniclastic
components in the Honda Fm (Setoguchi and Rosemberg, 1985). The Ro Negro surface is probably
older than 3 Ma. According to Soeters (1981a), the volcaniclastic Mesa Gp was dated palinologically in
the Pliocene (Dueas and Castro, 1981) and Thouret et al. (1985) obtained 4.30.4 and 3.50.4 Ma K-
Ar ages from it. The Ro Negro surface was truncated therefore around 4 Ma.
9. Garzn Massif. Van der Wiel (1991) and Van der Wiel and Andriessen (1991) studied
apatite, zircon and sphene fission tracks from 14 samples of the Precambrian metamorphic and Jurassic
igneous rocks of the massif. Samples are from San Vicente del Cagun (11a), Tres Esquinas Ventanas
road (11b) and Guadalupe-Florencia road (11c). Zircon ages range between 715 and 89 Ma and sphene
ages range between 183 and 146 Ma. Apatite ages range between 13.9 and 9.2 Ma. Using several
geological constraints and thermal fission track modelling to differentiate between mixed ages and
cooling ages she interpreted three exhumation and cooling events. The first exhumation event of
approximately 10.000m occurred 900 Ma ago probably related to orogenic processes. Between 850 and
210 Ma the Garzn Massif was denuded and eroded. The second differential 3500 m exhumation event
of faulted blocks occurred during the Cretaceous around 100 Ma. I discussed these ages in Chapter 2.
(Kohn et al., 1984 and Shagam et al., 1984 interpreted those ages as mixed ages). Finally between 12
Ma and present time the massif was exhumated by approximately 6500 m. After the first Neogene (12
Ma) exhumation pulse a second pulse occurred 6.4 Ma ago. The rate of very fast exhumation must have
exceeded the rate of thermal diffusion, producing thermal updoming of the isotherms, and the ages
should be interpreted as cooling following exhumation (Van der Wiel, 1991). Westward palaeocurrent
directions in the Villavieja Fm of the Upper MV indicate that the southern EC was already a mountain
range that completely closed the basin in its eastern side at about 11.8 Ma (ages from 40Ar/39Ar dates
and magnetostratigraphy, Guerrero, 1993). Changes of these palaeocurrents indicate that exhumation of
the southern EC began at 12.9 Ma (Guerrero, 1993).
10. Upper Magdalena Valley. Schwabe et al. (2001) reported zircon fission-track ages from the
Saldaa Fm. in this area. According to these authors Palaeozoic ages (235 to 395 Ma) represent
reworked zircon grains derived from Precambrian rocks of the Garzn Massif. Jurassic ages (between
136 and 176 Ma) are in agreement with the age of the volcanic rocks of the Saldaa Fm. Cretaceous
ages (100 4 Ma) are in agreement with the age of the Yav Fm. The presence of reworked Jurassic
zircons within the Neogene Honda Fm. indicates erosion of uplifted Jurassic rocks as a local source
area during the Neogene.
In conclusion the fission track ages combined with geological evidence suggest the following:
(1) Fission track ages and other geologic evidence indicates initiation of Andean exhumation of
blocks at the end of Cretaceous-Paleocene with local phases of exhumation following during
Eocene to Miocene time, terminating with regional exhumation in all blocks in the Pliocene-
Pleistocene. Shagam et al. (1984) interpreted this as the result of uncoupled crustal blocks being
uplifted in response to local stress until regional compression led to interlocking of the blocks
and their simultaneous rock-uplift during the Pliocene-Pleistocene time.
(2) Geological evidence suggests that Palaeogene exhumation of the Central Cordillera was regional,
supplying detritus to the eastern area. In the area of the Mesozoic extensional basins exhumation
was only local and probably of low elevation as suggested by local evidence of detrital source
areas.
(3) In the Perij and Mrida Andes fission track data suggest initial Oligocene exhumation of one of
the flanks, later Miocene exhumation of the other flank and finally exhumation of the whole area
including the axial central region during Pliocene time. Data suggest also a similar picture for the
EC. Deformation, and probably rock-uplift of the western flank started during Eocene/Oligocene
(38-32 Ma) and possibly Paleocene. Rock-uplift and deformation of the eastern flank started
during Oligocene and followed during Miocene (25 to 10 Ma) and regional rock and surface-
uplift including the central axial Sabana de Bogot area during the Pliocene (5-3 Ma).

123
Chapter 3
10.2. COMPARISON WITH MODEL RESULTS

Both model results and fission track data suggest a rock-uplift history starting at the end of
Cretaceous/Paleocene and following during Palaeogene time. However rock-uplift was moderate and
only affected tectonic blocks. Particularly model results and fission track data suggest initial moderate
rock-uplift in the flanks of the EC as expected in a scenario of inversion of the former extensional fault
systems delimiting the Mesozoic basin. As a first approach in a regional view, therefore, fission track
data seem to support model results. However, if a very moderate surface-uplift of 100 to 200-m
occurred in the flanks of the palaeo-EC as suggested by the models during Paleocene, it was not
recorded by fission track data. A possible explanation for this is that such a low topography was not
enough to produce enough exhumation and cooling. Because the lowest temperature associated with a
cooling age is 50-100 C (Faure, 1986), the amount of exhumation determined from cooling ages is
likely to be at least 1-3 km (England and Molnar, 1990). Fission track data only suggest Paleocene
exhumation of the Santander Massif and there is evidence that at that time the massif started to erode
and supply detrital sediments (Fabre, 1986, 1987) Deformation/uplift history during Oligocene and
early Miocene seems to agree with model results. However, more data are necessary to test model
results and define in detail the Palaeogene uplift history of the EC.

11. STRUCTURAL KINEMATIC MODEL OF PALAEOGENE DEFORMATION,


INCLUDING SEDIMENTATION AND EROSION EVENTS

In the flexural models we have recognized some local differences between the observed and
predicted subsidence. We interpreted these short wave length effects as produced by local relative
surface-uplift of fault limited blocks. Basin-wide scale flexural models can not predict them. As an
additional modelling test to try to understand these short wave effects of the Palaeogene subsidence
uplift pattern of the EC and to test if incipient inversion of the Mesozoic extensional basin is a viable
process consistent with available stratigraphic and structural data, I modelled the kinematic structural
deformation, sedimentation and erosion during Palaeogene along the regional cross-section of the EC
published by Cooper et al. (1995). This model incorporates structural geology to try to bridge basin
wide and sub-basin local scales.

I applied a 2D structural kinematic model of deformation, assuming inversion along the border faults of
the former extensional basin, including also sedimentation and erosion events. Sedimentation, erosion
and active compressional deformation processes are obviously interrelated. The purpose of such
modelling is to reconstruct step by step the structural, sedimentary and erosion evolution of the basin
during Palaeogene time. These models can be useful to infer the deformational history based on the
knowledge of the sedimentary record. In this model I accepted the hypothesis of early contractional
deformation of the former extensional basin during Palaeogene time, as suggested by local evidence
(Restrepo-Pace, 1999a,b; Gomez et al., 1999) and previous model results. I estimated the amount of
shortening necessary to generate local low topography as predicted by flexural models and also local
subsidence to create accommodation space for the observed Palaeogene sedimentary record.

11.1 STRUCTURAL, SEDIMENTARY AND EROSION MODEL

I used the program Thrustpack version 6.2 developed by Sassi et al. (1998) to model the
structural history of the EC during Palaeogene time. I applied the model to a regional balanced cross-
section. The program reconstructs through time the kinematic, structural, sedimentary and erosion
history along a cross-section. Such a reconstruction requires knowledge of the initial geometry of the
system before tectonic deformation occurred and the final deformed geometry (Sassi et al., 1998). The
initial and final geometries are usually taken from a previous balanced cross-section interpretation and
its restored state before deformation. The program simulates the kinematics of deformation,
sedimentation and erosion in a forward sense, starting from the initial undeformed section and ending
to the final, present day deformed section (Sassi et al., 1998). Because a good correspondence between
the two cross-sections can strongly depend on the fault geometry assigned in the present day deformed

124
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera Colombian Andes
cross-section, it is necessary to use a section where the fault geometry at depth has been interpreted. I
selected the regional balanced cross-section interpretation of Cooper et al. (1995) because it is the only
section that includes an interpretation of faults at depth until reaching a deep detachment in the upper
boundary of the lower crust.
The program Thrustpack version 6.2 allows the kinematic structural modelling of the geometry
of structures using both the fault-bend-fold deformation mechanism with kink fold geometry (Suppe,
1983) and the Foldis algorithm (Divies, 1997). In the Foldis algorithm the progressive folding of rock
layers is performed using a model that allows discrete slip between layers and strain concentration in
fold hinges. It simulates flexural slip mechanism that produces curved fold geometry and it can include
compaction of the lithologies due to an increase of burial depth trough time (Sassi et al., 1998). A
complete description of the algorithm Foldis can be found in Divies (1997). The sedimentation and
erosion simulation also is purely geometric and consists in the addition or erosion or material up to a
new earth surface defined by the user as a poly-line on the cross-section. The program also allows
simulation of thermal conductivity, organic matter maturity, generation and expulsion of hydrocarbons
from an oil source rock. However, in this study I only used the structural, sedimentation and erosion
module, which requires as input data the initial undeformed cross-section geometry, its dimensions, the
definition of each stratigraphic layer (including its properties), the initial topographic surface and the
fault trajectories geometry. All layers and faults are defined as poly-lines. Faults should be defined
starting from the most internal (hinterland) toward the most external (foreland). Each fault block or
thrust sheet is defined, together with certain amount of displacement per time increment defined by the
user, and applied to each fault block or thrust sheet starting with the autochthonous block. For each
time increment the user can apply displacement to each fault block as well as sedimentation and
erosion events. All displacements are referenced to the autochthonous block. Large-scale subsidence
and uplift such as flexural effects are introduced as a file indicating the rock-uplift or subsidence as a
function of the horizontal distance along the section. A complete description of the program can be
found in Sassi et al. (1998).
In order to apply the program to the regional balanced cross-section by Cooper et al. (1995) I
divided the section in two parts. I defined the faults and thrust sheets in each half of the section as
shown in Figure 3.36. I selected similar time intervals as those applied in the flexural models. For each
fault block (thrust sheet) and for each time interval, I applied several combinations of horizontal
contraction displacements until I found acceptable displacements that can be compared to the observed
subsidence or modelled palaeotopography. The amount of shortening in kilometers applied to each
thrust sheet is shown in Table 3.3. I tried to generate an amount of topography similar to that suggested
by flexural modelling and an amount of subsidence similar to that of the observed Palaeogene
sedimentary record.

11.2. RESULTS

After applying several combinations of contractional displacements to each thrust sheet for
each time interval to best fit the observed subsidence and the calculated topography from flexural
models, I arrived to the results shown in Figure 3.37. In this Figure the two halves of the section have
been assembled in a single section, some minor mismatch between the two parts have been artificially
eliminated in the drawing. This mismatch was produced by dividing artificially the original section in
two parts and applying eastward displacement to the eastern part relative to autochthonous block in the
eastern LLA, and westward displacement to the western part, relative to a western autochthonous in the
western Middle MV (Figure 3.37). In nature the less deformed autochthonous block is the eastern LLA,
which is part of the South American craton, and all the Palaeogene deformation occurred by relative
displacement toward this more stable block. The western Middle MV probably was the most deformed
block during Palaeogene, as evidenced in subsurface (Fig. 3.15), contrary to this artificial assumption.
However, the division of the section in two parts only was a necessary artifact to handle the section
with the program. A simple assumption was applied reducing the number of active faults and their
geometry, as shown in Figure 3.36.

125
W E

Middle Magdalena Valley Mesozoic extensional basin Llanos

9
4 1
10 8
7 6
5 3 2

11 (Relative autochthonous block) 0 (Relative autochthonous block)

126
Westward compressional Eastward compressional displacement
displacement of blocks in of blocks in the eastern half of the section
Chapter 3

the western half of the section

0 100 km

Figure 3.36: Simplified assumptions applied for Thrustpack modelling to the regional cross section of the Eastern Cordillera restored
to its undeformed state at the end of Cretaceous (from Cooper et al., 1995). The undeformed section has been divided into two parts
that were modelled independently. Some faults were eliminated. A displacement of each thrust sheet or tectonic block has been
applied relative to the autochthonous block during modelling.
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera Colombian Andes

Western Part Eastern Part Total


Thrust Sheet
11 10 9 8 7 Sh 6 5 4 3 2 1 0 Sh Sh

End Cretaceous 65
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
Ma

End Paleocene
54.8 Ma 0 0 -3 -3 -3 -3 1 1 1 1 0 0 0 4 4

End Eocene 33.7


Ma 0 0 -3.8 -3.8 -3.8 -3.8 2 2 2 2 0 0 0 2 6

End Early
0 0 -4 -6 -6.3 -6.3 3 3 3 3 0 0 0 3 9.3
Miocene 16.4 Ma
Sh: shortening

Table 3.3. Amount of shortening in kilometres applied to each thrust sheet in the Thrustpack program. Positive
numbers indicate left to right displacement, negative values indicate opposite sense of displacement.

Particularly the deep detachment interpreted by Cooper et al. (1995) below the MV was not considered
in the model. Following Cooper et al. (1995) interpretation I applied displacement only to those faults
interpreted by these authors as active during compression tectonics. For example I applied
displacement to the Chmeza Fault considered to be active during Tertiary compression, but I did not
applied displacement to the Guaicramo Fault interpreted as a Mesozoic extensional fault passively
transported with a short-cut basement block during Palaeogene compression deformation. A
consequence of this interpretation and the geometry of faults at depth is the absence of a clear inversion
of the faults defining the eastern border of the former extensional basin, contrary to the results
suggested by flexural models, which are more in agreement with the structural cross-section
interpretation by Colletta et al. (1990) who interpreted inversion of the Guaicramo Fault. The results
show that more clear inversion structures were developed in the western border of the former
extensional basin. Although some Mesozoic normal faults were passively transported with short-cut
basement blocks (e.g. Las Esmeraldas Fault, ESRI and Ecopetrol, 1994), flexural models support their
inversion (as interpreted by Colletta et al., 1990), which is more probable in a scenario of transpression.
Normal faults that usually develop at dip angles close to 60 are not inverted in frontal compression,
but they are inverted in oblique compression. Because of the very reduced thickness of the Palaeogene
deposits compared to the deep detachment interpreted for the section it is difficult to reproduce with
accuracy the geometry of these Palaeogene deposits, which at least partially have been exaggerated in
the model. Similarly the topography in the model has been at least partially exaggerated compared with
that suggested by flexural models, even when applying considerable amounts of erosion. An important
result of the model is the relatively small amount of total shortening (up to 9.3 km at the end of early
Miocene) which has produced local topography and subsidence greater than those observed or
estimated through flexural models. Therefore, assuming the dip angle of the faults is as interpreted by
Cooper et al. (1995), the estimate of 9.3 km of shortening at the end of early Miocene is an upper limit.
Obviously the amount of shortening necessary to generate a fixed amount of rock-uplift/subsidence
increases if the dip angle of the faults decreases.

127
Chapter 3

End of Cretaceous (65 Ma)

End of Paleocene (54.8 Ma)

4 km

End of Eocene (33.7 Ma)

6 km

End of Early Miocene (16.4 Ma)

9.3 km

La Salin a Fault B o y a c a F P e s c a F Chameza F Guaicaramo F


Present (0 Ma) Cusiana F

70 km

100 km
Figure 3.37: Kinematic structural models obtained with the Thrustpack program, simulating the evolution of the
Eastern Cordillera during Palaeogene time. The present day section is from Cooper et al. (1995) and has not been
modelled.

Fault dip angles close to 30, as interpreted by Cooper et al. (1995) are reasonable for
compression faults. Palaeogene shortening could be even less taking an average of dip values of 60 for
former normal faults. An additional limitation of the Cooper et al. (1995) interpretation for the deep
geometry of the faults is that the lower crustal detachment interpreted by these authors is deeper that
the crustal thickness suggested by stretching models applied to the Mesozoic history of the area.
In conclusion, it appears that all the limitations described for the model make it insufficiently
accurate to predict details of the structure, palaeotopography or the stratigraphic relationships between
Palaeogene stratigraphic units and related unconformities. However, an important result of the model is
that the small amount of compressional shortening that occurred during Palaeogene was enough to
generate the amount of subsidence/rock-uplift suggested by the Palaeogene sedimentary record and the
inferred flexural-modelled palaeo topography.

12. DISCUSION

12.1. CONSTRAINING DIFFERENT TECTONIC SCENARIOS FOR THE PALAEOGENE

To constrain the possible alternative tectonic scenarios on Palaeogene tectonic history, I used
quantitative thermal subsidence and flexural models to test how plausible these different tectonic
scenarios are. Although the observed subsidence could not be explained by any individual hypothesis, a

128
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera Colombian Andes

combination of all effects gave a satisfactory result. The observed local flexural component of
subsidence is interpreted as produced by uplift of local faulted blocks (Fig. 3.1). Van der Hammen
(1961) was the first geologist who recognized Palaeogene tectonic activity in the area of the EC. This
interpretation is supported by sedimentological (Anderson, 1970; 1972; Butler and Schammel, 1988;
Hoorn, 1988; Laverde, 1989; Alfonso, 1989; Ecopetrol-ICP, 1996; Gmez et al., 1999), palynological
(Van der Hammen, 1961; Sarmiento, 1992, 1993), petrographical (Porta, 1965; Anderson, 1970, 1972),
mineralogical (Cheilletz et al., 1993, 1997; Branquet et al., 1996), structural (Morales et al., 1956;
Julivert, 1970; Butler and Schammel, 1988; Guillande, 1988; Branquet et al., 1996, 1999b; Casero et
al., 1995, 1997; Rathke and Coral, 1997; Corredor, 1997; Restrepo-Pace, 1999a,b; George et al., 1997;
Branquet, 1999; Gmez et al., 1999) and fission track (Shagam et al., 1984; Toro, 1990; Van der Wiel,
1991; Van der Wiel and Andriessen, 1991; Hossack et al., 1999; Gmez et al., 1999) evidence for pre-
Andean Palaeogene compression deformation or slight local rock-uplift of the EC and MV. Incipient
inversion of Mesozoic extensional basins (Fig. 3.1, e.g. Gomez et al., 1999) was probably the
mechanism responsible for uplift of local faulted blocks.

12.2. PALAEOGENE TECTONIC HISTORY AND PLATE-TECTONICS

Inverted extensional basins, upthrust basement blocks and whole lithospheric folds (Cloetingh
et al., 1999) are common intraplate compressional/ transpressional structures that affect continental
plate margins and/or plate interiors (Ziegler et al., 1998). Compressional features at plate margins are
commonly associated with collision-related orogenic processes. Compressional features at plate
interiors are usually not associated with orogeny. (Ziegler et al., 1995). In Colombia all plate-tectonic
interpretations (e.g. Pindell and Erikson, 1993; Pindell and Tabut, 1995) propose collision of the
Caribbean with northwestern South America. Collision was oblique and diachronous, becoming
younger northward (Pindell and Erikson, 1993; Pindell and Tabut, 1995). Therefore plate-tectonic
history suggests that Palaeogene basin inversion and upthrust of basement blocks (Fig. 3.1) were
collision-related, probably involved right-lateral transpressional deformation, and led to pre-Andean
orogeny in the Central Cordillera during the Palaeogene. Transpressionally deformed grabens are also
associated with zones of major wrench faulting (Ziegler et al., 1995). Probably some right-lateral
strike-slip faults (e.g. Palestina Fault, Irving, 1971) were active during the Palaeogene.
Compressional/transpressional stresses related to collisional plate interaction are responsible for
reactivation of pre-existing crustal discontinuities (Fig. 3.1), upthrust of basement blocks and inversion
of tensional hanging-wall basins (Ziegler et al., 1995). According to Ziegler et al. (1998) these
compressional structures can occur at distances up to 1600 km from the collision front, both in the
forearc (foreland) and backarc (hinterland) positions with respect to the subduction system controlling
the evolution of the corresponding orogen.
Compression in the region behind a magmatic arc is associated with Andean-type orogens and
occurs during periods of increased convergence rates between the subducting and overriding plates
(Ziegler et al., 1998). Inversion of rift tensional hanging-wall basins located behind a magmatic arc is
the result of acceleration of convergence rates between the colliding plates, their increased mechanical
coupling and the transmission of compressional stresses into the backarc domain of the overriding plate
(Uyeda and McCabe, 1983; Ziegler, 1993 in Ziegler et al., 1998). Rates of the Caribbean-South
America plate convergence changed during the Cenozoic. Periods of development of compressional
structures seem to correlate with times of high convergence rate, particularly during the Eocene (Daly,
1989) and during the Late Miocene-Pliocene, i.e. Andean orogeny (Cooper et al., 1995). Probably
during late Oligocene-Early Miocene, development of compressional/transpressional structures was
associated with the rupture of the Farallon Plate into the Cocos and Nazca plates about 25 Ma ago
(Wortel and Cloetingh, 1981; Duncan and Hardgraves, 1984). Ziegler et al. (1995) cited several
examples of intraplate compressional/transpressional structures in Europe developed during phases of
plate boundary reorganisations that ultimately lead to the break-up of plate assembly.
Build up of intraplate compressional/transpressional stresses in the NW margin of South
America during latest Cretaceous-Palaeogene could have been favoured by subduction impediment

129
Chapter 3
caused by the arrival of more buoyant oceanic crust, such as an oceanic plateau. Nivia (1987), Kerr et
al. (1996, 1997) and Sinton et al. (1998) based on geochemical evidence have proposed that the
accreted oceanic terranes of western Colombia (Amime Terrane in the werstern flank of the Central
Cordillera, Calima Terrane in the Western Cordillera and Cuna Terrane in the Serrana de Baud, Fig.
1.1) are similar to the anomalously thick and buoyant Caribbean Plate. According to these authors the
Caribbean Plate west of Colombia was young lithosphere and an oceanic plateau. Subduction processes
can be impeded if buoyant material, such as a spreading ridge, oceanic plateau or micro-continent,
collides with a mature arc-trench system (Muellert and Philips, 1991; Cloos, 1993, all in Ziegler et al.,
1998). Large obstacles as oceanic plateaus can provide sufficient subduction resistance to deform the
arc-trench, causing the build up of compressional stresses in the subducting plate and potentially its
imbrication (Ziegler et al., 1998). This can explain intense deformation in the accreted oceanic plateau
terranes of the Western Cordillera (Nivia, 1987). Build up of compressional/transpressional stress
probably also favored deformation in the continental plate margin represented by the Central
Cordillera. Ziegler et al. (1998) suggested that at any stage in a subduction zone the upper plate
continental margin is weaker than the oceanic lithosphere plate margin. This suggests that the upper
plate margin is the most likely candidate to be compressionally reactivated. If a magmatic arc was
present in the Central Cordillera, as suggested by Palaeogene magmatic evidence, the Central
Cordillera lithosphere was hot, weak and prone to deformation. According to Ziegler et al. (1998) for
the build up of intraplate compressional stresses in forearc and foreland domains, the following
collision-related scenarios are envisaged:

(1) During the initiation of a subduction zone along a passive margin or within an oceanic basin.
(2) During subduction impediment caused by the arrival of more buoyant crust, such as an oceanic
plateau or a micro-continent at a subduction zone.
This is applicable to Colombia during Late Cretaceous-Palaeogene time.

12.3. MECHANICAL ASPECTS OF BASIN INVERSION

Mechanical aspects of basin inversion depend on the interplay of stresses and rheology of the
lithosphere.

12.3.1. Stresses

We have discussed in terms of plate-tectonics the build up of collision-related intraplate


compressional stresses at the NW South American plate margin. According to Ziegler et al. (1998), this
build up is indicative for mechanical coupling between an orogenic wedge and its fore- and/or
hinterland (Ziegler et al., 1998). If this is applicable to the study area it would imply some mechanical
coupling between the orogenic wedge, represented mainly the Central Cordillera and the regions east of
it (MV, EC, and LLA) during Palaeogene time. However, the intensity of collisional coupling between
an orogen and its fore- and hinterland is temporally and spatially variable. This can be a function of
oblique collision (Ziegler et al., 1995, 1998).

12.3.2. Rheology of the lithosphere

Localization of collision-related compressional intraplate deformation is controlled by spatial


and temporal strength variations in the lithosphere, in which the thermal regime, the crustal thickness,
the pattern of pre-existing crustal and mantle discontinuities, as well as sedimentary loads and their
thermal blanketing effect play an important role (Ziegler et al., 1998).
The strength of the continental lithosphere is controlled by its depth-dependent rheological
structure in which the thickness and mineralogical composition of the crust, and the thickness of the
mantle lithosphere. The latter depends largely on the potential temperature of the asthenosphere, as
well as the presence of fluids and strain rates, play a dominant role (Stephenson and Cloetingh, 1991;
Ranalli, 1995; Ziegler et al., 1995; Cloetingh and Burov, 1996).

130
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera Colombian Andes
Important inferences about lithosphere rheology in the EC during Palaeogene can be inferred
from flexural modelling results. The parameter that characterizes the apparent strength of the
lithosphere is the flexural rigidity, which is commonly expressed through the effective elastic thickness
of the lithosphere. For an elastic plate the notion of the integrated strength is quite close to that of the
flexural rigidity. Burov and Diament (1995) have shown that the effective elastic thickness of the
continental lithosphere is dependent on:

(1) The thermal state/age of the lithosphere (thermal age defined as a period of time required for the
lithosphere to reach its present-day thermal state, assuming that the lithosphere was initially
melted). The thermal age controls the depth to a specific geotherm obtained from a plate cooling
model, assuming that the lithosphere did not undergo thermal re-setting during this time. The
thermal age gives the age of the last large-scale thermal event (Burov and Diament, 1995). The
thermal state of the lithosphere controls lithospheric strength, since temperature dependent creep
controls the ductile strength of the lower crust and lower mantle lithosphere. Flexural models in
the EC infer a weak lithosphere (EET 5 km in the area of the former Mesozoic extensional
basin). In this area during Paleocene the thermal age of lithosphere was very young since the last
stretching thermal event was Aptian (see Chapter 2) Weak lithosphere corresponds to thermally
destabilized lithosphere. Young rifts and volcanic areas are characterized by low elastic thickness
values (e.g. Ebinger et al., 1989, in Burov and Diament, 1995) due to thermal weakening and
necking of the lithosphere. Stretched lithosphere is weak and low elastic thickness values in the
EC may have inherited this weakness. This could explain the very low effective elastic thickness
( 5 km) in the area of the EC during Palaeogene time. However, the thermal anomaly associated
with an active rift decays rapidly upon termination of crustal extension; after 60 Ma about 65 %,
and after 180 Ma about 95 % of the anomaly had decayed (Ziegler et al., 1995).
(2) The coupling or decoupling state of the crust and mantle.
(3) The thickness and proportions of the mechanically competent crust and mantle. The thickness of
the mechanically competent crust and the degree of coupling or decoupling are generally
controlled by composition of the upper and lower crust, the total thickness of the crust, and by
the crustal geotherm. If decoupling take place, as is of common occurrence in continental
lithosphere, it permits as much as a 50 % decrease of elastic thickness, compared with elastic
thicknesses implied for conventional thermal profiles (Burov and Diament, 1995). Low values of
EET are usually associated with fault-controlled upper crustal flexure patterns (Van Wees and
Cloetingh, 1994). In the EC area during Palaeogene the reduced EET values (<10 km) inferred
from flexural models suggest that lithosphere strength was represented only by the upper crust
strength (c.f. Cloetingh and Burov, 1996). This compressional intraplate deformation probably
affected only the upper crust. Probably Palaeogene transpressional deformations in the area of
the EC may have involved simple shear type detachment of the crust at the level of the
rheologically weak lower crust from the mantle lithosphere, as suggested by Ziegler et al. (1995)
in many other areas.
(4) The local curvature of the plate, which is directly related to the bending stress (Burov and
Diament, 1995). The curvature of the plate depends also on the rheological structure and on the
distribution of external loads applied to it (e.g. topography, sediment fill and plate boundary
forces). Bending stresses created by major mountain belts are large enough to cause inelastic
deformation (brittle failure and ductile flow) in the underlying plate, which in turn, leads to a 30
to 80% decrease of elastic thickness beneath such belts (and less beneath the adjacent regions).
The boundary forces and moments may lead to more localized but even stronger reductions in
elastic thickness (Burov and Diament, 1995). Flexural models for the EC during Palaeogene infer
small bending stress and small plate curvature. Thus this effect probably was not important.

Additionally the great thickness of sedimentary fill in the area of the EC, could have produced a
blanketing thermal effect that inhibited lithosphere cooling, so preserving a relatively hot and weak
lithosphere. Ziegler et al. (1998) have shown that the integrated strength of the lithosphere at a
continental margin is markedly reduced if there is a thick syn- and post-rift sedimentary prism.

131
Chapter 3
12.3.3. Pre-existing crustal discontinuities

An additional important effect that controls lithosphere strength is the presence of pre-existing
crustal discontinuities that weaken the lithosphere and play a crucial role in localising intra-plate
compressional deformations. Standard rheological models that do not include pre-existing crustal
discontinuities are unable to explain inversion of rift basins lacking a thick post-rift sedimentary prism
(Van Wees, 1994; Ziegler et al., 1998). Deep reaching pre-existing crustal discontinuities, such as
lithological inhomogeneities associated with ancient suture zones, as well as extensional, thrust and
wrench faults, cause significant weakening of the upper crust. Such discontinuities are apparently
characterized by a reduced frictional angle, particularly in the presence of fluids (Van Wees, 1994;
Sibson, 1995; Ziegler et al., 1995). A major factor in the compressional reactivation of rifted basins is
the failure of the upper crust along pre-existing faults causing stress concentration on the mechanically
strong part of the upper mantle, resulting ultimately in its yielding. Reactivation of relatively steep
dipping normal faults occurs when the angle between their strike and the compressional stress
trajectory is smaller than 45 (Allemand and Brun, 1991; Ziegler et al., 1995). In the EC results of
flexural models suggest that the former Mesozoic extensional hanging-wall basins were slightly
inverted during Palaeogene time (Fig. 3.1). This corroborates the importance of pre-existing crustal
discontinuities to focus deformation. The reactivation of faults also suggests an oblique orientation of
the compressional stress trajectories, which is in agreement with oblique plate convergence.
In conclusion, incipient inversion of Mesozoic extensional basins and upthrusted basement
blocks in the EC area was favored by collisional right-lateral transpresssional stresses that affected a
weak lithosphere and reactivated previously existing discontinuities, i.e. the inversion of former
Mesozoic extensional faults (Fig. 3.1).
The earlier deformation in the western MV (e.g. George et al., 1997; Restrepo-Pace et al.,
1999a,b) compared to the EC or LLA foothills suggests that basin inversion prograded eastwards.
This requires that basins located more proximal to the collision front were characterized by a
mechanically-weaker pre-inversion lithosphere than more distal basins. It appears logical that the
weakest basins are prone to early inversion. (Ziegler et al., 1995).
The lack of important compressional intra-plate deformations in the LLA can be related either
to the absence of major crustal discontinuities in the respective foreland, or to the lack of mechanical
coupling between the LLA foreland and the orogen during the evolution of the latter (Ziegler et al.,
1995).
The Palaeogene stratigraphic record of the EC and neighbouring MV and LLA basins is useful
to date collision-related intraplate compressional deformations affecting the northwestern margin
margin of South America (c.f. Ziegler et al., 1995).

12.4. LARGE-SCALE BASIN WIDE FLEXURE AND STRUCTURAL GEOLOGY

For forward stratigraphic modelling of piggyback basins the importance of coupling large-scale
basin wide flexure with thrusting on sub-basin scale and sedimentation has been demonstrated by
Zoetemeijer et al. (1993). I have incorporated structural geology in basin modelling in order to bridge
the basin-wide and sub-basin scales (Cloetingh et al., 1994) applying the Thrustpack model. The results
are consistent with the interpretation of incipient inversion of the hanging-walls of major Mesozoic
normal faults during Palaeogene time (Fig. 3.1). The results of the modelling approach as well as the
increasing evidence indicates that an important compression/transpression event took place related to
inversion of the Mesozoic extensional basin in the area of the EC (Fig. 3.1). Also the results of this
study suggest that the Palaeogene deformation history of the EC is related to inversion processes
mainly of the faults limiting the original Mesozoic extensional basin (Fig. 3.1). If this interpretation is
true it implies a compressional stress-field oblique to the fault strike. This is consistent with oblique
plate convergence with a right-lateral strike-slip component. The total amount of shortening was
moderate (<10 km). Deformation started close to the Central Cordillera and prograded eastward.

132
Palaeogene Incipient Basin Inversion History of the Eastern Cordillera Colombian Andes
12.5. SUGGESTIONS FOR FUTURE STUDIES

Although there is an increasing amount of evidence of earlier pre-Andean deformation in the


EC, and there is an increasing interest in the study of the Tertiary sedimentary record of the EC, many
uncertainties remain. Main uncertainties are related to the partial erosion of the sedimentary record and
to age determinations and stratigraphic correlation. All these uncertainties make any attempt to use the
available data difficult. In the case of this study I tried to compile the best possible database.
Nevertheless some assumptions were necessarily applied to generate the palaeogeographic and
thickness maps that constitute the conceptual model on which the modelling approach is based. An
important constraint on the thermal, uplift and erosion history of any orogen is fission track data.
However, the number of analyses available for the EC is very small and track lengths in most
cases have not been neassured. Clearly they are inadequate. Understanding of the tectonic and
sedimentary history of the EC needs more and better stratigraphic and fission track data. Palaeogene
compressional deformation has been documented locally (Resterpo-Pace et al., 1998, 1999a,b) also for
the emerald mineralizations of the EC (Branquet, 1999). However, clear and detailed identification of
such Palaeogene deformational history and its structural style is difficult because it has been partially or
totally masked by the later and stronger Andean deformation. More and better surface and subsurface
data are also needed to understand Palaeogene deformation. Large volumes of Cretaceous organic-rich
shales (Mayorga and Vargas, 1995; Mora et al., 1996, 1997; Mora 1997) were mature enough to
generate and to expulse large volumes of hydrocarbons during Palaeogene time (Mora et al., 1996,
1997; Mora, 1997). The Palaeogene deformation history could have created structural traps for such
hydrocarbon charge. Understanding of such Palaeogene deformation history, its structural styles and
the preservation and modification of Palaeogene structures during Andean deformation could be a key
element for further oil exploration.

13. CONCLUSIONS

Flexural models and numerous evidence, such as the presence of Palaeogene unconformities,
e.g. the regional Eocene unconformity (which locally truncates structures and other local
unconformities), lateral changes of facies and thickness, local erosion indicated by detrital composition
of sandstone and limited fission track data suggest that an incipient inversion of Mesozoic extensional
basins occurred during Palaeogene time.
Palaeogene basin inversion was related to the collision of oceanic plateau terranes with the
northwestern margin of South America. Build up of intraplate compressional/transpressional stresses in
the NW margin of South America during latest Cretaceous-Palaeogene could have been favoured by
subduction impediment caused by the arrival of more buoyant oceanic crust, such as an oceanic plateau
(Nivia, 1987; Kerr et al., 1996, 1997; Sinton et al., 1998). Right-lateral transpressional deformation
likely lead to a pre-Andean orogeny in the Central Cordillera during the Palaeogene. Probably some
right-lateral strike-slip faults (e.g. Palestina Fault, Feininger, 1970; Irving, 1971) were active during
Palaeogene. Periods of development of basin-inversion and compressional structures seem to correlate
with times of high convergence rate, particularly during Eocene (Daly, 1989) and Late Miocene-
Pliocene, i.e. during Andean orogeny (Cooper et al., 1995). During late Oligocene-Early Miocene,
development of compressional/transpressional structures was associated with the rupture of the
Farallon Plate into the Cocos and Nazca Plates about 25 Ma ago (Wortel and Cloetingh, 1981; Duncan
and Hardgraves, 1984). Development of compressional/transpressional structures suggests some
mechanical coupling between the orogenic wedge, represented mainly the Central Cordillera, and the
regions east of it (MV, EC, and LLA) during Palaeogene time. However, the intensity of collisional
coupling between the orogen and its fore- and hinterland is temporally and spatially variable,
depending on the obliquity of the collision (Ziegler et al., 1995, 1998).
Subsidence in the pre-Andean LLA, palaeo-EC and MV during Palaeogene time cannot be
explained only by the topographic load of the palaeo-Central Cordillera as if the basin were a simple
foreland basin (as interpreted in literature e.g. Cooper et al., 1995). Palaeogene subsidence only could
be modelled assuming three different subsidence components: (1) Residual thermal subsidence after
Mesozoic rifting; (2) Flexural subsidence of the lithosphere due to topographic load of the Central

133
Chapter 3
Cordillera; and (3) Flexural subsidence of the lithosphere produced by incipient topography generated
during the Palaeogene. Flexural subsidence models assuming the lithosphere behaves as an elastic plate
of laterally variable thickness, and including these three subsidence components along four regional
cross-sections, suggests that during the Palaeogene local topography (up to 500 m) was developed close
to the borders of the former Mesozoic extensional basin, probably by inversion along the border
extensional faults. However, such topography was probably discontinuous and low. It did not
significantly disturb the sedimentary and palaeocurrent pattern in the Palaeogene basin. Modelling
results require very low values of effective elastic thickness in the area of the former extensional basin.
The Mesozoic rifting events reduced significantly the strength of the lithosphere, making it very prone
to Palaeogene deformation and to further Andean deformation. Low values of effective elastic
thickness are usually associated with fault-controlled upper crustal flexure patterns (Van Wees and
Cloetingh, 1994). In the EC area during Palaeogene the reduced EET values (<10 km) inferred from
flexural models suggest that lithosphere strength was represented only by the upper crust strength (c.f.
Cloetingh and Burov, 1996). This compressional intraplate deformation was restricted to crustal levels
involving simple shear-type detachment of the crust at the level of the rheologically weak lower crust
from the mantle lithosphere. The great thickness of sedimentary fill in the area of the EC could have
produced a blanketing thermal effect that inhibited lithosphere cooling, thus preserving a relative hot
and weak lithosphere during the Palaeogene. Incipient inversion of Mesozoic extensional basins
occurred by reactivation of previously existing discontinuities, such as Mesozoic extensional faults.
The occurrence of earlier deformation in western the MV (e.g. George et al., 1997; Restrepo-Pace et
al., 1999a,b) as compared to the EC or LLA foothills suggests basin inversion progradation. This
requires that basins located more proximal to the collision front were characterized by a mechanically
weaker pre-inversion lithosphere than that of more distal basins.
Structural kinematic modelling along a regional cross-section suggests that the amount of
shortening during Palaeogene necessary to produce that Palaeogene topography is small and dependent
of the dip angle of the Palaeogene contraction faults. Assuming a dip angle close to 30 and using the
structural cross-section interpretation by Cooper et al. (1995) the modelled total amount of shortening
at the end of Early Miocene appear to has been less than 10 km. A direct consequence of this
interpretation is the possibility of generation of hydrocarbon traps during Palaeogene, a time when
petroleum generation and migration occurred according to petroleum system modelling results
published in the literature. Oil generation in Cretaceous rocks in portions of the EC happened a long
time before formation of late Oligocene-Miocene anticlines, the oldest Tertiary traps preserved along
the foothills (Gomez et al., 1999).

134
CHAPTER 4

NEOGENE BASIN INVERSION HISTORY OF THE EASTERN CORDILLERA,


COLOMBIAN ANDES

1. INTRODUCTION

Inverted extensional basins and upthrust basement blocks are common intraplate
compressional/transpressional structures that affect continental plate margins (Ziegler et al., 1998).
Compressional /transpressional stresses that are related to collisional plate interaction are responsible
for inversion of tensional hanging-wall basins (Ziegler et al., 1995). Compression in the region behind
a magmatic arc is associated with Andean-type orogens during periods of increasing convergence rates
between the subducting and overriding plates (Ziegler et al., 1998). During basin inversion, basin-
controlling faults reverse their movement due to compressional/transpressional stresses, and basins
become positive features.
The Neogene tectonic history of Colombia and particularly the history of the EC (Fig. 1.1) are
an excellent example of complete inversion of a former extensional basin and formation of a mountain
range related to the Andean Orogeny. According to the current literature (e.g. Cooper et al., 1995) the
former Mesozoic extensional basin was inverted during Neogene time to form the EC, while the LLA
and Middle MV became independent foreland basins related to flexural loading of the EC (Figs. 1.1
and 1.2). Up till now no flexural modelling has been carried out to test this hypothesis. A record of the
tectonic rock-uplift and exhumation history of the EC has been preserved in the synorogenic
sedimentary record of the neighbouring MV and LLA foreland basins. The study of the sedimentary fill
of these basins gives important information about basin formation mechanisms (e.g. Cloetingh et al.,
1993) and about the tectonic rock-uplift of the EC during Neogene time. The aim of this Chapter is to
contribute to understanding of the tectonic basin forming and inversion mechanisms during Neogene in
terms of geodynamic processes that govern deformation of the lithosphere. This Chapter studies the
tectonic subsidence of the MV and LLA Neogene basins. The hypothesis that tectonic subsidence of
these basins was the result of lithosphere flexure produced by the increasing topographic load
represented by the uplifting EC during Neogene time is successfully tested through modelling. The
contribution of thermal subsidence after Mesozoic stretching is also tested for Neogene time. The
surface-uplift evolution of the EC is delineated from flexural models (see location on Fig. 4.1) and its
exhumation is evidenced by scarce fission track data from literature (Kohn et al., 1984; Shagam et al.,
1984; Van der Wiel, 1991; Andriessen et al., 1993; Hossack et al., 1999). Finally the evolution of the
lithosphere strength through Neogene time is delineated and discussed.

2. TECTONIC SETTING

2.1. PLATE TECTONIC INTERPRETATIONS

2.1.1. Middle Miocene

Collision and accretion of the Choco Block (Serrania de Baudo in Fig. 1.1, accreted oceanic
Cuna Terrane according to terminology of Toussaint, 1995a,b) with the north-western margin of South
America occurred during the Middle Miocene (Duque-Caro, 1990) and may have contributed to
loading and have initiated deformation in the EC. An Andean foreland basin was developed since
Middle Miocene in the area of the LLA (Cooper et al., 1995).

2.1.2. Late MiocenePliocene

The South American Plate rate convergence was fast, inducing deformation in the Colombian
Andes (e.g. Cooper et al., 1995). Major deformation of the EC began at approximately 10.5 Ma
(Cooper et al., 1995). During this deformation phase the EC was uplifted and eroded (Fig. 1.2). Pre-
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes

Serrania Maracaibo

Lucas
de Perija Basin

r y)
1400

s
de
nda
Cucuta

de San

An
Barinas

bou

a
Basin

id
52

er
Serrania
te -

M
42 26
epla
23
+
53 - - 0 100 200 Km
1300 70 71 +
(F r e

43 + - Bucaramanga
+ 74 18
+ +
72 73 25 Arauca 1
+ - +
44
27 47 +
+ +
7
1200 Medellin -
Llanos Orientales
ra

Basin (LLA)
d i ll e
Co r

1100 Tunja Yopal

Manizales 3

Fi
e

1000 Bogota g.
u t ur

Ibague 4.
6
l S

4
Villavicencio
er a

Fig. 4.7
Ro m

900
5
al
nt r
Ce

Neiva 6
800
La Macarena
Serrania de

San Jose del 8


Guaviare
13 10 11 12 7

800 900 1000 1100 1200 1300 1400 1500

Figure 4.1: Location of stratigraphic columns and wells used (see Table 2.1) and location of lithosphere flexure 2D
model sections. Numbers along sections refer to labelling of stratigraphic transversal sections (Figs. 2.7, 2.8 and 2.9)
and also to labelling of lithosphere flexure 2D models (Figs. 4.10 to 4.12 and 4.14). The Romeral suture is the
westernmost boundary of continental crust. It was used as a free plate boundary for the 2D flexural models.

existing extensional faults were inverted and new compressional structures developed. On the western
flank of the EC and in the MV, middle Eocene folds were reactivated (Buttler and Shamel, 1988).
Erosional deposits from the EC are preserved in the Guayabo FM in the LLA. Deformation and
rock-uplift are still active, periodically causing earthquakes. The Neogene sedimentary record of the
Sabana de Bogot (Fig. 2.1) suggests that 1000 to 2000 m of surface-uplift occurred between 5 and 3
Ma (Van der Hammen et al., 1973; Van der Hammen and Hooghiemstra, 1997; Hooghiemstra, 1984,
1989; Helmens, 1988, 1990; Andriessesn et al., 1993). Present day plate velocity vectors measured
from satellite data indicate that the Colombian Andean region (Andean Block) is moving north-
eastward with a right-lateral transpressional movement relative to the cratonic South American Plate
(Freymueller et al., 1993; see Chapter 5).

136
Chapter 4

MIDDLE Maracaibo

San r a m a n
B uc
MIOCENE Basin

Lucas

tan
a
1400

s
PALEOGEOGRAPHY Cucuta

de
der
de San

An
Positive relief Barinas

Mas u lt

a
ga
Basin

id
Alluvial fan and fluvial

er
sif
fa
Serrania

M
Coastal plain predominanatly
sandstones 0 100 200 Km
1300 Coastal plain predominanatly Bucaramanga

y
lle
mudstones

Va

a
en
m
Littoral to inner shelf sandstones Arauca

ste

l
da
len
Shallow marine inner shelf

sy

g
da

Ma
carbonates

u lt
Ma

in
Shallow marine inner shelf

C) Su lazo
as
a

o
1200

in
le
mudstones and siltsones

os
b-B
b
a
dd

s
Ta

Re gam
n

Ba
u

Mi

li

C
Outer shelf shales or carbonates

Sa

n
ed

So
g io
c uy
ert
Turbiditic sandstones

La

(E

tem
nj a
Inv

Co
Tu
ra

ys
Tunja
bas in

ys te o

ille

ts
Sub- a
er a

Inv
d

ul
m

rd
Ho n

f au a Cam

fa
1100
Co
d ill

o
la e
Yopal
Co r

go d t e r n
lt s

pa
im

t
Manizales
ul
s

o
Bitu

fa
ba E a

m
ta e

ra
a
as in

ca
Sub-b t

ez
Bo na

Llanos Orientales
do

aI
m
Girar

Gu
ha
Sa

Bogota
C

1000 Ibague
y
alle

s y ena
a V

em
l
st

Villavicencio
da
len

ag
da

M
t
g

ul
Ma

900
fa
r
pe
tr a l

t
Up

ul
fa
C en

as in

i ra
Su e i v a

Neiva
m
b- b

ta
N

Al

La Macarena

San Jose del


Serrania de

800
Guaviare

800 900 1000 1100 1200 1300 1400 1500

Figure 4.2: Middle Miocene palaeogeography without palinpastic restoration. (modified from Geotec, 1992 and
Cooper et al., 1995).

3. STRATIGRAPHY

In this section I summarize the Neogene stratigraphy (Figures 2.7 to 2.11). Because the
deformation and rock-uplift history of the EC is recorded by the sedimentary record of the Sabana de
Bogot and the MV and LLA basins (Figs. 4.2 to 4.7), I also included additional geological evidence to
constrain the Neogene tectonic history of the EC.
The Middle Miocene to Present late Tertiary sedimentary record of the study area is present in
the LLA and MV Neogene Andean foreland basins (Cooper et al., 1995,) and locally also in the Sabana
de Bogot (Fig. 1.2 and Figs. 4.2 to 4.7).

3.1. LLANOS ORIENTALES (LLA)

According to Cooper et al. (1995), during Middle Miocene a global tectono-eustatic base-level
rise coincided with the first significant deformation in the EC. Evidence for at least partial emergence

137
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes
of the EC is the presence of more sand in the western foothills than in the east (Cooper et al., 1995). In
the LLA foothills a moderate compressional tectonic phase occurred during Middle Miocene, ending

Maracaibo

B uc a
Serrania
de Perija Basin
MIDDLE

ra m a
Lucas

s
Cucuta

de
An
MIOCENE

ng a
Barinas

de San

a
Basin

id
THICKNESS

fa ult

er
M
(meters)

Serrania
Bucaramanga
Arauca

m
sy a
n
st e

s yste r a
ali

m a
f au a S

m
L
lt

na T
Medellin
ra

C us ia
fa ult
di lle

t em eo
ys al a
lt s o p
Cor

f au ram
Tunja Llanos Orientales

ic a
Basin (LLA)
ys te bao

a
Gu
m
f au a Cam

Manizales Yopal
lt s

t
ul
im

fa
Bitu

a
ez
m

Bogota
ha
C

Ibague
sy ena
em

Villavicencio
l
st
u lt g d a

-
0 100 200 Km
a
M
fa

900
al
ntr
Ce

t
ul

Neiva
fa
ir a

200
m

La Macarena
Serrania de

800
ta

San Jose del


Al

Guaviare
100
0

800 900 1000 1100 1200 1300 1400 1500

Figure 4.3: Middle Miocene restored thickness (meters) without palinpastic restoration. Thick lines represent
palaeo-faults believed to be active during Middle Miocene time.

the previous monotonous sedimentation (Casero et al., 1995, 1997). The resultant loading tectonically
enhanced the highstand system tract and resulted in deposition of the green to dark grey Leon Fm (Figs.
2.7 to 2.11 and Figs. 4.2 and 4.3, Cooper et al., 1995) shales with iron nodules and locally organic-rich
shales, deposited in anoxic conditions (Moreno and Velzquez, 1993). In the upper part an upward
colour change from grey to red has been interpreted as a change from marine to continental deposition
(Cooper et al., 1995).
Later (since Late Miocene?) the coarse synorogenic continental clastic deposits of the Guayabo
Fm were deposited in the LLA basin (Figs. 2.7 to 2.11 and Figs. 4.4 and 4.5 to 4.7). These sediments
recorded the exhumation of the EC as evidenced by clasts of Cretaceous rocks (clasts of mudstones,
cherts) derived from erosion of the EC (Moreno and Velazquez, 1993). These authors recognized an
erosional unconformity between the lower and middle portions of the Guayabo Fm. in the LLA
foothills (near Nunchia).

138
Chapter 4
Seismic lines across the LLA foothills (Figs. 2.1 and 4.7) show indications of severe thrusting
of post-Leon age that seem to be overlain by Guayabo beds (Casero et al., 1995, 1997). Seismic (Figs.
4.6 and 4.7) and well data indicates a gradual westward increase of thickness to a maximum in the
thrust faults defining the eastern boundary of the EC, and a virtually undeformed Neogene section on
most of the LLA area (Figs. 4.3, 4.5 and 4.6). However, in the south-western LLA, NE of the Serrana
de La Macarena, seismic lines (e.g. Fig. 4.7) show that the upper part of the Neogene section (Guayabo
Fm) has been partially eroded. This explains the decrease in preserved thickness in this part of the
basin. Ecopetrol and Beicip (1995) have interpreted in this area some flower structures probably related
to the Andean deformation. Wrench faulting in the LLA also has been proposed by Cediel (1982).
Vsquez (1988) has interpreted a strike-slip component in several faults in the LLA Basin.
Development of local rock-uplift associated with development of positive flower structures may
explain the erosion of the Guayabo Fm in this area. It is interesting to note that southward in the
Putumayo Basin (Fig. 1.1) the thinner Ospina Fm is considered time equivalent to the Guayabo Fm (but
palaeontological arguments are weak) while the eastern flank of the Cordillera in that area is
characterized by dextral strike-slip Altamira/Algeciras faults (Casero et al., 1995, 1997).
LATE Serrania Maracaibo
de Perija
Lucas

MIOCENE Basin

San
B uc

s
1400 PALEOGEOGRAPHY Cucuta

de
tand
de San

ar a

An
Positive relief Barinas
ma

er

a
Basin

id
ng a

Alluvial fan and fluvial


Mas

er
Serrania

M
Coastal plain predominanatly
sif
f au

sandstones
1300
lt

Coastal plain predominanatly Bucaramanga Arauca


mudstones
a

Littoral to inner shelf sandstones


len
m

m ara
ste

C)
da

Shallow marine inner shelf


(E

m
g
sy

carbonates

sy ste
M

na Ta
a sin

Shallow marine inner shelf


lt

sin

er
f au

zo

1200
l

mudstones and siltsones


l
a

i
bla

rd

b-B
b-B
a

Co

Outer shelf shales or carbonates


Ta

Cus ia
fau lt
lin

Su

Su
Sa

ed

rn

Turbiditic sandstones
s te
e rt

cu y
La

Inv

Ea

Llanos Orientales
Co
in

ys te bao
da

Basin (LLA)
em

Tunja
-b as

Inv
m

1100
Ho n

f au a Cam

t
ys

Yopal
ts
Sub

ul
lt s

t
ul

Manizales
fa
im

fa

o
Bi tu

lae
ta e
as in

a
y Sub-b t
ra

Bo na d
o

ez

pa
Girard
d ille

am

o
m
ba
go

Ch

ra
Sa
Co r

1000 Ibague
ai

Bogota
Gu
t r al

lle
Va

s y ena
Ce n

em
na

l
st

Villavicencio
da

-
0 100 200 Km
ale

ag
gd

M
t
Ma

ul

900
fa
r
pe

ul t
Up

fa
ir a
m
ta
Al

La Macarena

Neiva
n

Serrania de
S u e iv a
as i

800
b- b

San Jose del


N

Guaviare

800 900 1000 1100 1200 1300 1400 1500


Figure 4.4: Late Miocene palaeogeography without palinpastic restoration. (modified from Geotec, 1992 and Cooper
et al., 1995).

Detrital source areas. Clast composition of the Guayabo Fm conglomerates changes upward.
The lower part contains clasts of mudstone (one Oligocene clast as dated by palynology) and muddy

139
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes
sandstones with coal fragments derived from the Carbonera Fm. The upper part contains clasts of chert
or clasts with glauconite derived from Cretaceous rocks of the EC (Moreno and Velazquez, 1993). The
coarsening upward trend reported for the Guayabo Fm (Ecopetrol and Beicip, 1995) might represent an
increase in grain size of clasts derived from a growing topography of the EC.

3.2. MAGDALENA VALLEY (MV)

Most of the Neogene record is continental (Figs. 2.7 to 4.2 and 4.4). Thicknesses of Neogene
strata are highly variable (Fig. 4.5). In the Middle MV the major units are the Real Gp and the Mesa Gp
In the Upper MV and the Honda sub-basin the major units are the Honda Gp and the Mesa Gp (Figs.
2.7 to 2.11). Although the general lithology and relative stratigraphic position of units with the same
name in different areas of the MV is similar, there is no evidence that these units are time equivalents
or that they represent the same sedimentary events. Miocene deposits unconformably truncate early
structures, evidencing deformation and erosion (Morales et al., 1956; Guillande, 1988). According to
Jordan and Gomez (1996) the bases of the Honda and Mesa Gps vary spatially from conformable to
angularly unconformable. This geometric variability, along with strong thickness and facies variations,
suggests both palaeogeographic and tectonic controls on their accumulation.

Serrania Maracaibo
Bu ca

LATE
s

Basin
an Luca

de Perija
rama
0

s
MIOCENE
500

1400

de
Cucuta
2000 150 000

An
n
1

Barinas
0

QUATENARY
ga fa
a de S

a
Basin

id
ult

er
THICKNESS
M
Serrani

(meters) 0 100 200 Km


1300 Bucaramanga
m
sy 00 a

Arauca
0

n
st e
a li

4
fa u a S
0

L
100 0

lt
50
150 0

Medellin
0

1200 4000
3500
u

C
em o
sy a m

00
30
00

ul t i ca r
st

00
ys t e bao

Llanos Orientales
10

00
25
a
m

20
Gu
f au a Cam

Tunja Basin (LLA)


00
fa

1100
15
0
500
0

lt s
150

Yopal
im

100

Manizales
Bit u

ult
fa
a
ez

500
m
ha

720
Ibague
C

1000 Bogota
s y ena
em

Villavicencio
l
st
ul g da
a
M
t

900
fa
ul t
fa
i ra
m
ta

La Macarena

Neiva
Al

Serrania de

800 0
San Jose
del Guaviare

800 900 1000 1100 1200 1300 1400


Figure 4.5: Late Miocene restored thickness (meters) without palinpastic restoration. Thick lines represent palaeo-
faults believed to be active during Late Miocene time.

140
NW SE
AR 91 3 SAN 84-42a RE 901553 C 80 1285 C 85 10
La Maria-1 La Cabana-1 La Gloria-1 Centauro-1 La Punta-1 Guarimena-1 Morichito-1 Cano Barulia-1

0 Guayabo 5 0
Guayabo 4
Guayabo 3
1
Guayabo 2 1
Guayabo 1
2 Top Leon Fm
2
Top Carbonera Fm
3
Top Mirador Fm 5 km
3

141
Top Palaezoic
4
5 km 5 km
Chapter 4

Top Basement
5
sec
5 km
5 km

Figure 4.6: Interpretation of a regional seismic line trough the Llanos Orientales Basin showing the stratigraphic markers reflectors
used in flexural modelling. Location of the section shown in Figure 4.1.
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes
3.2.1. Middle Magdalena Valley

Real Gp. According to Cooper et al. (1995) the Real Gp lies unconformably over the Colorado
Fm (Figs. 2.7 to 2.11). They assume a Middle Miocene hiatus and stated that such an unconformity is
evidence for deformation, exhumation and erosion of the EC. However, the age of the unit has not been
confidently determined. Morales et al. (1956) reported a Miocene age based on fossil leaves and a
horizon of small gastropods found in the Velasquez field. The Real Gp consists of conglomerate,
sandstone and minor mudstone, representing alluvial fan and braided fluvial deposition (Figs. 4.2 and
4.4, Geotec, 1994). Its maximum thickness is 3600 m near the EC but it thins westward (Fig. 4.5).
Lateral thickness variations in the Middle MV, particularly a minimum of thickness in the Cachira
palaeo-high (Fig. 4.5) suggest less subsidence in this area that was deformed simultaneously during
deposition. According to Geotec (1994), piggyback deposits are common. Some faults affecting the
Neogene sedimentary record confirm Neogene deformation. (Porta 1965, 1966; Servicio Geologico
Nacional, 1967a,b; Ward et al., 1973; Restrepo-Pace et al., 1999a,b). In the Honda sub-basin the Honda
Gp. consists of greenish gay sandstones (Wellman, 1970). From this Gp Gmez (1999) reported
radiometric and volcanic fission track ages between 11.92 0.11 Ma and 7.5 0.6 Ma.
Detrital source areas. Sandstones are characterized by a predominance of plagioclase and
volcanic rock fragments with more plagioclase than K-feldspar and non undulatory quartz, pyroxene
and other mafics as secondary components (Hathon and Espejo, 1997). Morales et al. (1956) reported
hornblende, augite and a large amount of igneous and volcaniclastic material. These sandstones fall
into the magmatic arc provenance field in a quartz, feldspar, lithic plot (Dickinson, 1985). Rare
plutonic grains suggest a dissected arc provenance. The fresh volcanic fragments indicate a near detrital
source inferred to be the Miocene age volcanics that blanket partially the Central Cordillera (Figs. 1.1
and 4.2; Hathon and Espejo, 1997). Palaeocurrent data also suggest dominant eastward flow of
sediments (Rubiano, 1998). There is also evidence of detrital source areas located in the EC.
Conglomerates contain black chert, mudstones and coal fragments (Morales et al., 1956) possibly
derived from the EC. According to Ecopetrol et al. (1994) palynological evidence suggests these
sediments were sourced predominantly from the EC.
Mesa Gp. This unit lies unconformably over Real Gp (Figs. 2.7 to 2.11; Cooper et al., 1995).
In the Honda sub-basin the Mesa Gp rests with angular unconformity. Its age is Early Pliocene in the
Honda sub-basin as determined radiometrically by Thouret (1988). In the Honda sub-basin its age is
Early Pliocene (radiometric age 3.5 to 4.3 Ma: Thouret,1988; palynology: Dueas and Castro, 1981). It
consists of lithic conglomerates, sandstones and andesitic tuffs containing pyroclastic fragments. These
sediments were derived from the Central Cordillera and EC (Figs. 1.1 and 4.5) and were deposited and
reworked in alluvial fans and fluvial systems (Fig. 4.5). Thickness reaches up to 575 m. (Morales et al.,
1956). Porta (1965, 1966) has reported occurrence of the Pliocene Mesa Fm, limited towards the east
by the Honda Fault, indicating fault-controlled sedimentation.

3.2.2. Upper MagdalenaValley

According to Geotec (1994) sedimentation of the Middle to Upper Miocene Honda Fm shows a
piggyback basin fill relationship with the west-verging thrusts that delimited the western flank of the
EC and the Upper MV (Prado, Agua de Dios, Chaparral, Pitalito, Garzon faults). In the northern
Girardot sub-basin Amezquita and Montes (1994) and Amaya and Santamaria (1994) demonstrated the
activity of both an east- and west-verging fold-and-thrust systems. Both systems were active before and
after sedimentation of the Middle to Upper Miocene Honda Fm. Lateral changes of thickness of the
Middle to Upper Miocene Honda Gr also suggest piggyback sedimentation.
In the southern Neiva sub-basin (Figs. 2.1, 4.2 and 4.4) sedimentation occurred on distal
volcanic aprons derived from the Central Cordillera volcanic arc, and braided river systems flowing
toward the east (Figs. 4.2 and 4.4, Van der Wiel, 1991). In the Neiva sub-basin the age of the Honda
Gp is latest early Miocene to late-Middle Miocene (16.1 to 11.5 Ma), based on palaeontologic
(Villarroel and Guerrero, 1984; Busbey, 1986; Czaplewski, 1989; Hirsfeld and Marshall, 1976;

142
W
Santa Maria Fault E
Tesalia Fault

Humadea-1
Chichimene-1
La Libertad
Suria-1
Negritos

Apiay-1
?

0 0
1 1
2 2
3 3

TW T(SEC)
4 4
5 0 10 20 km
5

Neogene (Leon and Guayabo Fms.)

143
Palaeogene (Mirador and Carbonera Fms.)
Chapter 4

Cretaceous
Cambrian-Ordovician
Precambrian Cristaline basement

Figure 4.7: Interpretation of a regional seismic section trough the southern part of the Llanos Basin north of the Serrania de la
Macarena. Note the local and partial erosion of the Neogene Guayabo Group. Location of the section shown in Figure 4.1 (from
Ecopetrol and Beicip, 1995).
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes
Setoguchi and Rosenberg, 1985, 1987; Kay et al., 1989), radiometric and magnetoestratigraphic data
(Hayashida, 1984; Guerrero, 1993). It consists of grey to tan feldspatic sandstone and drab to red brown
mudstone and conglomerate, locally more than 1600 m thick. Conglomerates and sandstones contain
abundant volcanic fragments, feldspar, quartz, chert, metamorphic and plutonic rocks. During
deposition dacitic to andesitic volcanic activity increased in the north part of the Neiva sub-basin
(Wellman, 1970).
The Huila Gp (former Mesa Gp proposed by Howe, 1974 in the Upper MV; Guerrero, 1993) in
the Neiva sub-basin it is of Late Miocene age, based on absolute ages (Takemura and Danhara, 1983;
Setoguchi and Rosenberg, 1985; Van der Wiel and Van der Bergh, 1992a,b; Guerrrero, 1993). The
lower part of the Huila Gp in the Neiva sub-basin (Neiva Fm; Guerrero, 1993)) consists of
conglomerates with clasts of intermediate volcanic rocks, locally with abundant granitic pebbles.
Volcanic arenite is the dominant sandstone with plagioclase. There are tan coloured mudstones. Detrital
composition and palaeocurent data suggest these volcaniclastic sediments were mainly derived from
the Central Cordillera, with some local exhumation of the Garzon Massif (Howe, 1974). The Huila Gp
rests on angular unconformity on the Honda Gp (Guerrero, 1993)
The Honda and Huila Gps were deposited in a piedmont plain by coalescing alluvial fans and
braided to meandering fluvial systems (Figs. 4.2 and 4.4), related to pulses of exhumation of the
Central Cordillera. These units also record andesitic to dacitic volcanism of the Central Cordillera. The
record of volcanism is best preserved in the Huila Gp (Van Houten and Travis, 1968; Howe, 1969; Van
der Wiel and Van der Bergh, 1992a,b; Guerrero, 1993).
Very coarse Pliocene to Recent clastic deposits are found along the western and eastern
margins of the basin. The Upper Cenozoic record (Ibague Fan, Chaparral Fan, Ceibas Formation and
Gigante Fm.) contains abundant volcaniclastic deposits derived from stratovolcanoes on the Central
Cordillera. About 7 to 5 Ma ago several polymictic debris-flows and torrential sediments formed fans
that spread across the Upper Magdalena Valley (Van Houten, 1976).
In the southern part of the Upper Magdalena Valley changes in palaeocurrents and facies show
that the exhumation of the Garzon Massif (Fig. 2.1) began during accumulation of the upper part of the
Honda Gp (late Middle Miocene) and continued during deposition of the Mesa Gp (Van Houten and
Travis, 1968; Van der Wiel and Van der Bergh, 1992a,b; Guerrero, 1993).
As mentioned in Chapter 3 westward palaeocurrent directions in the Villavieja Fm (upper part
of Honda Gp of the Upper MV) indicate that the southern EC was already a mountain range that
completely closed the basin in its eastern side at about 11.8 Ma (ages from 40Ar/39Ar dates and
magnetostratigraphy, Guerrero, 1993). Changes of these palaeocurrents indicate that exhumation of the
southern EC began at 12.9 Ma (Guerrero, 1993). Fission track data indicate that around 12 Ma. the
Garzon Massif and the Central Cordillera (Fig. 2.1) were simultaneously exhumated. Such rock-uplift
and exhumation caused rapid relative subsidence of the intermontane Neiva sub-basin and deposition of
increasingly coarse detritus from the west (Van der Wiel, 1991). After 12 Ma tectonic activity of both
the Central Cordillera and the Garzon Massif (Fig. 2.1) waned and sediments began to exhibit eastward
and northward palaeocurent directions (Van der Wiel, 1991). Volcanic activity and exhumation of the
Central Cordillera was resumed 10-9 Ma at the start of deposition of the Gigante Fm (Neiva member).
Exhumation of the Central Cordillera occurred in its southern part where it merges with the Garzon
Massif (Figs. 1.1 and 2.1; Van der Wiel, 1991). Between 8 and 6.4 Ma activity of the Central Cordillera
volcanic arc reached a maximum and because of the high input of volcanic and volcaniclastic material
into the basin, the north-flowing palaeo-Magdalena River was forced by volcanic aprons (Los Altares
member of Gigante Fm; Van der Wiel, 1991) to flow along the eastern border of the basin. Around 6.4
Ma ago a new pulse of tectonic exhumation affected the Garzon Massif (Fig. 2.1), which supplied an
increasing amount of metamorphic pebbles and boulders into the basin (Garzon Member of the Gigante
Fm) although sediment also came from the Central Cordillera. Rock-uplift of the massif culminated 6
Ma ago (Van der Wiel, 1991).
During rock-uplift strong SE-NW compression produced folding and reverse faulting in the
basinal sediments. Later sediments (lower unit of the Las Vueltas Fm) were immediately deformed
after deposition, forming a small angular unconformity with older deposits. Rock-uplift of the Garzon
Massif (Fig. 2.1) probably occurred along several faults of the Garzon-Suaza fault system. The latter
had intermittent pulses of strike-slip movement, leading to the formation of several generations of

144
Chapter 4
alluvial fans coming from the east. Strike-slip activity of the Garzn-Suaza fault system affected
Quaternary sediments in local basins on the Garzn Massif (Fig. 2.1). In the basin Quaternary
sediments have been affected by reverse faulting, indicating NW-SE compression (Van der Wiel,
1991).
3.3. SABANA DE BOGOT

The Sabana de Bogot (Figs. 2.1, 4.4 and 4.5) is a high plain in the axial part of the EC at an
altitude of 2600 m. It is an intermontane basin consistent in a broad NNE-SSW-striking synclinorium
probably formed during Oligocene to Late Miocene time. During late Pliocene and Pleistocene
approximately 600 m of mainly lacustrine sediments accumulated (Helmens, 1988, 1990; Andriessen et
al., 1993). The Neogene and Quaternary fluvial to lacustrine sedimentary section of the Sabana de
Bogot registers major tectonic surface-uplift for the period between 5 and 3 Ma (Helmens, 1988,
1990; Andriessen et al., 1993). A summary of this stratigraphy is the following:
The Neogene-Quaternary stratigraphic record of the Sabana de Bogot area starts with the Late
Miocene Marichuela Fm, which represents very large synorogenic debris flows and gravity flows that
aggraded on broad alluvial plains and lakes deposited during a period of increased regional tectonic
activity, with locally strong deformation of strata (Helmens, 1990; Andriessen et al., 1993).
The earliest Pliocene Tequendama Member of the Lower Tilata Formation represents a period
of relatively quiet fluvial sedimentation. Palynological data indicate forest vegetation of tropical
lowland type with an altitude of deposition that did not exceed 500 m (Helmens, 1990; Andriessen et
al., 1993). Tilata Fm rests with a pronounced angular unconformity over a variety of older strata,
confirming that some deformation preceded its deposition (Cooper et al., 1995).
The fluvial and alluvial fan sediments of the early Pliocene Tibagota member of the Lower
Tilat Fm were deposited at the beginning of the final major upheaval of the EC. Palynological and
macrobotanical data indicate a depositional environment in the lower tropical to lower sub-Andean
forest belt, between 1000 m and 1500 m (Helmens, 1990).
The Late Pliocene Guasca Member of the Upper Tilat Fm recorded fluvial to lacustrine
sedimentation in the outer valleys of the present high plain area. Palynological data from that unit
indicates deposition in the upper sub-Andean forest belt, at an elevation of about 2200 m (Helmens,
1990; Andriessen et al., 1993). Magnetostratigraphic data indicates that the Guasca Member was
deposited between 3.2 Ma and 2.58 Ma (Helmens et al., 1997). Later the Late Pliocene unnamed Upper
Member of the Upper Tilata Fm recorded sedimentation in the central part of the present high plain of
Bogot, when the main surface-uplift of the Sabana de Bogot area had ceased. The Tilata Fm and
correlative sediments are found on the slopes surrounding the Sabana de Bogot, and their beds are
often tectonically disturbed (Helmens, 1990).
During the Early Pleistocene, lacustrine fluvial deposition of the Subachoque Fm occurred in a
large area under alternately glacial and 'interglacial conditions at an altitude similar to the present
day elevation of the Sabana de Bogot plain (which is 2600 m). The Subachoque Fm and younger
stratigraphic units are tectonically undisturbed (Helmens, 1990; Andriessen, 1993).
Later deposition of Pleistocene Sabana Fm in the central part of the present high plain of
Bogot occurred in a lake environment (Helmens, 1990; Andriessen et al., 1993).
Palynological evidence indicates the replacement of a tropical lowland flora by a high mountain
flora occurred during the Pliocene. This trend in vegetation change has been interpreted as being
primarily the result of tectonic surface-uplift (Van der Hammen et al., 1973). Wijninga and Kuhry
(1993), however, note than warmer climate in the Pliocene should have placed the altitudinal
vegetation belts several hundred metres higher. Probably surface-uplift took place in various phases.
Phases of high tectonic activity, with coarse gravity flow deposition alternated with phases of fluvial
lacustrine sedimentation and soil formation (Helmens, 1990).

4. TECTONIC SUBSIDENCE DURING THE NEOGENE

The study of the sedimentary fill of a basin and its associated tectonic subsidence signal give
important information about basin formation mechanisms (Cloetingh et al., 1993). In order to quantify
the tectonic component of subsidence of the Neogene basins of the study area, a 1D-backstripping

145
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes
technique was used (Steckler and Watts, 1978; Bond and Kominz, 1984). Details of the backstripping
subsidence analysis, the procedure and data are described in Chapter 2.
The late Tertiary sedimentary record of the study area is present in the LLA and MV basins and
locally in the Sabana de Bogot. Because the Neogene sedimentary record of the LLA and Middle MV
has not been studied, there are no reliable age data and I only used a limited number of sections.

4.1. RESULTS

4.1.1. Llanos Orientales (LLA)

The convex upward shape of tectonic subsidence curves (Fig. 3.16) clearly shows the
acceleration of tectonic subsidence in a typical subsidence pattern, as observed in foreland basins
(Allen and Allen, 1990). Maximum increase in tectonic subsidence occurred during the Neogene (last
10 Ma, see Fig. 3.16) during the Andean orogeny. The amount of tectonic subsidence is maximum (up
to 1500 m, in well Arauca-1) near the thrust front of the Cordillera and gradually decreases eastward
(e.g. well La Tortuga-1 450 m, Fig. 3.16).

4.1.2. Magdalena Valley (MV)

Middle Magdalena Valley. The thickness distribution map (Fig. 4.5) indicates a general
increase in thickness toward the EC in a similar pattern to foreland basins. However, some lateral
thickness variations in the Middle MV, particularly a minimum of thickness in the Cachira palaeo-high
(Fig. 4.5), suggest that less subsidence took place in this area that, which was deformed simultaneously
during deposition.
Curves of tectonic subsidence (Fig. 3.16) show large variability. Some curves show a convex
upward shape with acceleration of tectonic subsidence especially during Neogene time, as observed in
foreland basins (Allen and Allen, 1990). At other localities tectonic subsidence and sedimentation were
compensated by local erosion resulting in a horizontal segments in the tectonic subsidence curves.
Upper Magdalena Valley. Lateral changes of thickness of the Middle to Upper Miocene Honda
Gr Also confirm piggyback sedimentation.
In the Neiva sub-basin (Fig. 2.1) subsidence in the area east of the Chusma thrust system
started during the Early Miocene, probably more than 16 Ma with deposition of the Honda Fm just
after the last phase of activity of the Chusma thrust system. The western limit of the subsiding area was
the easternmost Dina thrust. Further to the east, the subsiding area was probably bordered by very low
isolated hills represented by erosional remnants of the Garzon Massif (Fig. 2.1), which had been
exhumated about 100 Ma ago. Between <16 Ma and approximately 12 Ma subsidence was not yet very
large (Van der Wiel, 1991).
Tectonic subsidence curves of the Upper MV also show variability suggesting tectonic
deformation and emergence of local blocks due to thrusting, with local erosion and simultaneous
sedimentation in other parts of the area. Results of thermal modelling of vitrinite reflectance data
(Buitrago, 1994) suggest that instead of flat subsidence curves, sedimentation and later erosion affected
some local thrust faulted blocks. This latter case of subsidence and sedimentation was compensated by
local erosion and indicates local exhumation and erosion of blocks, this suggests that tectonic
deformation affected some blocks while other areas were subjected to sedimentation.
Figure 4.8 shows a map of the observed tectonic subsidence during Neogene time. This figure
and the thickness maps (Figs. 4.3 and 4.5) indicate subsidence in the Neogene LLA and MV basins and
absence of subsidence in the area of the EC.

5. NEOGENE REMAINING THERMAL SUBSIDENCE AFTER MESOZOIC RIFTING

In Chapter 3 I concluded that during Palaeogene time there was a partial component of
subsidence produced by thermal equilibration of the lithosphere after the Mesozoic extensional events.
I also tested through modelling the possibility of some Neogene remaining thermal subsidence after
Mesozoic lithosphere extension. Here I used the same methodology applied to Chapters 2 and 3: I

146
Chapter 4
calculated the remaining thermal subsidence due to thermal re-equilibration of the lithosphere during
Neogene. A description of the forward modelling technique and procedures developed by Van Wees et
al. (1998) is given in Chapter 2. I use the lithosphere stretching factors calculated for Mesozoic rifting
episodes (see tables Chapter 2).

5.1. RESULTS

Results of these calculations appear in Figure 4.9. Comparison of the calculated thermal
subsidence during Neogene time (Fig. 4.9) with the observed tectonic subsidence (Fig. 4.8) shows that
the latter cannot be explained by thermal re-equilibration of the lithosphere. Subsidence maps show that
the remaining thermal subsidence assuming tectonic quiescence would have affected the area of the
EC, where the Mesozoic extensional basin was located. The observed tectonic subsidence is just the
opposite with zero values in the area of the EC, and maximum values in the Neogene LLA and MV
basins. This opposite trend is also clearly seen comparing the thickness maps of the Lower Cretaceous
(Figs. 2.13, 2.15 and 2.17) and the Neogene (Figs. 4.3 and 4.5). These maps illustrate that the Mesozoic
extensional basin located mainly in the area of the EC was completely inverted during Neogene time to
form the mountain range. The Neogene sedimentation of the Sabana de Bogot area (Figs. 4.4 and 4.5),
which apparently is an exception, just recorded the surface-uplift history of the Cordillera.
B uc

Serrania Maracaibo
Lucas

de Perija
ara

Basin
BACKSTRIPPED

s
Cucuta
m an

de
MIDDLE
de San

An
MIOCENE Barinas
ga

a
QUATERNARY Basin

id
f au

er
Serrania

SUBSIDENCE
M
lt

(m) 0.00 0.00


440.00
88.00 160.00 Bucaramanga
Arauca
134.00 472.00
0.00 m ara
67.00 900.00
0.00 0.00 m
0.00
0.00
s ys te
520.00 201.00
na T

0.00
ys t alaeo

Medellin
C us ia
fault

509.00
ra

0.00 0.00
em

1.00
p
d i ll e

0.00
f au am o

457.00
0.00
lt s

0.00
r

485.00 169.00
Cor

aic

0.00
Gu

0.00
Tunja Llanos Orientales
ys te bao

0.00
495.00
0.00 366.00
m

Basin (LLA)
f au a Cam

0.00 646.00
0.00 318.00
0.00 638.00
0.00
Yopal
lt s

497.00
im

485.00
ult

Manizales 327.00
Bi tu

fa

332.00 726.00 516.00


0.00 658.00 349.00
0.00 105.0 528.00 369.00 343.00
a

424.00 360.00
ez

0.00 0
538.00 373.00
-1669.00 641.00 399.00
am

270.00
0.00 0.00 467.00 354.00 337.00
-1505.00 478.00 370.00
Ch

472.00 325.00
1.00 1.00 283.00
Ibague Bogota 492.00 501.00
288.00
294.00
0.00 0.00
0.00
167.00 414.00
306.00 275.00 216.00
153.00
323.00
Villavicencio 0.00 141.00
300.00 167.00
183.00 208.00
248.00 0.00 254.00
0.00 53.00 196.00 169.00
62.00 240.00
548.00
0.00 219.00
900
al

0.00 295.00 178.00


ntr

177.0
224.00 0 170.00
ult
Ce

fa

178.00
233.00
ira

215.00 130.00
216.00 216.00 0
m

Neiva
ta
Al

800 107.00
La Macarena

356.00
-
0 100 200 Km
Serrania de

San Jose del


489.00 Guaviare
489.00

800 900 1000 1100 1200 1300 1400 1500


Figure 4.8: Observed tectonic subsidence, in meters, during Middle Miocene to Recent. Observed tectonic
subsidence calculated from backstripping analysis.

147
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes

Serrania Maracaibo

s
de Perija Basin

an Luca
THERMAL

s
de
MIDDLE Cucuta

An
MIOCENE Barinas

a de S

a
QUATERNARY Basin

id
er
SUBSIDENCE

M
Serrani
(m) 0.00 -
-104.00
0 100 200 Km
0.00
0.00 0.00 Bucaramanga
-172.00

- 0.00
0.00 0.00
50.00
-
0.00
Arauca
0.00 50.00
- 0.00
0.00 0.00
47.00
0.00
0.00
-139.00
Medellin -
0.00 149.00 0.00
0.00

0.00
-69.00 0.00
- -
124.00 26.00 0.00 0.00 0.00
-80.00 -119.00
0.00

0.00 -121.00 Llanos Orientales


0 0.00
-302.00 Tunja 0.00
0.00
0.00
0.00
Basin (LLA)
0.00
-18.00 0.00
0.00
0.00
-291.00 Yopal 0.00
Manizales -309.00
0.00

0.00
-281.00
0.00 0.00
0.00
-289.00 0.00
0.00 0.00
0.00
Ibague 0.00
0.00
0.00 0.00
- -125.00 Bogota 0 0.00
102.00
-83.00 -109.00 0.00
0.00 0.00
0.00

Villavicencio
0 0.00 00.00
0.00
-
900 109.00
0.00

0.00 0.00
0.00

0.00
0.00

Neiva
La Macarena

0.00
Serrania de

800
San Jose del
Guaviare

800 900 1000 1100 1200 1300 1400 1500


Figure 4.9: Middle Miocene to Recent predicted thermal subsidence after Mesozoic rifting, in meters, calculated
from forward modelling using lithosphere-stretching factors discussed in Chapter 2. Negative values mean tectonic
rock-uplift.

6. NEOGENE FLEXURAL SUBSIDENCE

If the EC was generated as a mountain range during Neogene as proposed in literature (e.g.
Cooper et al., 1995), the lithosphere would bend due to the topographic loading (Price, 1973). As a
result, depressions would be formed at both sides of a mountain chain (MV and LLA Neogene basins)
in which sediments would accumulate. In this section I quantitatively test this hypothesis using a model
that assumes the lithosphere behaves as an elastic sheet floating on a fluid asthenosphere (Beaumont,
1981) and the observed Neogene subsidence of the MV and LLA basins. I tested the following
scenarios:
(1) Observed tectonic subsidence of the MV and LLA basins was produced by present day
topographic load of the Central and EC.
(2) Observed tectonic subsidence of the MV and LLA basins was only produced by present day
topographic load of the EC.
(3) Observed tectonic subsidence of the MV and LLA basins was only produced by present day
topographic load of the EC and the lithosphere behaved as a broken elastic plate under the EC.

148
Chapter 4
6.1. FLEXURAL MODELS PRODUCED BY THE PRESENT-DAY TOPOGRAPHIC LOAD
OF THE EASTERN AND CENTRAL CORDILLERAS

Initially I test the hypothesis that the Neogene tectonic subsidence of the MV and LLA basins
was produced by lithosphere flexural subsidence due to the present day topographic load of the Eastern
and Central cordilleras.

6.1.1. Methods

I modelled the flexural subsidence along 13 2D sections using the program Cobra developed by
Zoetemeijer (1993, 1998). An explanation of the flexural modelling principles and a description of the
program is included in Chapter 3. Location of the sections is shown on Figure 4.1. The modelling
parameters used are shown in Table 4.1. I applied the present day topography of the Eastern and
Central cordilleras along each cross-section to an elastic plate representing the lithosphere. I compared
the calculated flexural deflection to the thickness of the Middle Miocene to present-day sedimentary fill
of the LLA and MV basins. I changed the effective elastic thickness of the plate until an acceptable fit
was obtained between the calculated flexural deflection and the Neogene sedimentary thickness.
Initially I tried to use a constant effective elastic thickness, but after several runs the fit was not
satisfactory. Then I changed the thickness of the elastic plate as a function of the horizontal distance
along the section. After several trials I obtained an acceptable fit. The model assumes the flexural
deflection of the elastic plate is instantaneous. Because I applied the present-day topography as a load
the model represents the flexural response of the lithosphere at the present time. In continental foreland
basins the topographic load alone is usually enough to explain the observed flexural subsidence, but in
flexural basins associated with subduction of oceanic lithosphere, subsurface loads, plate boundary
forces or moments are usually required (Karner and Watts, 1983; Royden, 1993). The observed
subsidence of Neogene basins was fitted assuming surface topographic loads without any subsurface
load, boundary force or moment. As an additional constraint to the model the program allows to
calculate the gravity anomaly produced by the flexured lithosphere. For that purpose the program
applied the algorithm developed by Parker (1972), making use of additional data. These additional data
are: (1) The crustal thickness as a function of the horizontal distance along the section; and (2)
Densities of the crust, the mantle, the mountain range (topographic load) and the sediments filling the
flexural basins created. Gravity data were taken from the gravity map by Kellogg et al. (1991), crustal
thickness was taken from a Moho depth map calculated by Salvador (1991) using published gravity and
refraction sections in the south of Colombia (Meyer et al., 1973; Ocola et al., 1973; Ramirez and
Aldrich, 1973; Meissner et al., 1973, 1976; Mooney et al., 1979; Flueh et al., 1981). Density values
applied in the flexural models are displayed in Table 4.1.

ASSUMPTIONS

-Continuous vs. broken plate model, two possibilities tested: (1) Broken plate with a free boundary located
westward on the Romeral suture zone for models assuming a continuous plate under the Eastern Cordillera and
(2) Broken plate under the Eastern Cordillera.
-Topographic load: two possibilities tested: (1) Eastern and Central cordilleras load, (2) Eastern Cordillera load.
-No water depth (Neogene depositional environments were dominantly continental to transitional) except 60 to
80 m paleo-water depth during deposition of shallow marine mudstones of the Leon Fm. (Middle Miocene).
-No horizontal intraplate force.
-No vertical shear force at the free end of the plate.
-No bending moment at the free end of the plate.
-Sediment density applied as a load filling the basin, sediment thickness was used to compare to the calculated
flexural deflection.
-No subsurface loads.

Table 4.1: Assumptions and parameters used in lithosphere flexural models.

149
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes

PARAMETERS

-Length of the plate: 1000 km.


-Size of the finite difference grid cell: 5 km.
-Densities (kg/m3)
Sediments: 2250
Topographic load: 2550 to 2650
Crust: 2770
Mantle: 3300
-Effective elastic thickness and topographic load variable as shown in Figs. 4.10 to 4.14 and Tables 4.2 to 4.4.
-Young modulus of the plate: 7x10 10 N/m2
-Poisson ratio of the plate: 0.25

Table 4.1: Continued

6.1.2. Results

Results of these models are presented in Figure 4.10. Table 4.2 shows the effective elastic
thickness adopted to obtain the fit shown in Figure 4.10. In general the fit between the calculated
flexural subsidence of the lithosphere and thickness of the Middle Miocene to present-day sediments of
the LLA and MV is very good for the eastern LLA basin. However, the model results contain some
misfit in the MV area. Comparison of the calculated gravity anomaly with the observed Bouguer
anomaly indicates a good correlation at a wavelength of the same order of magnitude as the lithosphere
flexural deflection. However the observed gravity represents short wavelength variations that are not
comparable to the theoretical gravity values. Probably this high frequency gravity variation is due to
shallow lateral changes in crustal density as can be inferred from a surface geological map. The surface
geology of the EC is characterized by heterogeneous rocks of different densities. Metamorphic and
igneous rocks crop out in the Santander, Floresta, Garzon and Quetame massifs, forming the pre-
Mesozoic basement, while less dense sedimentary rocks of Palaeozoic to Cenozoic age constitute most
geological units at surface. These shallow crustal lateral density changes are not considered in the
model to calculate the theoretical gravity anomaly. Also the model includes a calculation of the stress
produced by bending the elastic plate (Fig. 4.10).

Distance Distance Distance


EET EET
from from from EET (km)
Section (km) Section (km) Section
Romeral Romeral Romeral
(km) (km) (km)
1 230 15 5 204 5 9 231 5
1 600 15 5 360 5 9 300 5
1 700 50 5 370 25 9 330 15
1 800 50 5 500 25 9 400 15
1 1200 50 5 740 55 9 600 40
2 230 5 5 1200 55 9 1200 40
2 390 5 6 173 5 10 218 5
2 400 20 6 370 5 10 390 5
2 500 20 6 390 25 10 400 15
2 550 30 6 520 25 10 600 15
Section number is shown in Fig. 4.1. EET: effective elastic thickness.

Table 4.2: Calculated effective elastic thickness (EET) in km.

150
Chapter 4
Distance Distance Distance
EET EET
from from from EET (km)
Section (km) Section (km) Section
Romeral Romeral Romeral
(km) (km) (km)
2 600 50 6 900 55 10 700 40
2 1200 50 6 1200 55 10 1200 40
3 172 5 7 240 5 11 296 2
3 380 5 7 420 5 11 400 2
3 390 25 7 430 25 11 410 15
3 525 25 7 650 25 11 571 15
3 750 50 7 800 55 11 700 40
3 1200 50 7 1200 55 11 1200 40
4 288 5 8 332 5 12 239 3
4 480 5 8 450 5 12 340 3
4 500 25 8 470 25 12 350 15
4 550 25 8 600 25 12 509 15
4 850 50 8 900 50 12 650 40
4 1200 50 8 1200 50 12 1200 40
13 237 5
13 480 5
13 530 40
13 1200 40
Section number is shown in Fig. 4.1. EET: effective elastic thickness.

Table 4.2: Continued

6.2. FLEXURAL MODELS PRODUCED BY THE PRESENT-DAY TOPOGRAPHIC LOAD


OF THE EASTERN CORDILLERA ONLY

In order to improve the misfit between calculated flexural deflection and thickness of Neogene
deposits in the MV, I re-calculated the models including only the topographic load of the EC. This also
was helpful to understand the effect of the topographic load represented by the Central Cordillera (on
the assumption that the topography of the Central Cordillera remained the same since Palaeogene time,
thus assuming no additional load was applied during Neogene time in this region). Results show a good
fit between the calculated flexural deflection of the lithosphere and the thickness of the Middle
Miocene to present-day sedimentary fill of the LLA and MV basins, significantly reducing the misfit
from earlier models (compare fit in the MV in Fig. 4.10 and Fig. 4.11).

6.2.1. Flexural deflection of the lithosphere as a mechanism of tectonic subsidence during Neogene
time

The good fit obtained for the model (Fig. 4.11) indicates that flexural subsidence produced by
the present-day topographic load of the EC can explain the observed Neogene subsidence of these
basins. A better fit between the calculated flexural deflection and the observed thickness of Neogene
sedimentary record support the hypothesis that the topography of the Central Cordillera has not
significantly changed since Palaeogene times. Of course, Neogene sediments derived from erosion of
the Central Cordillera suggest a reduction of topography that must have been balanced by tectonic
rock-uplift without significantly changing the topographic elevation. Almost all the sedimentary cover

151
t
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes

(o

n
e r
S ECTI O N 1 S ECTI O N 2
(t
n BOUGU E R A N O M A L Y BOUGU E R A N O M A L Y
ref.leve l : 0. m ref.level : 0. m
0 0
e

-70 -70

-14 0 -140

a
ai

0. 200. 400. 600. 800. 1000. 1200. 0. 200. 400. 600. 800. 1000. 1200.
distance (km) distance (km)

FLEXURE FLEXURE
5000 5000
o l)
e
n

0 0
c
d

-500 0
-5000
-100 00

0. 200. 400. 600. 800. 1000 1200. 0. 200. 400. 600. 800. 1000 1200.
. .
EFFECT I V E E L A S TIC THI C K N E S S EFFECT I V E E L A S TIC THI C K N E S S
80 80
60 60
f

40 40
20
ne

20
0
0
t

0. 200. 400. 600. 800. 1000 1200. 0. 200. 400. 600. 800. 1000 1200.
distance (km) . distance (km) .

S ECTI O N 3 S ECTI O N 4
BOUGU E R A N O M A L Y BOUGU E R A N O M A L Y
t

ref.level : 0. m ref.leve l : 0. m
0 0.

-70 -80 .
e

-140 -160 .
a

0. 200. 400. 600. 800. 1000 1200. 0. 200. 400. 600. 800. 1000 1200.
. .
distance (km) distanc e (k m )

FLEXURE FLEXURE
5000 5000
)
ne

0 0
c
d

-5000
-5000

0. 200. 400. 600. 800. 1000. 1200. 0. 200. 400. 600. 800. 1000 1200.
.
EFFECT I V E E L A S TIC THI C K N E S S EFFECT I V E E L A S TIC THI C K N E S S
80 80

60 60
40 40
20 20
e

0 0
0. 200. 400. 600. 800. 1000. 1200. 0. 200. 400. 600. 800. 1000. 1200.
distanc e (k m ) distance (km)

Figure 4.10: 2D flexural models produced by the present day topographic load of the Eastern and Central
cordilleras. Top panel: Dots represent observed Bouguer anomaly and continuous line calculated Bouger anomaly in
mgals. Middle panel: basement deflection and topography in meters. Dots represent the observed thickness of
Middle Miocene to Recent deposits in meters. Bottom panel: effective elastic thickness in km. Location of each
section in Figure 4.1.

of the Central Cordillera was removed. The eastward increase in thickness of Neogene sediments of the
Middle MV and their onlaping relations on the Central Cordillera suggest that the Middle MV was

152
o
Chapter 4

f
eastward tilted. Model results (Fig. 4.11) suggest that the Central Cordillera produced mainly a local
effect but not regional flexural subsidence effect. The reduced elastic thickness obtained for the western
n
border of the MV and Central Cordillera explain this more local subsidence and make flexural model
results difficult to extrapolate westward in the area of the Central Cordillera.

n
S ECTI O N 5 S ECTI O N 6
BOUGU E R A N O M A L Y BOUGU E R A N O M A L Y

0. ref.level : 0. m ref.leve l : 0. m
0.
e

e
-100. -80.

-200. -160.
a

a
0. 200. 400. 600. 800. 1000. 1200. 0. 200. 400. 600. 800. 1000. 1200.
distance (km)
distance (km)
FLEXURE FLEXURE
5000 5000

0 0
d

-5000

d
-5000

0. 200. 400. 600. 800. 1000 1200. 0. 200. 400. 600. 800. 1000 1200.
. .
EFFECT I V E E L A S TIC THI C K N E S S EFFECT I V E E L A S TIC THI C K N E S S
80 80

60 60

40 40
20 20
0 0
0. 200. 400. 600. 800. 1000 1200. 0. 200. 400. 600. 800. 1000 1200.
. .
distance (km)
distanc e (k m )

S ECTI O N 7 S ECTI O N 8
BOUGUER ANOMAL Y BOUGU E R A N O M A L Y
ref.leve l : 0. m ref.leve l : 0. m
0. 0.

-80. -100.

-160. -200.

200. 400. 600. 800. 1000. 1200. 0. 200. 400. 600. 800. 1000 1200.
distance (km) .
distance (km)
FLEXURE FLEXURE
5000
5000

0
0

-5000
-5000

200. 400. 600. 800. 1000. 1200. 0. 200. 400. 600. 800. 1000 1200.
.
EFFECT I V E E L A S TIC THI C K N E S S EFFECT I V E E L A S TIC THI C K N E S S
80 80

60 60
40
40
20 20
0 0
200. 400. 600. 800. 1000. 1200. 0. 200. 400. 600. 800. 1000 1200.
.
distance (km) distance (km)

Figure 4.10: Continued

153
f

f
t n
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes

( t
(
er

ne r
(

(
S ECTI O N 5 S ECTI O N 6
t

t
BOUGU E R A N O M A L Y BOUGU E R A N O M A L Y

)
0. ref.level : 0. m ref.leve l : 0. m
0.
e

y e
-100. -80.
i

i
-200. -160.
a

m a
0. 200. 400. 600. 800. 1000. 1200. 0. 200. 400. 600. 800. 1000. 1200.
m

distance (km)
distance (km)
FLEXURE FLEXURE
o l

n e ) o l
5000 5000
n e )

0 0
c

c
d

-5000

d
-5000

0. 200. 400. 600. 800. 1000 1200. 0. 200. 400. 600. 800. 1000 1200.
. .
EFFECT I V E E L A S TIC THI C K N E S S EFFECT I V E E L A S TIC THI C K N E S S
80 80
f

f
60 60

40 40
n

t en
e

20 20
t

0 0
0. 200. 400. 600. 800. 1000 1200. 0. 200. 400. 600. 800. 1000 1200.
. .
distance (km)
distanc e (k m )

S ECTI O N 7 S ECTI O N 8
t

BOUGUER ANOMAL Y BOUGU E R A N O M A L Y


ref.leve l : 0. m ref.leve l : 0. m
0. 0.
e

-80. -100.

-160. -200.
a

200. 400. 600. 800. 1000. 1200. 0. 200. 400. 600. 800. 1000 1200.
distance (km) .
distance (km)
FLEXURE FLEXURE
5000
5000
)

)
ne

ne

0
0
c

-5000
d

-5000

200. 400. 600. 800. 1000. 1200. 0. 200. 400. 600. 800. 1000 1200.
.
EFFECT I V E E L A S TIC THI C K N E S S EFFECT I V E E L A S TIC THI C K N E S S
80 80
)

60 60
40
40
20
e

20
0 0
200. 400. 600. 800. 1000. 1200. 0. 200. 400. 600. 800. 1000 1200.
.
distance (km) distance (km)

Figure 4.10: Continued

154
f
t n
Chapter 4

S EC TIO N 1 3

t
BOUGU E R A N O M A L Y

0. ref.leve l : 0. m

e
-160.

-320

0. 200. 400. 600. 800. 1000 1200.

a
.
FLEXURE
5000
ne )

0
c

-5000
d

-10000
0. 200. 400. 600. 800. 1000 1200.
.
EFFECT I V E E L A S TIC THI C K N E S S
80

60
40

20
e

0
0. 200. 400. 600. 800. 1000 1200.
.
distance (km)

Figure 4.10: Continued

6.2.2. Calculated bending stress

The program also calculated the bending stresses of the assumed thin elastic plate (Fig. 4.11).
In figures showing results of flexural models, the fiber stress of the modelled profiles relative to the
distance from the plate edge, includes indications for the onset of yielding at * = 1Gpa. For most of
the profiles the bending stress generated in the flexed thin elastic plate is close or even greater than the
normal rock strength in the place where the EC reaches its maximum topographic elevation (Fig. 4.11).
Probably this result suggests that a model with a broken plate is more appropriate to represent the
flexural deflection produced by the present-day topography of the EC. Additionally except for the most
northern profiles the calculated deflection in the Middle MV is a little in excess of the observed
subsidence. If the lithosphere in the western part of the model were thinner than in the eastern part,
deflection would have given a better fit. This suggests that probably an initial subduction of the LLA
lithosphere is occurring below the EC and the Middle MV. From structural analysis of a regional
balanced cross-section and comparison with analogue model experiments, Colleta et al. (1990) have
suggested a possible continental subduction under the EC. These authors did not reach conclusions
about the direction of such a possible subduction. Flexural modelling results support this hypothesis.
However it is difficult to constrain the subduction direction. I also tested with modelling the hypothesis
of a broken plate under the EC.

6.3. FLEXURAL SUBSIDENCE MODELS WITH A BROKEN PLATE UNDER THE


EASTERN CORDILLERA

6.3.1. Methods

The possibility of a broken plate under the EC as discussed above was also tested through
flexural models assuming a broken plate with a free end under the Cordillera, using an additional
feature of the Program Cobra not mentioned before. This is the possibility to model the lithosphere as a
thin elastic broken plate to simulate plate-subduction (Zoetemeijer, 1993). The program allows
applying

155
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes
forces or bending moments to the free end of the thin elastic broken plate model. However, I did not
apply forces or bending moments at the end of the plate.

SECTION 1 SECTION 2
anomaly ( mgal )

anomaly ( mgal )
BOUGUER ANOMALY BOUGUER ANOMALY
0. 0.
ref.level : 0. m
-70. -70. ref.level : 0. m

-140. -140.

500. 1000. 0. 500. 1000. 1500.


distance (km) distance (km)
FLEXURE FLEXURE
deflection (m)

deflection (m)
5000. 5000.

0. 0.

-5000.
-5000.

-10000.
500. 1000. 0. 500. 1000. 1500.
distance (km) distance (km)

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


80. 80.
eet (km)

eet (km)

60. 60.
40. 40.
20. 20.
0. 0.
500. 1000. 0. 500. 1000. 1500.
distance (km) distance (km)

FIBER STRESS FIBER STRESS


stress (GPa)

1. 1.
stress (GPa)

0. 0.

-1. -1.

500. 1000. 0. 500. 1000. 1500.


distance (km) distance (km)

Figure 4.11: 2D flexural models produced by the present day topographic load of the Eastern Cordillera only. Top
panel: Dots represent observed Bouguer anomaly and continuous line calculated Bouguer anomaly in mgals. Second
panel: basement deflection and topography in meters. Dots represent the observed thickness of Middle Miocene to
Recent deposits in meters. Third panel: effective elastic thickness in km. Bottom panel: Calculated bending (fiber)
stress in Gpa. Location of each section in Figure 4.1.

6.3.2. Results

Results of these models are presented in Figure 4.12. The fit obtained for the eastern side of the
model was excellent and very similar to the previous models in which the elastic plate under the EC
was continuous. For the western part of the model, except for the most northern two sections, the fit
between the flexural deflection and the thickness of Neogene sediments of the MV is also very good
and better than previous models (compare Figs. 4.11 and 4.12). This supports the hypothesis of
mantle lithosphere subduction under the EC.

156
Chapter 4

SECTION 3 SECTION 4
anomaly ( mgal )

anomaly ( mgal )
BOUGUER ANOMALY
BOUGUER ANOMALY
0.
0.
-80. ref.level : 0. m
-70. ref.level : 0. m

-140. -160.

500. 1000. 1500.


0. 500. 1000.
distance (km) distance (km)
FLEXURE
FLEXURE
5000.

deflection (m)
deflection (m)

5000.

0.
0.

-5000.
-5000.

500. 1000. 1500.


0. 500. 1000.
distance (km) distance (km)

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


80. 80.
eet (km)

eet (km)

60. 60.
40. 40.
20. 20.
0. 0.
0. 500. 1000. 500. 1000. 1500.
distance (km) distance (km)
FIBER STRESS
FIBER STRESS
stress (GPa)
stress (GPa)

1.
1.
0.
0.
-1.
-1.
500. 1000. 1500.
0. 500. 1000.
distance (km) distance (km)

Figure 4.11: Continued

6.4. LATERAL DISTRIBUTION OF EFFECTIVE ELASTIC THICKNESS AT PRESENT

The estimated effective elastic thickness, reaching a maximum value of 50-55 km in the eastern
LLA basin is shown on Figure 4.13. In comparison the effective elastic thickness in the EC shows
values of 25 km or less. In the area of the Central Cordillera results indicate low elastic thickness
of 5 km. Values of estimated effective elastic thickness of less than 10 km occur in the northern and
southern parts of the EC, where major strike-slip faults such as the Santa Marta-Bucaramanga, Bocono
and Altamira faults are located. These results will be interpreted in terms of lithosphere strength.

157
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes

SECTION 5 SECTION 6

anomaly ( mgal )
anomaly ( mgal )

BOUGUER ANOMALY BOUGUER ANOMALY


0. 0.
ref.level : 0. m ref.level : 0. m
-90. -80.

-180. -160.

500. 1000. 0. 500. 1000.


distance (km) distance (km)

FLEXURE FLEXURE
5000. 5000.
deflection (m)

deflection (m)
0. 0.

-5000. -5000.

500. 1000. 0. 500. 1000.


distance (km) distance (km)

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


80. 80.
eet (km)

eet (km)

60. 60.
40. 40.
20. 20.
0. 0.
500. 1000. 0. 500. 1000.
distance (km) distance (km)

FIBER STRESS FIBER STRESS


stress (GPa)

stress (GPa)

1. 1.

0. 0.

-1. -1.
0.
500. 1000. 500. 1000.
distance (km) distance (km)

Figure 4.11: Continued

7. FLEXURAL SUBSIDENCE PRODUCED BY GRADUAL SURFACE-UPLIFT OF THE


EASTERN CORDILLERA DURING NEOGENE TIME

Foreland basins often contain a valuable record of lithosphere processes in their sedimentary
sequences. Especially because of the continual flux of sediments, foreland basins are excellent
recorders of changes in geodynamic processes, such as changes in lithosphere strength and mountain
building processes (Price, 1973; Beaumont, 1981). The geometry of the internal stratigraphy of the
Neogene sedimentary fill during several time intervals recorded the progressive lithospheric flexural
deflection as a response to the progressive growth of the topographic load during surface-uplift of the
EC. I modelled the evolution of lithosphere flexure due to a growing topography of the EC along a 2D
regional section using several stratigraphic markers interpreted from a regional seismic traverse of the

158
Chapter 4

SECTION 7 SECTION 8
anomaly ( mgal )

anomaly ( mgal )
BOUGUER ANOMALY BOUGUER ANOMALY
0. 0.
ref.level : 0. m
ref.level : 0. m
-80. -70.

-160. -140.

0. 500. 1000. 1500. 200. 400. 600. 800. 1000. 1200.


distance (km) distance (km)
FLEXURE FLEXURE

deflection (m)
5000. 5000.
deflection (m)

0. 0.

-5000. -5000.

0. 500. 1000. 1500. 200. 400. 600. 800. 1000. 1200.


distance (km) distance (km)

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


80. 80.
eet (km)

eet (km)

60. 60.
40. 40.
20. 20.
0. 0.
0. 500. 1000. 1500. 200. 400. 600. 800. 1000. 1200.
distance (km) distance (km)
FIBER STRESS FIBER STRESS
stress (GPa)

stress (GPa)

1. 1.

0. 0.

-1. -1.

0. 500. 1000. 1500. 200. 400. 600. 800. 1000. 1200.


distance (km) distance (km)

Figure 4.11: Continued

LLA calibrated with well data (Fig. 4.6). The evolution of the growing topography and lithosphere
strength through Neogene time can be outlined from this model.

7.1 DATA

The section is located approximately close to section 6 (Fig 4.1). Fig 4.6 shows a line drawing
interpretation from the regional seismic traverse crossing the LLA basin. I identified the Leon Fm and
five stratigraphic intervals within the Guayabo Fm. I measured the two-way travel time for each of the
marker reflectors and using depth information for several wells on the section (Table 4.3), I calculated
depth of each stratigraphic marker for several points along the section. Taking into account that the
Leon Fm was deposited under shallow marine conditions I applied also a water depth load as indicated
in Table 4.4. For the continental facies of the Guayabo Fm I did not apply any water load.

159
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes

anomaly ( mgal ) SECTION 9 SECTION 10

anomaly ( mgal )
BOUGUER ANOMALY BOUGUER ANOMALY
0. 0.
ref.level : 0. m ref.level : 0. m
-80. -90.
-160. -180.

200. 400. 600. 800. 1000. 200. 400. 600. 800. 1000.
distance (km) distance (km)
FLEXURE FLEXURE
deflection (m)

5000. 5000.

deflection (m)
0. 0.

-5000. -5000.

-10000.
200. 400. 600. 800. 1000. 200. 400. 600. 800. 1000.
distance (km) distance (km)

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


eet (km)

80. 80.
eet (km)
60. 60.
40. 40.
20. 20.
0. 0.
200. 400. 600. 800. 1000. 200. 400. 600. 800. 1000.
distance (km)
distance (km)
stress (GPa)

stress (GPa)

FIBER STRESS FIBER STRESS


1. 1.

0. 0.

-1. -1.

200. 400. 600. 800. 1000. 200. 400. 600. 800. 1000.
distance (km) distance (km)

Figure 4.11: Continued

Unfortunately the internal stratigraphy of the Neogene sedimentary fill has not been dated and
age determinations of stratigraphic markers interpreted from the seismic are absent. On the assumption
of a constant sedimentation rate during Neogene, I estimated the age of these stratigraphic markers in
order to try to compare them with the surface-uplift history of the Sabana de Bogot area. Such a
history has been recorded by the Neogene sedimentary and palynological record (e.g. Helmens, 1990,
Andriessen et al., 1993).

7.2. METHOD

For each of the stratigraphic markers I modelled the flexural deflection of the lithosphere
necessary to fit the thickness accumulated between the bottom of the Leon Fm and each one of the
selected stratigraphic markers. I applied the same principles and program as the previous modelling. I
systematically changed the topography and the effective elastic thickness until an acceptable fit was
obtained between the calculated flexural deflection and the thickness of each stratigraphic interval.

160
Chapter 4

anomaly ( mgal )
SECTION 11 SECTION 12
anomaly ( mgal )

BOUGUER ANOMALY BOUGUER ANOMALY


0. 0.

ref.level : 0. m -80. ref.level : 0. m


-90.

-180. -160.

200. 400. 600. 800. 1000. 200. 400. 600. 800. 1000.
distance (km) distance (km)
FLEXURE FLEXURE
5000. 5000.

deflection (m)
deflection (m)

0. 0.

-5000. -5000.

-10000.
200. 400. 600. 800. 1000. 200. 400. 600. 800. 1000.
distance (km) distance (km)

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


80. 80.
eet (km)
eet (km)

60. 60.
40. 40.
20. 20.
0. 0.
200. 400. 600. 800. 1000. 200. 400. 600. 800. 1000.
distance (km) distance (km)
FIBER STRESS FIBER STRESS
stress (GPa)
stress (GPa)

1. 1.

0. 0.

-1. -1.

200. 400. 600. 800. 1000. 200. 400. 600. 800. 1000.
distance (km) distance (km)

Figure 4.11: Continued

7.3. RESULTS

Figure 4.14 shows the modelled results and Table 4.4 contains the effective elastic thickness
and topography obtained for the best fit of each model. In general the fit between calculated flexural
deflection and thickness of sediments filling the LLA basin is good. A comparison of the model results
for different stratigraphic markers allows to outline the evolution of the growing topography of the EC
through Neogene time, and the lithosphere strength evolution revealed by temporal changes of the
effective elastic thickness. I tried to correlate such a history to the rock-uplift history of the EC as
revealed by the stratigraphic and palynological record of the Sabana de Bogot area. In order to do such
a correlation I estimated ages for the stratigraphic markers of the Guayabo Fm assuming constant
sedimentation rates.

161
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes

SECTION 13

anomaly ( mgal )
BOUGUER ANOMALY
0.
ref.level : 0. m
-100.

-200.

200. 400. 600. 800.


distance (km)
FLEXURE
deflection (m)

5000.

0.

-5000.

-10000.
200. 400. 600. 800.
distance (km)
EFFECTIVE ELASTIC THICKNESS
80.
eet (km)

60.
40.
20.
0.
200. 400. 600. 800.
distance (km)

FIBER STRESS
stress (GPa)

1.

0.

-1.

200. 400. 600. 800.


distance (km)

Figure 4.11: Continued

8. UPLIFT EVOLUTION FROM FLEXURAL MODELING, FISSION TRACK AND


GEOLOGICAL DATA

8.1. SURFACE-UPLIFT EVOLUTION OF THE EASTERN CORDILLERA AS INFERRED


FROM FLEXURAL MODELLING

In the LLA and MV, Neogene flexural subsidence created accommodation space for sediment
accumulation. Probably the maximum topographic elevation in those basins occurs at present time
close to 200 m above sea-level.
At the end of deposition of the Leon Fm (end of Middle Miocene; Cooper et al., 1995) the
model predicts 500 m of elevation in the area of the EC (Fig. 4.14).
Later at the time of deposition of the stratigraphic marker 1 of the lower part of the Guayabo
Fm (Late Miocene?), the model predicts a surface-uplift of 2000 m in the western flank of the former
extensional basin and 2500 m in its eastern flank (Fig. 4.14). In the middle axial Tunja region of the

162
Chapter 4
EC, where the Palaeogene sedimentary record has been partially preserved, I assumed a topography
similar to that interpreted for the palynological record of the Late Miocene Marichuela Fm (Helmens,
1990; Andriessen et al., 1993), with an elevation not greater than 500 m (Fig. 4.14).

Dist
Reflector Reflector Reflector Reflector
from Top Len Fm.
Well Guayabo 1 Guayabo 2 Guayabo 3 Guayabo 4
Romeral
(km) (sec) (m) (sec) (m) (sec) (m) (sec) (m) (sec) (m)
216.40 2.00 4088.86 1.46 1933.07 1.07 1052.84 0.73 591.91 0.42 316.31
221.40 1.97 3933.70 1.44 1876.19 1.06 1035.87 0.70 560.76 0.41 308.67
226.40 1.84 3313.40 1.38 1713.88 1.02 970.33 0.68 540.65 0.39 293.52
La Mara-1 229.50 1.84 3453.38 1.38 1713.88 1.03 986.36 0.68 540.65 0.40 301.07
La Cabaa-1 233.75 1.80 2926.08 1.33 1587.85 1.03 986.36 0.70 560.76 0.41 308.67
235.85 1.78 3054.39 1.31 1539.70 1.01 954.52 0.69 550.64 0.41 308.67
240.85 1.65 2548.35 1.31 1539.70 0.97 893.49 0.69 545.63 0.40 301.07
La Gloria-1 243.35 1.62 2441.77 1.29 1492.80 0.97 893.49 0.68 540.65 0.39 293.52
248.35 1.60 2372.75 1.29 1492.80 0.97 893.49 0.67 530.79 0.39 293.52
253.35 1.54 2175.19 1.28 1469.81 0.98 908.42 0.68 540.65 0.39 293.52
254.20 1.52 2112.43 1.27 1447.13 0.98 908.42 0.68 540.65 0.39 293.52
Centauro-1 259.20 1.49 2176.27 1.23 1359.38 0.98 908.42 0.68 540.65 0.40 301.07
264.20 1.48 1991.39 1.21 1317.26 0.96 878.77 0.68 540.65 0.39 293.52
269.20 1.44 1876.19 1.18 1256.20 0.93 835.85 0.67 530.79 0.39 293.52
274.20 1.42 1820.71 1.14 1178.61 0.92 821.95 0.63 492.53 0.40 301.07
277.80 1.39 1740.08 1.13 1159.88 0.91 808.25 0.62 483.25 0.38 286.00
La Punta-1 279.40 1.35 1706.88 1.11 1123.19 0.90 794.74 0.62 483.25 0.37 278.52
284.40 1.31 1539.70 1.09 1087.52 0.90 794.74 0.62 483.25 0.35 263.66
289.40 1.29 1492.80 1.06 1035.87 0.88 768.29 0.62 483.25 0.34 256.27
294.40 1.28 1469.81 1.05 1019.13 0.86 742.58 0.61 474.07 0.32 241.56
299.40 1.23 1359.38 1.04 1002.63 0.85 730.00 0.60 465.01 0.32 241.56
Guarimena-1 301.05 1.23 1319.78 1.03 986.36 0.84 717.60 0.59 456.04 0.32 241.56
306.05 1.24 1380.88 1.03 986.36 0.83 705.36 0.58 447.17 0.31 234.24
311.05 1.22 1338.18 1.03 986.36 0.82 693.30 0.57 438.40 0.31 234.24
Morichito-1 311.85 1.21 1150.62 1.03 986.36 0.81 681.41 0.55 421.14 0.30 226.93
315.55 1.19 1276.27 1.03 986.36 0.82 693.30 0.53 404.22 0.31 234.24
328.55 1.07 1052.84 0.95 864.26 0.79 658.11 0.50 379.44 0.29 219.63
333.55 1.02 970.33 0.91 808.25 0.74 602.58 0.49 371.33 0.27 205.06
338.70 0.99 923.57 0.87 755.35 0.70 560.76 0.44 331.75 0.23 175.94
343.70 0.92 821.95 0.81 681.41 0.65 511.42 0.39 293.52 0.19 146.68
Cao Barulia-1 348.70 0.90 853.44 0.80 669.67 0.64 501.92 0.38 286.00 0.18 139.33
353.70 0.86 742.58 0.78 646.69 0.62 483.25 0.35 263.66 0.17 131.95
355.90 0.86 742.58 0.78 646.69 0.62 483.25 0.35 263.66 0.17 131.95

Table 4.3. Two way travel time and depth of the reflectors selected from seismic lines along section B. Guayabo 5
reflector is not included since represent the present day earth surface

163
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes

Distance from Top Fm Len Guayabo 1 Guayabo 2 Guayabo 3 Guayabo 4 Guayabo 5


Romeral (end Middle (Late (Early (Middle (Late (present
Region / sub- Miocene) Miocene?) Pliocene?) Pliocene?) Pliocene?) day)
basin
top EET pwd top EET top EET top EET top EET top EET
(km)
(m) (km) (m) (m) (km) (m) (km) (m) (km) (m) (km) (m) (km)
Middle
Magdalena 397 0 10 0 0 15 0 17 0 23 0 24 pdt 25
Valley
" " " 397.2 0 10 0 0 15 0 17 0 20 0 24 " 25
" " " 420 0 10 0 0 15 0 17 0 int 0 24 " 25
W flank
Eastern 452.5 int 10 int int 15 1500 17 2000 18 2000 24 " 25
Cordillera
" " " 460 int 10 int int 15 int 17 int 18 1500 24 " 25
" " " 462.5 500 10 int 2000 15 int 17 int 18 1500 24 " 25
" " " 466 500 10 int int 15 int 17 int 18 1500 24 " 25
Tunja Axial
515 500 10 int 500 15 500 17 1300 18 2200 24 " 25
region
" " " 520 500 10 int int 15 500 17 int 18 2200 24 " 25
" " " 530 500 10 int int int 500 17 1300 int 2200 24 " int
" " " 550 500 10 int int int 3000 17 3300 int 3000 24 " int
E flank
Eastern 568.5 500 10 int int int 3000 17 3300 int 3000 24 " int
Cordillera
" " " 588.5 500 10 80 2500 int 3200 17 3300 int int 24 " int
" " " 590 int 10 int int int Int 17 3300 int 3300 24 " int
Llanos basin 610 0 10 int 0 int 0 17 0 int 0 24 " int
" " " 615.5 0 int 0 int 0 17 0 int 0 24 " int
" " " 620 0 50 int 0 int 0 17 0 int 0 24 " int
" " " 650 0 50 60 0 int 0 17 0 17 0 24 " int
" " " 660 0 50 int 0 int 0 int 0 20 0 24 " int
" " " 669 0 50 int 0 int 0 int 0 int 0 50 " int
" " " 670 0 50 int 0 int 0 int 0 50 0 50 " int
" " " 700 0 50 0 0 30 0 int 0 50 0 50 " int
" " " 730 0 50 0 0 int 0 50 0 50 0 50 " int
" " " 900 0 50 0 0 50 0 50 0 50 0 50 " 55
" " " 1200 0 50 0 0 50 0 50 0 50 0 50 " 55
EET: effective elastic thickness; top: palaeotopograhy; pwd: palaeowater depth, for the other stratigraphic markers
palaeowater depth is zero; pdt: present day topography; int: linearly interpolated value depending on the distance
from the Romeral suture

Table 4.4: Evolution of the topography of the Eastern Cordillera along a regional section according to the 2D
flexural models for different reflectors in the Llanos Basinsrepresenting different times during Neogene.

164
Chapter 4

anomaly ( mgal ) SECTION 2 SECTION 3

anomaly ( mgal )
BOUGUER ANOMALY BOUGUER ANOMALY
0. ref.level : 0. m 0. ref.level : 0. m

-120. -80.

-240. -160.

300. 400. 500. 600. 250. 300. 350. 400. 450. 500. 550.
distance (km) distance (km)

FLEXURE FLEXURE
5000.
deflection (m)

deflection (m)
5000.

0. 0.

-5000.
-5000.
-10000.
-10000.
300. 400. 500. 600. 250. 300. 350. 400. 450. 500. 550.
distance (km) distance (km)
EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS
80. 80.
eet (km)

eet (km)
60. 60.
40. 40.
20. 20.
0. 0.
300. 400. 500. 600. 250. 300. 350. 400. 450. 500. 550.
distance (km) distance (km)
FIBER STRESS FIBER STRESS
stress (GPa)

stress (GPa)

1. 1.

0. 0.

-1. -1.

300. 400. 500. 600. 250. 300. 350. 400. 450. 500. 550.
distance (km) distance (km)

Figure 4.12: 2D flexural models of the western part of the Eastern Cordillera assuming a broken plate under the
Eastern Cordillera. Top panel: Dots represent observed Bouguer anomaly and continuous line calculated Bouguer
anomaly in mgals. Second panel: basement deflection and topography in meters. Dots represent the observed
thickness of Middle Miocene to Recent deposits in meters. Third panel: effective elastic thickness in km. Bottom
panel: Calculated bending (fiber) stress in Gpa. Location of each section in Figure 4.1.

Calculated bending stress started to exceed normal rock strength since the time of deposition of this
stratigraphic marker (Fig. 4.14). If some subducting lithosphere exists under the EC probably, it began
to subduct during deposition of stratigraphic marker 1.
Later during deposition of the stratigraphic marker 2, between the lower and middle parts of the
Guayabo Fm a palaeo-topography of maximum 1500 m and 3200 m respectively in the western and
eastern flanks of the palaeo-EC fits the observed subsidence (Fig. 4.14). In the middle axial Tunja
region the model requires only 500 m of palaeo-topography (Fig. 4.14) as interpreted by reference to
the elevation of the Sabana de Bogot area during deposition of the Tequendama member of the Tilata
Fm during Early Pliocene (Helmens, 1990; Andriessen et al., 1993).
At the time of deposition of the stratigraphic marker 3, between the middle and upper part of
the Guayabo Fm, the model requires a palaeo-topography varying between 1300 and 2000 m in the
western flank and 3300 m in the eastern flank of the palaeo-EC (Fig. 4.14). In the axial Tunja region a

165
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes

anomaly ( mgal ) SECTION 4 SECTION 5

anomaly ( mgal )
BOUGUER ANOMALY BOUGUER ANOMALY
0. ref.level : 0. m 0. ref.level : 0. m

-110. -120.

-220. -240.

350. 400. 450. 500. 550. 600. 650. 300. 350. 400. 450. 500.
distance (km) distance (km)
FLEXURE FLEXURE
5000. 5000.
deflection (m)

deflection (m)
0. 0.

-5000. -5000.

-10000. -10000.

350. 400. 450. 500. 550. 600. 650. 300. 350. 400. 450. 500.
distance (km) distance (km)

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


eet (km)

80. eet (km) 80.


60. 60.
40. 40.
20. 20.
0. 0.
350. 400. 450. 500. 550. 600. 650. 300. 350. 400. 450. 500.
distance (km) distance (km)
FIBER STRESS FIBER STRESS
stress (GPa)

stress (GPa)

1. 1.

0. 0.

-1. -1.

350. 400. 450. 500. 550. 600. 650. 300. 350. 400. 450. 500.
distance (km) distance (km)

Figure 4.12: Continued

palaeo-topography of 1300 m fits the observed subsidence (Fig. 4.14). This palaeo-elevation is similar
to the palaeo-elevation between 1000 and 1500 m interpreted from the palynological record of the early
Late Pliocene Tibagota member of the Tilata Fm in the Sabana de Bogot area (Helmens, 1990;
Andriessen et al., 1993).
Model results calculated at the time of deposition of stratigraphic marker 4, in the upper part of
the Guayabo Fm suggest 1500 to 2200 m of topography in the western flank of the Cordillera, 2200 to
3300 m in the eastern flank, and 2200 m of topography in the axial Tunja region (Fig. 4.14). 2200 m of
palaeo-elevation also was interpreted from the palynological record of the Late Pliocene Guasca
Member of the upper Tilata Fm in the Sabana de Bogot region (Helmens, 1990; Andriessen et al.,
1993).
Finally the present day topography along the section with elevations between 1000 to 3000 m
in the western and eastern flanks of the EC and between 2500 to 3000 in the Tunja axial region fits the
Neogene thickness up to the present day surface of the LLA basin (Fig. 4.14). Probably evidence of
small rock-uplift and erosion in some places of the MV and LLA basin (e.g. Fig. 4.7), evidenced by
fluvial terraces, are related to regional isostatic flexural rebound produced by recent erosion of the
mountain range.

166
Chapter 4

SECTI0N 6 SECTION 7

anomaly ( mgal )
anomaly ( mgal )

BOUGUER ANOMALY BOUGUER ANOMALY


0. 0. ref.level : 0. m
ref.level : 0. m
-100. -90.

-200. -180.

250. 300. 350. 400. 450. 500. 550. 350. 400. 450. 500. 550.
distance (km) distance (km)
deflection (m)

FLEXURE FLEXURE

deflection (m)
5000. 5000.

0. 0.

-5000. -5000.

-10000. -10000.

250. 300. 350. 400. 450. 500. 550. 350. 400. 450. 500. 550.
distance (km) distance (km)

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


80. 80.
eet (km)

eet (km)

60. 60.
40. 40.
20. 20.
0. 0.
250. 300. 350. 400. 450. 500. 550. 350. 400. 450. 500. 550.
distance (km) distance (km)
FIBER STRESS FIBER STRESS
stress (GPa)

stress (GPa)

1. 1.

0. 0.

-1. -1.

250. 300. 350. 400. 450. 500. 550. 350. 400. 450. 500. 550.
distance (km) distance (km)

Figure 4.12: Continued

8.2. COMPARISON OF THE UPLIFT AND EXHUMATION EVOLUTION INFERRED FROM


FISSION TRACK AND GEOLOGICAL DATA

In Chapter 3, fission track data and other evidence of deformation/uplift-exhumation have been
discussed. According to these data, during Neogene time pronounced exhumation occurred at all
localities of the northern Andes. In the Mrida Andes an exhumation of approximately 2300 m ocurred
during the Pliocene (Kohn et al., 1984; Shagam et al., 1984). In the Sabana de Bogot (axial zone of
the EC) the fluvial-lacustrine sediment record registers major tectonic surface-uplift for the period
between 5 and 3 Ma (Helmens, 1990; Andriessen et al., 1993). In the Garzn Massif, Van der Wiel
(1991) has shown that between 12 Ma and the present time the massif was exhumated approximately
6500 m. After the first Neogene (12 Ma) exhumation pulse a second pulse occurred 6.4 Ma ago. The
rate of very fast exhumation must have exceeded the rate of thermal diffusion, producing thermal
updoming of the isotherms, and the ages should be interpreted as cooling following exhumation
(Shagam et al., 1984; Van der Wiel, 1991). The most complete data from these three localities indicate
rapid regional rock-uplift, which agrees with the model results.

167
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes

anomaly ( mgal ) SECTION 8 SECTION 9

anomaly ( mgal )
BOUGUER ANOMALY BOUGUER ANOMALY
0. ref.level : 0. m 0. ref.level : 0. m

-110. -110.

-220. -220.

400. 450. 500. 550. 250. 300. 350. 400.


distance (km) distance (km)
FLEXURE FLEXURE
deflection (m)

deflection (m)
5000. 5000.

0. 0.

-5000. -5000.

-10000. -10000.

400. 450. 500. 550. 250. 300. 350. 400.


distance (km) distance (km)
EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS
80. 80.
eet (km)

eet (km)
60. 60.
40. 40.
20. 20.
0. 0.
400. 450. 500. 550. 250. 300. 350. 400.
distance (km) distance (km)
stress (GPa)

FIBER STRESS
stress (GPa)

FIBER STRESS
1.
1.
0.
0.
-1.
-1.
400. 450. 500. 550.
250. 300. 350. 400.
distance (km)
distance (km)

Figure 4.12: Continued

Fission track ages and other geologic evidence indicate initiation of Andean rock-uplift and exhumation
of blocks at the end of Cretaceous-Paleocene with local phases of rock-uplift during Eocene to Miocene
time, terminating with regional rock-uplift and exhumation in all blocks in the Pliocene-Pleistocene
(Shagam et al., 1984). These authors interpreted this to result from uncoupled crustal blocks being
uplifted in response to local stress until regional compression led to interlocking of the blocks and their
simultaneous rock-uplift during the Pliocene-Pleistocene time.
Model results (Fig. 4.14) agree with these observations. Probably during the Neogene the
normal fault systems that delimited the Mesozoic extensional basin were completely inverted in a
compressional/transpressional stress field. The most uplifted areas were close to these fault systems.
The eastern flank of the Cordillera illustrates these processes. The Chmeza thrust fault system, which
was one the eastern extensional Mesozoic master faults limiting the basin, started an episode of total
inversion 25 Ma ago (Hossack et al., 1999). Fission track data reported by Hossack et al. (1999)
indicates that the sedimentary fill of the former extensional basin close to the fault began to uplift and
exhumate at that time. Younger apatite data (15 to 3 Ma) from the more internal parts of the Cordillera
suggest later exhumation (Hossack et al., 1999). Finally the surface of the central axial Sabana de
Bogot region was uplifted during Pliocene time (5 to 3 Ma; Andriessen et al., 1993). However, better
data are needed to refine models and unravel more precisely the Neogene uplift history.

168
Chapter 4

anomaly ( mgal ) SECTION 10 SECTION 11

anomaly ( mgal )
BOUGUER ANOMALY BOUGUER ANOMALY
0. ref.level : 0. m 0. ref.level : 0. m

-120. -120.

-240. -240.

300. 350. 400. 450. 500. 350. 400. 450.


distance (km) distance (km)
FLEXURE FLEXURE
deflection (m)

5000.

deflection (m)
5000.

0. 0.

-5000. -5000.

-10000. -10000.

400.
300. 350. 400. 450. 500. 350. 450.
distance (km) distance (km)

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


eet (km)

80. 80.
eet (km)

60. 60.
40. 40.
20. 20.
0. 0.
300. 350. 400. 450. 500. 350. 400. 450.
distance (km) distance (km)
stress (GPa)

FIBER STRESS FIBER STRESS


stress (GPa)

1. 1.

0. 0.

-1. -1.

300. 350. 400. 450. 500. 350. 400. 450.


distance (km) distance (km)

Figure 4.12: Continued

In the Middle MV Middle and Upper Miocene, fluvial sediments were deposited by a northward-
directed fluvial system, which was fed by the Eastern and Central cordilleras (Figs. 1.1 and 4.4).
Sedimentation was contemporaneous with Central Cordillera volcanism between 10.92 and 6.2 Ma
(Gmez et al., 1999). However, the dominantly eastward palaeocurrent directions measured in the
Palaeogene (La Paz Esmeraldas, Mugrosa and Colorado Fms of the Middle MV, e.g. Rubiano, 1998;
Gualanday Gp of the Upper MV, e.g. Anderson, 1970, 1972) and Neogene (Real Fm of the Middle
MV, Rubiano, 1998; Honda, Neiva and Gigante Fms of the Upper MV, e.g. Wellman, 1970; Howe,
1974; Van der Wiel, 1991) sedimentary record of the MV seems to be in contradiction with the
hypothesis of gradual surface-uplift of the EC. Van der Wiel (1991) found eastward palaeocurrent
directions in the Neogene Honda Fm of the Upper MV and fission track data indicating simultaneous
rock-uplift and exhumation of the EC Garzon Massif (Fig. 2.1). She proposed the following possible
explanations for this apparent contradiction: (1) The Garzn Massif is an eastward tilted block, with
most of the present-day drainage flow eastward toward the continent interior and only minor amount of
sediment flow toward the western side in the Upper MV; (2) Within the EC Garzon Massif there are
small sedimentary basins that trapped sediments inhibiting their westward distribution in the Upper

169
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes
MV; (3) The volume of sediments and volcanic products from the Central Cordillera was greater than
the volume of sediments from the rising EC. The MV surface was tilted toward the east; the Proto-
Magdalena River flowed close to the eastern border of the basin capturing all the eastward flowing
drainage systems. Westward drainage was poorly developed and possibly only occurred in the
easternmost part of the MV in areas of poor exposure (Van der Wiel, 1991). Probably these
explanations may also be extended to the whole MV. The axial Sabana de Bogot region (Figs. 2.1 and
4.4) was a sedimentary basin until recent time (Helmens, 1990; Andriesen et al. 1993) trapping
sediments and preventing their distribution on the MV. The eastward thickening of the sedimentary
record for several Tertiary time intervals in the MV indicates a progressive tilting of the depositional
surface of the MV.

SECTION 12
anomaly ( mgal )

BOUGUER ANOMALY
0. ref.level : 0. m

-110.

-220.

300. 350. 400.


distance (km)
FLEXURE
deflection (m)

5000.

0.

-5000.

-10000.

300. 350. 400.


distance (km)
EFFECTIVE ELASTIC THICKNESS
80.
eet (km)

60.
40.
20.
0.
300. 350. 400.
distance (km)

FIBER STRESS
stress (GPa)

1.
-1.
0.

300. 350. 400.


distance (km)

Figure 4.12: Continued

170
Chapter 4
Serrania Maracaibo
LITHOSPHERE de Perija Basin s
de
EFFECTIVE An

Lucas
1400
Cucuta a Barinas
ELASTIC
e rid
THICKNESS M Basin

de San
(km)

Serrania
1300 Bucaramanga

5
Arauca
20

1200 Medellin
a
ller
rd i
Co

20

Tunja

30
1100
Yopal

50
10
10

Manizales

Bogota
1000 Ibague

Villavicencio

900
la
ntr
Ce

Guyana Shield
Neiva
La Macarena
Serrania de

50
800 San Jose del -0 100 200 km
Guaviare

800 900 1000 1100 1200 1300 1400 1500

Figure 4.13: Map of calculated effective elastic thickness. Note location of important strike slip faults associated to
pre-Mesozoic outcrops (shaded areas: Santander-Floresta and Garzon-Quetame massifs) and location of possible
mantle lithosphere subduction under the Eastern Cordillera.

9. DISCUSSION

9.1. NEOGENE TECTONIC HISTORY AND PLATE-TECTONICS

During Neogene the Mesozoic extensional basin was completely inverted (Colletta et al., 1990;
Dengo and Covey; 1993; Cooper et al., 1995; Casero et al., 1995, 1997;). Upthrust of the basement
blocks represented by the Garzon Massif and the Sierra de La Macarena (Fig. 2.1) also occurred during
Neogene (Van der Wiel, 1991). Inversion of extensional basins and upthrusts of basement blocks
occurred by reactivation of pre-existing crustal discontinuities in response to
compressional/transpressional stresses on the NW margin of the South American plate. These processes
have been proposed to explain basin inversion and upthrust of basement blocks in Europe and many
other areas by Ziegler et al. (1995, 1998). According to the plate tectonic history of Colombia,
convergence rate of the South American plate relative to the Pacific plates was particularly fast during
Late Miocene and Pliocene, triggering Andean orogeny (Cooper et al., 1995). Backarc compression
resulting in inversion of rift tensional hanging-wall basins is associated with Andean-type orogens

171
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes
occurs during periods of increased convergence rates between the subducting and overriding plates
(Ziegler et al., 1998). This inversion of backarc rift basins results from acceleration of convergence
rates between the colliding plates, their increased mechanical coupling and the transmission of
compressional stresses into the backarc domain of the overriding plate (Uyeda and McCabe, 1983;
Ziegler, 1993; Ziegler et al., 1998). In addition Neogene basin inversion and Andean orogeny have
been correlated to the collision of the Choco Block (Cuna Terrane according to terminology of
Toussaint, 1995) with the northwestern margin of South America (Duque-Caro, 1990, Cooper et al.,
1995). Kerr et al. (1997) have studied the geochemistry of the volcanic rocks of the oceanic Cuna
Terrane and proposed that it represents part of an oceanic plateau. Build up of intraplate
compressional/transpressional stresses in the NW margin of South America during Neogene and
collision of the Cuna Terrane could be favoured by impeded subduction caused by the arrival of the
more buoyant oceanic plateau. Probably right-lateral transpressional deformation contributed to
Andean orogeny and allowed inversion of Mesozoic normal faults. An example of a Mesozoic normal
fault system, right-lateral transpressionally inverted during Neogene, is the Guaicramo-Chmeza fault
system (limiting the EC), evidenced by earthquake focal mechanisms solutions (Pennington, 1981).
Some right-lateral faults (e.g. Altamira Fault, Casero et al., 1995, 1997) were also active during
Neogene.

9.2. LITHOSPHERE STRENGTH EVOLUTION AS INFERRED FROM FLEXURAL


MODELLING

The parameter that characterizes the apparent strength of the lithosphere is the flexural rigidity,
which is commonly expressed through the effective elastic thickness (EET) of the lithosphere. The
distribution of strength in the lithosphere varies vertically as a result of its thermal and compositional
layered structure. However an appropriate first order description of the flexural behaviour of the
lithosphere can be formulated assuming it as a thin elastic plate (e.g. Turcotte and Schubert, 1982). The
effective elastic thickness is a useful and easy parameter to estimate that provides important
information on the mechanical and thermal state of the lithosphere. For an elastic plate the notion of the
integrated strength is quite close to that of the flexural rigidity (Burov and Diament, 1995). These
authors have shown that the effective elastic thickness of the continental lithosphere is dependent on:
(1) The thermal state/age of the lithosphere (thermal age defined as a period of time required for the
lithosphere to reach its present-day thermal state, assuming that the lithosphere was initially melted.
The thermal age controls the depth to a specific geotherm obtained from a plate cooling model
assuming that the lithosphere did not undergo thermal re-settings during this time. The thermal age
gives the age of the last large-scale thermal event (Burov and Diament, 1995); (2) The coupling or
decoupling state of the crust and mantle; (3) The thickness and proportions of the mechanically
competent crust and mantle; and (4) The local curvature of the plate, which is directly related to the
bending stress (Burov and Diament, 1995). An additional important control on lithosphere strength is
the presence of pre-existing discontinuities, which are prone to reactivate reducing strength (Van Wees,
1994; Ziegler et al., 1995, 1998). The thermal state of the lithosphere controls lithospheric strength
since temperature dependent creep controls the ductile strength of the lower crust and lower mantle
lithosphere (Ranalli, 1995). The thickness of the mechanically competent crust and the degree of
coupling or decoupling are generally controlled by composition of the upper and lower crust, total
thickness of the crust, and by the crustal geotherm. If decoupling take place, as is of common
occurrence in continental lithosphere, it permits as much as 50 % decrease of elastic thickness,
compared with elastic thickness derived for conventional thermal profiles (Burov and Diament, 1995).
The curvature of the plate depends of the rheological structure and on the distribution of external loads
applied to the plate (e.g. topography, sediment fill and plate boundary forces). Bending stresses created
by major mountain belts are large enough to cause inelastic deformation (brittle failure and ductile
flow) in the underlying plate, which in turn, leads to a 30 to 80% decrease of elastic thickness beneath
such belts and less beneath the adjacent regions. The boundary forces and moments lead to more
localized but even stronger reductions in elastic thickness (Burov and Diament, 1995).

172
Chapter 4
LEON GUAYABO 1
FLEXURE FLEXURE
deflection (m)

deflection (m)
1000. 4000.

500. 2000.
0. 0.
-500. -2000.
-1000.
-4000.
-1500.
200. 400. 600. 800. 200. 400. 600. 800.
distance (km) distance (km)

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


eet (km)

80. 80.

eet (km)
60. 60.
40. 40.
20. 20.
0. 0.
200. 400. 600. 800. 200. 400. 600. 800.
distance (km) distance (km)
FIBER STRESS FIBER STRESS
stress (GPa)

1. stress (GPa) 1.

0. 0.

-1. -1.

200. 400. 600. 800. 200. 400. 600. 800.


distance (km) distance (km)

GUAYABO 2 GUAYABO 3
FLEXURE FLEXURE
5000.
deflection (m)

deflection (m)

4000.
2000.
0. 0.
-2000.
-4000. -5000.
-6000.

200. 400. 600. 800. 200. 400. 600. 800.


distance (km) distance (km)

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


eet (km)

80. 80.
eet (km)

60. 60.
40. 40.
20. 20.
0. 0.
200. 400. 600. 800. 200. 400. 600. 800.
distance (km) distance (km)
stress (GPa)

FIBER STRESS FIBER STRESS


stress (GPa)

1. 1.

0. 0.

-1. -1.

200. 400. 600. 800. 200. 400. 600. 800.


distance (km) distance (km)
Figure 4.14: 2D flexural models produced by the gradual surface-uplift of the Eastern Cordillera during Neogene
time using data from the seismic section shown in figure 4.6. Location of the section in Figure 4.1

173
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes

GUAYABO 4 GUAYABO 5
FLEXURE FLEXURE
5000. 5000.
deflection (m)

deflection (m)
0. 0.

-5000. -5000.

200. 400. 600. 800. 0. 500. 1000.


distance (km) distance (km)

EFFECTIVE ELASTIC THICKNESS EFFECTIVE ELASTIC THICKNESS


eet (km)

eet (km)
80. 80.
60. 60.
40. 40.
20. 20.
0. 0.
200. 400. 600. 800. 0. 500. 1000.
distance (km) distance (km)

FIBER STRESS FIBER STRESS


stress (GPa)

stress (GPa)

1. 1.

0. 0.

-1. -1.

200. 400. 600. 800. 0. 500. 1000.


distance (km) distance (km)

Figure 4.14: Continued

9.2.1. Effect of lithosphere thermal age on strength

One of the major controls on the effective elastic thickness of the study area is the thermal age
of the lithosphere, which varies laterally. Based on this and the values obtained for elastic thickness
three regions can be distinguished in the study area:
Llanos Orientales (LLA) Basin. The estimated effective elastic thickness has a maximum value
of 50 to 55 km in the eastern LLA basin (Fig. 4.13). The thermal age of the eastern LLA lithosphere is
older; the last significant thermal event in this region probably was Cambrian rifting (Ecopetrol and
Beicip, 1995, mainly in the western part of the LLA area). This correlation suggests a relatively stable
strong lithosphere in this area. However in the western part of LLA basin values of effective elastic
thickness are intermediate between those of the Andean region and those of the eastern part of LLA
basin (Fig. 4.13). In the eastern part of LLA basin a strong continental lithosphere is indicated by
effective elastic thickness with a constant value of 50 km for most of the Neogene and 55 km at present
time (Table 4.4). The westernmost part of LLA basin is characterized by the transition between the
strong (EET 50 Km) and weak (increasing EET from 10 to 25 km during Neogene time, Table 4.4)
continental lithosphere.
Eastern Cordillera (EC)-Magdalena Valley (MV). The effective elastic thickness in the EC has
values of 25 km or less (Fig. 4.13) indicating a weak lithosphere. The age of the last large-scale thermal
event related to the Mesozoic rifting phases which started 248 Ma and ended 102 Ma ago. Because the
last stretching/heating event occurred during Aptian time, the estimated thermal age is about 120 20
Ma. This weak lithosphere is prone to deformation as indicated by its well-developed earthquake
activity (e.g. Pennington, 1981). Weak lithosphere in the area of the EC and the western part of the MV

174
Chapter 4
correlates with the area of the former Mesozoic extensional basins (Figs. 2.13, 2.15 and 2.17 Chapter
2). In general weak lithosphere in the Andean region probably corresponds to thermally destabilized
lithosphere. Young rifts and volcanic areas are characterized by low elastic thickness values (e.g.
Ebinger et al., 1989, in Burov and Diament, 1995) due to thermal weakening and necking of the
lithosphere. Stretched lithosphere is weak and low elastic thickness values in the EC may have
inherited this weakness. Similar results were obtained by Van der Beek and Cloetingh (1992), who
demonstrated that preceding thermo-tectonic events explain the present-day flexural response of the
lithosphere in the Betic cordilleras of Spain.
Central Cordillera (including westernmost MV). Weakest lithosphere in the area of the Central
Cordillera (EET values of 5 km, Fig. 4.13) also correlates to the recent volcanic arc in the Central
Cordillera. The thermal age of the Central Cordillera is very small since at the present-day there is
active volcanism on this mountain range. Volcanism and very low elastic thickness suggest a hot and
weak lithosphere in the area of the Central Cordillera.

9.2.2. Coupling/decoupling state of the crust-mantle, thickness and proportions of mechanically


competent crust and mantle

Burov and Diament (1995) have compared analytical and numerical solutions for the flexural
behaviour of the lithosphere for both the elastic model and a stratified yield stress envelope model with
realistic brittle-elasto-ductile rheology. In order to compare results of both models deflection of the
brittle-elasto-ductile plate was also modelled by these authors as the deflection of an equivalent elastic
plate with space and temporally variable effective elastic thickness. Quantitative relationships between
thermal age of the lithosphere, effective elastic thickness and crustal thickness showing the reduced
elastic thickness effect produced by crust-mantle decoupling obtained by Burov and Diament (1995),
are shown in Figure 4.15. These authors showed that the lithosphere is decoupled when the crustal
thickness is greater than a critical value, and coupled, when the crust is thinner than such critical value.
The critical value corresponds to the transition zone in Figure 4.15. It increases with age of the
lithosphere until it reaches an asymptotic value close to 35 km for ages greater than 750 Ma. Plotting
the thermal age, crustal thickness and obtained elastic thickness for the EC, Central Cordillera and LLA
area for different times during the Tertiary and present-day indicates a relatively young lithosphere and
crust-mantle decoupling (Fig. 4.15). Thus Tertiary pre-Andean and Andean deformation of the upper
crust of the EC was largely independent of upper mantle deformation. These zones were separated by
ductile flow in the lower crust. In areas of young lithosphere like the Alps (e.g. Okaya et al., 1996) the
elastic thickness estimates (5-30 km) are smaller that the depth of the base of the mechanically strong
upper crust, corresponding to the isotherm 200 -300 C (Burov and Diament, 1995). The similar age
and elastic thickness obtained for the Colombian EC suggests a similar picture for the EC. According to
Burov and Diament (1995) for very young lithosphere, as is the case for the Central Cordillera and the
EC during Palaeogene time, the thickness of the competent crust may be greater than, or comparable
with, the thickness of the competent mantle lithosphere. Consequently, the integrated strength of the
young plate is controlled at a large extent by the strength of the crust (Kusznir and Karner, 1985;
Burov and Diament, 1995). The mantle lithosphere perhaps does not contribute to the plate strength in
these cases. For ages larger than 100-150 Ma, the lithospheric strength is significantly controlled by the
strength of the mantle (Burov and Diament, 1995). If the thermal age of the EC lithosphere is about 120
Ma, probably mantle began to contribute to lithospheric strength only since Neogene time, as
suggestedby the obtained elastic thickness values. According to Ziegler et al. (1998) compressional
deformation restricted to crustal levels involves simple shear type detachment of the crust at the level
of the rheologically weak lower crust from the mantle lithosphere. This mechanism probably occurred
in the EC area during Palaeogene and beginning of Neogene time, as suggested by reduced EET values
at those times.

175
hp
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes

Ther mal Ag e of Litho sphe re (Ma )

0 25 0 50 0
10

70

CO UPL ING
80

90
o

1 00

1 10
4
4
20 Tr
an
s it
ion
5 8, 9 9 zo
8 ne
5
h k

3
3
62

70
7

60
e

7 50
30 1 Cent ral
o

40

Cordiller a
3 0
M

2 0

DEC

10
10

10
10
O

Llanos
EE T

2
U PL

Bas in
6
( k m)

IN G

1
40
6 7
3 8, 9
2
5

End C ret a c e ou s En d N e o g en e
E nd P al e og en e 3 Inv e rted bas in com pa rt m e nt

Figure. 4.15: Coupling/decoupling state of the lithosphere as a function of the age of the lithosphere and the Moho
depth. One can see that for each age there is a critical value of crustal thickness (transition zone curve). If the crustal
thickness is larger than this critical value, the lithosphere is practically always decoupled, resulting in low EET. In
turn, thin crust results in mechanical coupling between the crust and mantle and in high EET values. For a
lithosphere older than 750 Ma the value of the critical crustal thickness is practically constant and equal to 35 40
km (from Burov and Diament, 1995). Colombian data: numbers atached to a symbol (circles, squares and crosses)
indicate basin compartments shown in Fig 2.18. Numbers 2 to 9 refer to basin compartments the former Mesozoic
extensional basin in the area of the EC

176
Chapter 4
9.2.3. Effect of local curvature of the plate on lithosphere strength

Radius of plate curvature, which is related to the gradient of the bending stress within the plate,
is a parameter that also controls effective elastic thickness according to Burov and Diament (1995).
These
authors calculated the relationship between the radius of curvature, the thermal age and the equivalent
elastic thickness. Figure 4.16 shown their result for a case of crust-mantle decoupling. Lithospheric
flexure may be responsible for localized reductions of the effective elastic thickness that may be
between 0.1 and 0.5 orders (Burov and Diament, 1995). In their figure there are two families of curves:
grey curves represent the concave downward flexure, with compression in the uppermost crust and
mantle, while the black curves represent concave upward flexure, with extension in the uppermost crust
and mantle. The downward flexed plate generally cannot be so highly flexed as the subducting plate,
because the stresses required for the concave downward flexure are much higher than those for the
concave upward flexure. Plotting the respective values obtained for the EC (Table 4.5 and Fig. 4.16)
may explain the drastic reduction in strength (modelled assuming the end of a broken plate (elastic
thickness = 0 under the Cordillera) in the place where the continuous plate model under the Cordillera
predicts maximum curvature and maximum bending stress. Also the reduction of elastic thickness in
the Middle Magdalena Valley may be at least partially explained by this effect.

EET Bending stress EET/ Bending stress Radius


Curvature Radius curvature Log10(radius)
(km) (GPa) (km/GPa) (km)
Antiform min 15 0.60 25.00 0.625 -0.20412
Antiform max 50 0.20 250.00 6.250 0.79588
Synform min 15 1.20 12.50 0.313 -0.50515
Synform max 50 0.65 76.92 1.923 0.28400
EET: effective elastic thickness.

Table 4.5: Estimated radius of curvature of the flexed lithosphere under the Eastern Cordillera calculated from the
modelled bending stress.

EC

Log1 0 (Radius of curvature (Km))


Figure 4.16: Decoupled lithosphere (lower crust with low temperature of creep activation and/or young plates). The
decoupled model fits most of the continental data excepts few cases where the lithosphere is likely to be coupled
(Himalaya). The lithospheric flexure can dramatically decrease the effective elastic thickness of the lithosphere.
(from Burov and Diament, 1995). Colombian data: LL Llanos basin. EC: Eastern Cordillera.

177
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes
9.2.4. Effect of pre-existing discontinuities on lithosphere strength

Van Wees (1994) has shown the important role of pre-existing discontinuities is reducing
lithosphere strength. I have discussed this in Chapter 3. This effect is probably very important in the
case of the EC since the leading mechanism of orogeny has been basin inversion with compressional
reactivation of former extensional faults. The reactivation of faults also suggests an oblique orientation
of the compressional stress trajectories, which is in agreement with oblique plate convergence.
Stretched lithosphere is weak in part due to those discontinuities and low elastic thickness values in the
EC may inherit this weakness.
Additionally in the EC values of estimated effective elastic thickness less than 10 km occur in
the northern and southern parts of the EC, where major strike-slip faults such as the Santa Marta-
Bucaramanga, Bocono and Altamira faults are located. Regionally Andean deformation affected weak
lithosphere. However, the weakest lithosphere in the EC seems to be associated with strike-slip faults
(Fig. 4.13). Such faults probably are deep enough into the lithosphere to reduce its total strength.
Outcrops of pre-Mesozoic igneous and metamorphic basement rocks also are associated with strike-slip
faults and weak lithosphere (Fig. 4.13). Probably such faults also have vertical components that brought
basement rocks to surface.

9.2.5. Other effects on lithosphere strength

Elevated topography of mountain ranges creates normal stresses high enough to cause
significant flexural deformations and bending stresses on the underlying lithosphere (e.g. Burov and
Diament, 1995). Inelastic deformation below the EC is suggested by the elevated irrealistic bending
stresses greater than rock strength estimated from flexural models (Figs. 4.10 and 4.11). Burov and
Diament (1992, in Burov and Diament, 1995) have demonstrated strength reduction under mountain
ranges. The better fit obtained for a broken plate end elastic model under the EC (Fig. 4.12) than for a
continuous plate (Fig. 4.11) suggests a dramatic strength reduction (to zero) under the Cordillera and a
possible subduction under the EC as hypothesized by Colleta et al. (1990).
The presence of rhyolitic and andesitic volcanic rocks of Neogene age near Paipa and Iza
(Renzoni et al., 1967; Romero and Rincon, 1990) demonstrates Neogene magmatic activity in the EC
at least locally. Additional thermal heating from below (evidenced by the Paipa-Iza volcanic rocks) also
contributes to a reduction of lithospheric strength under the EC.
In zones of collision, the lithosphere is subjected to significant plate boundary loads (Burov and
Diament, 1995, Ziegler et al., 1998) creating significant local strength variations (Burov and Diament,
1995). Although such plate boundary forces may have existed in the flexural history of the EC
horizontal stresses may also have a significant role on the flexural history of the EC and on the
lithosphere strength (Cloetingh, 1988; Burov and Diament, 1995). However, in the present situation
models do not require plate boundary forces or subsurface loads, and for simplicity I neglected their
effect.

9.2.6. Temporal changes of lithosphere strength

In the weak EC region lithospheric strength increased during Neogene time (Table 4.4).
Cooling and lithosphere thickening after the Mesozoic rifting probably dominated this increase in
strength, although other mechanisms also participated as already discussed. Pre-existing discontinuities
also decrease lithosphere strength. The effective elastic thickness obtained in the EC region gradually
increased from 10 km at the end of Middle Miocene during deposition of the top of Leon Fm up to 25
km at present-time (Table 4.4). During Palaeogene time small EET values (<10 km) imply that
lithospheric strength was only represented by the upper crust (c.f. Cloetingh and Burov, 1996). Gradual
increase of EET values during Neogene suggests a gradual participation of the mantle lithosphere in the
total lithospheric strength (c.f. Cloetingh and Burov, 1996).

178
Chapter 4
The progressive increase in EET values in the Andean region and constant EET in the LLA region
imply a general increase of lithospheric strength. Apart of the apparent bulge migration because of
convergence of the foreland toward the mountain belt, such general increase of lithospheric strength
would produce an increase in width of the foreland basin and migration of the flexural bulge away from
the load. In general this trend is observed in the LLA basin during the whole Tertiary (Figs. 3.4, 3.10,
3.12. 4.2, and 4.4 and Table 4.4). In the LLA the Paleocene basin was narrow (<100 km) and there is
evidence in its eastern part for erosion of the Cretaceous sedimentary record probably due to erosion of
the uplifted flexural bulge. Later during Palaeogene the LLA foreland width increased to approximately
300 km at the end of deposition of Carbonera Fm (end of Early Miocene), and up to a maximum of
approximately 400 km in the north during deposition of Guayabo Fm in Neogene time (Figs. 3.4, 3.10,
3.12, 4.2, 4.4 and 4.8). During the whole Tertiary maximum width of the LLA basin is located in the
northern Arauca area and decreases toward the south suggesting a southward decrease in lithospheric
strength as also indicated by flexural models (Figs. 3.4, 3.10, 3.12, 4.2, 4.4 and Table 4.2). In the
southwestern LLA basin close to the Serrania de La Macarena there is seismic evidence of partial
erosion of the Neogene Guayabo Fm (Fig. 4.7). Seismic lines of other parts of the basin (e.g. Fig. 4.6)
do not show evidence of similar erosion, which affects only this part of the basin close to the Serrania
de La Macarena. Probably this was the result of flexural rebound induced by an erosional decrease in
topographic load in that part of the EC (Garzon-Quetame Massif, Fig. 2.1). The Serrana de La
Macarena can be regarded as a foreland upthrust block. Because in this SW part of the basin this
erosion of Guayabo Fm implies a reduction in sediment thickness and estimated subsidence the flexural
deflection, the width of the basin and EET value have been underestimated for the southernmost
modelled profile.
In the westernmost part of the LLA basin the calculated EET value is lower than the 50 km
estimated for the eastern continental lithosphere. In this region model results (Table 4) indicate that the
zone of weak lithosphere (characterized by a reduced elastic thickness 50-km) moved craton-ward
possibly as the result of heat transmission from the hot Andean region to the cold cratonic region.
The characteristic time of these thermal relaxation processes is about 50 Ma (McKenzie, 1978). An
alternative explanation may be found in the flexural behaviour of a viscoelastic plate. In such model
stress relaxation leads that a load initially isostatically compensated by a regional flexural deflection
with time to evolve to local isostatic compensation. If a thin elastic plate model is applied in this case
for several time intervals, the EET will decrease with time (Walcott, 1970; Watts et al., 1982; Garcia
Castellanos, 1998). Characteristic time of this viscous stress relaxation is 1 Ma. Taking into account the
Neogene time interval studied (16.4 Ma) the viscous stress relaxation hypothesis is the most probable
explanation. Additionally most of the thermal relaxation occurred during Palaeogene time (48.6 Ma).

9.2.7. Suggestions for further studies

Detailed stratigraphic studies of the Neogene sediments, their relations with deformation, and
fission track data are needed to accurately constraint models about the deformation and uplift history of
the EC.

10. CONCLUSIONS

Comparison of the observed Neogene tectonic subsidence with the remaining thermal
subsidence assuming tectonic quiescence indicates that the Mesozoic extensional basin was completely
inverted during Neogene time to form the EC. Increased South American plate convergence rate and
possibly collision of the oceanic plateau Cuna Terrane favoured this compressional/transpressional
deformation. Flexural subsidence produced by the topographic load of the EC explains the observed
Neogene subsidence in the LLA and MV. Irrealistic high bending stresses predicted by flexural models
under the EC and the better fit obtained by flexural models assuming a broken plate under the EC,
support the hypothesis of mantle lithosphere subduction (Colletta, et al., 1990) under the EC.
Flexural model results for different stratigraphic markers within the Neogene sedimentary
record, and comparison with available fission track data and other evidence of deformation/rock-uplift,
suggest that during Miocene time important surface-uplift (> 1000 m) occurred mainly at the margins

179
Neogene Basin Inversion History of the Eastern Cordillera, Colombian Andes
of the EC. In the middle axial Tunja-Sabana de Bogot region, where the Palaeogene sedimentary
record has been partially preserved, important surface-uplift occurred later during Pliocene time as
recorded by the exceptional palynological record of the Sabana de Bogot (Helmens, 1990; Andriessen
et al., 1993). Such uplifting history can be explained in terms of basin inversion: contractional
reactivation of Mesozoic extensional faults initiated during Palaeogene times leading to an initial slight
basin inversion; then during Neogene time the compressional deformation increased in rate and
magnitude, leading to a complete inversion of the original extensional basin. Scarce fission track data
indicate exhumation and the results of flexural models suggest that moderate (< 1000 m but mostly
200 m) and local surface-uplift of blocks, mostly associated with the master normal fault system that
delimitated the Mesozoic extensional basins, initiated during Palaeogene times (see Chapter 3).
However, during Neogene times increased surface-uplift (> 1000 m) affected the former Lower
Cretaceous depocenters adjacent to the master normal fault systems delimiting the basin. Complete
inversion of these master normal fault systems led to extrusion of the sedimentary fill of half graben
basins that now form the eastern and western flanks of the EC. Regional structural balanced cross-
section interpretations (e.g. Colletta et al., 1990; Cooper et al., 1995) suggest that maximum dip-slip
displacement and shortening occurred on the thrust faults connected to inverted Mesozoic normal faults
that now approximately delimit the Lower Cretaceous outcrops of both flanks of the EC. In these
uplifted flanks Lower Cretaceous or older exposed rocks indicate that all Late Cretaceous or younger
sediments have been eroded. However, only in the Santander, Quetame and Garzon massifs all the
sedimentary cover has been removed by erosion. Proximity of these massifs to important strike-slip
faults suggests that inversion was associated with important strike-slip motions. Preservation of the
Palaeogene sedimentary record in the axial Bogot-Tunja zone of the EC, as well as the Neogene
sedimentary record of the Sabana de Bogot area suggest that this axial region remained low during
Miocene time and was uplifted until the Pliocene.
One of the major controls on the effective elastic thickness of the study area is the thermal age
of the lithosphere. Based on the thermal age and the values obtained for elastic thickness, three regions
can be distinguished in the study area: (1) The LLA basin where effective elastic thickness has a
maximum value of 50 to 55 km and the thermal age of the eastern LLA lithosphere is Palaeozoic (this
correlation suggests a relatively stabilized strong lithosphere in this area); (2) The Eastern Cordillera-
MV, where effective elastic thickness has values of 25 km or less, indicating a weak lithosphere and the
thermal age is about 120 +-20 Ma (this weak thermal destabilized lithosphere is prone to deformation;
weakness there is inherited from the former Mesozoic extensional basins); (3) The Central Cordillera
(including westernmost MV) with EET values of 5 km and very young thermal age as indicated by a
recent volcanic arc in the Central Cordillera.
EET values lower than 10 km occur in the northern and southern parts of the Eastern
Cordillera, where major strike-slip faults such as the Santa Marta-Bucaramanga, Bocono and Altamira
faults are located. Regionally Andean deformation affected weak lithosphere; however, the weakest
lithosphere is associated with strike-slip faults. Such faults probably reach deep enough into the
lithosphere to reduce its total strength. Results of flexural modelling indicate crust-mantle decoupling.
In the Eastern Cordillera mantle began to contribute to lithospheric strength probably only since
Neogene time. Existence of pre-existing discontinuities probably significantly reduced lithosphere
strength. Maximum radius of curvature of the flexed plate under the Eastern Cordillera also reduced its
strength. In the weak Eastern Cordillera region, lithospheric strength increased during Neogene time. A
local lithospheric weakening effect in the western LLA basin can be interpreted as viscous stress
relaxation.

180
CHAPTER 5

MAP VIEW RESTAURATION OF TRANSPRESSIONAL BASIN INVERSION IN THE


EASTERN CORDILLERA

1. INTRODUCTION

In areas of compressional tectonics, the traditional method for restoring the deformation is to
use balanced cross-sections (Dahlstrom, 1969). Such restoration takes into account both folding and
faulting. The main assumption is conservation of lengths or surface areas of the layers in the cross-
section. This assumption is not valid if there are differential motions of material in directions
perpendicular to the plane of the section, such as occur in strike-slip faulting or block rotations about
vertical axes. This restriction makes the methods of doubtful use in areas of non-planar deformation
such as the Northern Andes and, in particular, the EC (Fig. 1.1), which is located in a broad zone of
deformation resulting from the interaction of the South American, Nazca and Caribbean plates (Fig.
5.1). Such interaction has fragmented the NW corner of the South American plate into a number of
micro-plates or tectonic blocks (e.g. Kellogg et al., 1985; Freymueller et al., 1993). Relative
movements of the micro-plates has resulted in deformed belts with a combination of compressional
thick-skinned (e.g. Julivert, 1970; Colletta et al., 1990), thin-skinned thrusting (Dengo and Covey,
1993; Fajardo-Pea, 1998), folding (Campbell and Burgl, 1965; Julivert, 1970; Kammer, 1999a, b) and
strike-slip faulting (Campbell, 1968; Kammer, 1993a, 1996, 1999 a,b; Acosta et al., 2000) which form
several mountain ranges or strike-slip fault zones in the Northern Andes.
The variable orientation of these deformed belts, as well as the combination of compression and
strike-slip movements, implies non-planar deformation, which cannot be, restored properly using
balanced cross-sections alone. Some algorithms and commercial computer programs have been
developed for 3-D structural restoration. Rouby (1995), Rouby et al. (1998) among others have
developed algorithms for unfolding non-cylindrical folds. They allow unfolding individual blocks.
Numerical procedures for fitting blocks in plan view, applicable to non-planar deformation, have been
developed for regions of continuous deformation (Cobbold and Percevault, 1983) and adapted to
discontinuous deformation of regions dominated by strike-slip motion (Audibert, 1991), normal
faulting (Rouby et al., 1996) and reverse faulting (Bourgeois et al., 1997). However, all these methods
require in addition to balanced structural sections, structure contour maps of the layer to be restored. In
the case of the EC such methods have up to now been impossible because structure contour maps do
not exist for the EC. These methods would require a detailed knowledge of the 3-D structure, which is
beyond present-day knowledge of this mountain range. In this Chapter I use a modified method for map
view structural restoration that does not require structure contour maps and applied it to restore the
Neogene Andean deformation of the EC.
The aim of this chapter is to restore in map view the Neogene Andean deformation of the EC
and neighbouring areas (Figs. 1.1 and 5.1) using all data available in literature. I compiled more than 50
balanced cross-sections and used the amounts of shortening estimated in those cross-sections.
Understanding the Neogene compressional Andean deformation mechanisms is important to
understand development of petroleum structural traps during the Neogene and the preservation of early
formed traps.

2. TECTONIC SETTING

2.1. PRESENT DAY PLATE-TECTONIC SETTING

The Colombian Andes lie in a complex tectonic area of convergence between the South
American, Nazca and Caribbean plates (Fig. 5.1, e.g. Freymueller et al., 1993). Kellogg et al. (1985)
have interpreted the tectonics of the region in terms of two additional microplates, Panama and the
North Andes, both of which are subject to internal deformation (Fig. 5.1). At present, subduction (70
mm/yr) of the oceanic Nazca Plate occurs at the western margin of South America, and the trench is
seismically active (Fig. 5.2; Freymueller et al., 1993). The North Andes are bounded by the Colombia-
Map View Restauration of Transpressional Basin Inversion in the Eastern Cordillera

Colombian Basin
South Caribbean deformed belt
CARIBBEAN PLATE

elt
Belt
Maracaibo Block

mad B
North Panama

mad
fold belt

defor
def or
PANAMA
Sinu
es
BLOCK nd

into
aA
San Jac ir d
e
M
Venezuela

RA
LE
CHOCO DIL
OR
BLOCK
NC

NAZCA
ER

SOUTH
ST

PLATE AMERICAN PLATE


Colombia
EA
BIA
LOM
CO
ch

N
ren

ER

Detail in Fig. 5.3


or t

ST
uad

WE
-Ec

PLATES
OF
bia

ES
lom

Continental plate
AN
Co

RR

Oceanic plate
TE

DEFORMATION ZONES
IC
AN
CE

Eastern Andes
-O

O
L AE
Accreted oceanic terranes
PA

0 500 km ESTRUCTURES
Thrust fault
Strike-slip fault
Normal fault

Figure 5.1. Present day plate-tectonic setting and major structures of the Northern Andes in the NW corner of South
America.

Ecuador Trench and Panam Block on the west, the South Caribbean deformed belt to the north, and
the Bocon and East Andean fault zones to the east (Figs. 5.1 and 5.2). Within the northern part of the
Andes, significant displacements have occurred on the Oca (right-lateral, east-west trending) and Santa
Marta-Bucaramanga (left-lateral, northwest-southeast trending) faults (Fig. 5.2; e.g. Campbell, 1968;
Irving, 1971; Schubert, 1982). The Bocon Fault, Santa Marta-Bucaramanga Fault, and Oca-El Pilar
faults (Fig. 5.2) define the wedge-shaped Maracaibo Block which has moved northward relative to
stable South America during late Cenozoic (Bowin, 1976; Mann and Burke, 1984; Pindell and Erikson,
1993). On the east side of the North Andes, right-lateral transpressive movement is taking place on the
Bocon and east Andean (Guaicramo) fault systems (Fig. 5.2, e.g. Pennington, 1981). The Bocon
Fault in the Mrida Andes of Venezuela is mainly a right-lateral strike-slip fault (Fig. 5.2, Schubert,

182
Chapter 5

CARIBBEAN PLATE

Colombian Basin
South Caribbean deformed belt

Oca fault

lt
El Pilar Fault

d Be
lt

Buc
Maracaibo

rme
d Be

San anga Fa
North Panama Block

defo Sinu

aram
defo
rme
fold belt

ta M
U ram ita

into

arta ult
s
de

Jac
An

em
fault syst al
Panama Block rida

Romer
fault M e ult
fa 0 500 km
San
no
co
BUCM
Bo
Choco Block

tal
on
Fr
m o)
ste am
EC

Sy ar
ult uic
BOGO
NAZCA PLATE VILL Fa n (G
a
de
An

CA LI
st
Ea
h
nc
re
rT
o
ad
cu

PA S T
-E

SOUTH AMERICAN PLATE


bia

ES M E
lom

JE R S
PLATE VELOCITY VECTORS W ITH RESPECT TO
Co

STABLE SOUTH AMERICA


(from Kellogg et al., 1985 and Freimueller et al., 1993)
LA T A

REL ATIV E DIS PACEM E NT O F SEVERAL G PS


S TATI ONS W I TH RESPECT TO VIL L A VICENCI O
ult STATI O N (STABL E S O UT H AM ERI CA)1 9 9 1 T O 19 9 4
a Fa (fro m Mora, 19 95 , in Aco s ta et al., 2000)
niz ult
Lli
Fa 20 mm/yr
ga
tan ESTRUCTURES
ala
P Plate velocity vector relative
to South American Plate Thrust fault
Plate velocity vector relative Strike-slip fault
to the North Andes Block Normal fault

Figure 5.2. Tectonic features and boundaries of the North Andean Block. All plate velocity vectors are relative to
the South American plate, except the grey vector, which represents the Caribbean-North Andes vector of Kellogg et
al. (1985, modified by Freymueller et al., 1993). Relative displacement measurements (mm/year) from 1991 to 1994
of several Global Positioning System (GPS) stations located within the Andes with respect to the Villavicencio
station located in the Llanos Basin (stable South American Plate). (modified from Mora, 1995 in Acosta et al.,
2000).

1982). Sub-parallel thrust faults flank the Mrida Andes suggesting significant compression normal to
the Bocon Fault (Henneberg, 1983). The East Andean fault system (faults limiting the eastern border
of the EC) consists of west-dipping reverse faults (Fig. 5.2). Based on a study of focal mechanisms
Pennington (1981) showed that transpressive right-lateral slip is occurring on these faults, and that the
North Andes Block is moving NNE relative to the South American Plate. To the south the Ecuadorian
Andes include a system of sub-parallel northeast-trending right-lateral strike-slip faults and north-
trending thrust faults (Fig. 5.2). The relationships between these faults are not well understood. Focal
mechanism studies suggest thrust faulting on a N-S trending fault as well oblique transpressional

183
Map View Restauration of Transpressional Basin Inversion in the Eastern Cordillera
displacement (Lyon-Caen et al., 1990). Based on observed slip rates, Kellogg et al. (1985) predicted
that the North Andes relative to South American Plate is moving approximately 10 mm/yr toward the
NE (Fig. 5.2). The Southern Caribbean Plate boundary is generally accepted to be a distributed zone of
deformation, although there remains disagreement as to how much of the North Andes should be
considered Caribbean. The principal part of the deformation zone consists of a series of deformed
belts along the Caribbean coast of South and Central America (Pindell and Barret, 1990; Mann et al.,
1990). In Colombia the Southern Caribbean deformed belt includes the Sin and San Jacinto belts (Fig.
5.2). Kellogg et al. (1985) used a block vector model to estimate the convergence rate between the
Caribbean and North Andes to be 17 mm/yr to the southeast (Fig. 5.2). Dewey and Pindell (1985)
found a similar convergence between the Caribbean Colombian Basin and the Maracaibo Block (Fig.
5.2).

3. STRUCTURE OF THE EASTERN CORDILLERA

In the EC (Figs. 1.1, 5.1 and 5.3) geological mapping is largely based on photogeologic
interpretations. Structural field studies are scarce and seismic data is available only for the axial region
and eastern and western foothills. Frequently seismic data are of poor quality (Dengo and Covey,
1993). Structural interpretations are thus based on limited data and conceptual models. All structural
interpretations recognize reverse faults limiting the mountain range in its eastern and western borders.
Earlier interpretations (Campbell and Burl, 1965; Julivert, 1970) emphasized high-angle faults and
assume a dip increase with depth for the fault surfaces. In contrast most recent structural interpretations
based on local seismic data (Colletta et al., 1990; Dengo and Covey, 1993; Cooper et al., 1995; Roeder
and Chamberlain, 1995) emphasize thin-skin deformation and flattening at depth. Presence of both
types of faults has been assumed in other interpretations (Kammer, 1993b, 1996, 1999a, b; ESRI and
Ecopetrol, 1994; Linares, 1996; Kammer and Mojica, 1996; Fajardo-Pea, 1998).
Based on outcrop observations Campbell and Burgl (1965) and Julivert (1970) recognized the
general double vergence of the EC. They recognized two main thrust systems developed in the margins
of the Cordillera (Eastern Guaicramo thrust system limiting with the LLA basin and western La Salina
thrust system limiting the MV Basin; Fig. 5.3) and identified the main thrusts. They recognized some
asymmetry in the EC, for example the Quetame Massif (Fig. 2.1) is only present in the eastern margin
of the mountain range. In the western margin they recognized low angle east-dipping thrusts parallel to
the mountain front, with a series of splays from the mountain front extending into the Magdalena
Valley at a 20 angle (Fig. 5.3). The splays pass into north-plunging anticlines (Campbell and Burgl,
1965). These faults are in an en-chelon arrangement (Julivert, 1970). Within the Middle Magdalena
Valley the structures are gentle and consist of a series of anticlines (containing oil), usually associated
with reverse faults. These structures tend to converge in a southerly direction with the La Salina Fault
(Fig. 5.3; Julivert, 1970).
Julivert (1970) suggested that the Eastern Cordillera shows a fan structure, with some similarity
to that of the Pyrenees and differentiated the mechanically different behaviour of the heterogeneous
Precambrian and Palaeozoic metamorphic and sedimentary mechanical basement from the Mesozoic
and Cenozoic sedimentary cover. In fact he suggested that the main structural features of the chain are
determined by the relationships between basement and sedimentary cover: (1) Passive behaviour of
undeformed tilted cover on basement blocks or simple draping structures occur in regions where the
thickness of the preserved sedimentary cover is small; (2) Complex tight structures, where the
basement has less rigid behaviour occur in areas where the thickness of the preserved sedimentary
cover is larger; (3) Independent behaviour of the cover, due to gravity (small-scale collapse structures
driven by gravity and favoured by erosion), to salt (salt in anticlines but without clear evidence of
typical salt domes in the Sabana de Bogot area; Fig. 2.1 for location), and to disharmonic relationships
occur in areas where the thickness of the preserved sedimentary cover is large (e.g. axial region Sabana
de Bogot). In these regions the dcollement of the cover from its basement can be facilitated by thick
shales and accompanied by horizontal translation. The basement crops out within the mountain range in
two fault-bounded massifs arranged en-echelon: the southward-plunging Santander-Floresta and the
northward-plunging Garzn-Quetame massifs (Fig. 5.3). A rather narrow strip between the two massifs
in the

184
Chapter 5
1700
Sierra
Nevada
de Santa
Marta

1600

Lower
Magdalena
Valley
s
1500 de
An
id a
er
M

1400

TA

1300
PP

SO

NM

BF
h
FC
a r
dille

1200 OP
Cor

a
FS
FC

AC
h
FD

1100
o
FS

FT

CF
1000
FG
FB

LLANOS ORIENTALES
BASIN (LLA)

900
0 100 km
al
Centr

1.200.000 1.300.000
Neogene structures
800
FA

Paleogene or
La Macarena
Serrania de

older structures

ESTRUCTURE
S Thrust
700 Strike-slip
fault
Normal
fault
Guyana Shield
fault
Anticlin
S
eynclin
e
800 900 1000 1100 1200 1300 1400

Figure 5.3. Structure map of the EC and surrounding areas. (from Geotec, 1976; ESRI and Ecopetrol, 1994) and
subsurface maps of the Llanos (Ecopetrol and Beicip, 1995), Magdalena Valley (Ecopetrol et al., 1994; Geotec,
1994) and Maracaibo basins (Roure et al., 1997). In the Llanos, Magdalena Valley and Maracaibo basins
Palaeogene structures have been recognized in the subsurface below Neogene sediments. BF Bucaramanga Fault.
Fsa La Salina Fault, FB Bituima Fault, FT Cusiana-Tamara Fault, FCh Chucur Fault, FC Cambras Fault, FDh Dos
Hermanos Fault, FG Guaicramo fault, FSo Soapaga Fault, FA Altamira Fault, NM Nuevo Mundo Syncline, AC
Arcabuco Anticline, PP Paya-Provincia Anticline, OP Opn Oil Field, SO La Salina Oil Field, CF Cupiagua Oil
Field.

185
Map View Restauration of Transpressional Basin Inversion in the Eastern Cordillera
axial part of the range, shows a more tightly folded synclinorium structure where Tertiary rocks have
been preserved (Julivert, 1970).
Based on surface geology and limited seismic data Colletta et al. (1990) concluded that the
Cordillera could be interpreted as the result of inversion of Jurassic-Cretaceous basins during the
Miocene-Pliocene shortening event (Andean orogeny). Reactivation of normal faults could have
induced ramp anticline structures such as the Arcabuco Anticline (Fig. 5.3). They recognized that the
opposite vergence of the mountain range is not symmetric. The amount of shortening they estimated
from a regional balanced cross-section (Fig. 5.4) is 105 km (L1/L0 66%). In a later paper (Colletta et
al., 1995) suggested based on analogue models that in the Llanos foothills (Fig. 5.3) backthrusts and
forward thrusts developed at the same time and that their intersection corresponds with a potential
dcollement horizon. They suggested that due to the presence of a regional dcollement level at the
base of the Carbonera Fm and to intense erosion, triangle zones and passive roof duplexes could
develop along the Andean foothills. Tectonic inversion could have produced short cut low-angle
faults while some of the existing normal faults are expected to have been passively transported without
reactivation.
Dengo and Covey (1993) extrapolated two dominant styles of deformation to the entire
mountain range: (1) Basement-detached fold-and-thrusts, characterized by low-angle thrust faults and
fault-ramp anticlines that involve Jurassic through Tertiary rocks. They assume, without proof, that the
geometric models of fault-bend and fault-propagation folding (Suppe, 1983; Suppe and Medwedeff,
1990) are applicable to the whole chain; (2) Basement-involved reverse faults, many of which have a
high-angle dip at surface. They propose that these basement-involved reverse faults were active during
Pliocene-Pleistocene and displace earlier (Oligocene? to Miocene) thrust faults. Their regional
palinspastic restoration suggests approximately 40% shortening (150 km) in the sedimentary cover and
one third of this value as basement shortening. In order to explain the difference in shortening between
basement and cover they proposed as responsible for basement shortening a hypothetical low-angle
mid-crustal detachment that extends westward to the Central Cordillera up to the western edge of the
South American Plate. They speculated without evidence the existence of a decapitated graben below
the basement rocks of the MV.
Based on the assumption that the whole Cretaceous sedimentary record represents a thermal sag
basin fill following Jurassic rifting and ignoring earlier work by Fabre (1987, who demonstrated the
existence of an Early Cretaceous extensional basin), Roeder and Chamberlain (1995) reconstructed the
geometry of the Cretaceous basin assuming it originated from the flexural subsidence produced by a
line load. They assumed that this flexural model was wholly responsible for the thickness variations
within the Cretaceous, and positioned stratigraphic columns reported by Colletta et al. (1990) within
this template. As a result of these assumptions they constructed a more than 350 km (their figure 6)
long cross-section for the Cretaceous basin that, compared to the present-day width of the EC, would
imply 230 km of shortening. They extended downward the Dengo and Covey (1993) structural section
assuming low-angle thrust faults affecting the whole upper crust fitting a Moho geometry produced by
a line loading flexural deflection. They postulated that a polyphase history involving foreland upthrusts
redeforming the fold-and-thrust belt is unlikely and suggested that, if anything, foreland upthrusting
occurred prior to the main deformation event. Their speculative interpretation shows three major
basement-involved thrust sheets in the Cordillera. The shallow, westward-dipping slices imbricate
basement and sedimentary cover and truncate pre-existing rift geometry.
Based on surface geology and limited seismic data Cooper et al. (1995) returned to a rift
inversion interpretation (Fig. 3.37 Chapter 3). Their shortening estimate is 68 km. They also recognized
thickening of Jurassic, Early Cretaceous and probably pre-Jurassic (?) sediments in the hanging-walls
of the Guaicramo, Arcabuco (Soapaga Boyac) and La Salina fault systems (Fig. 5.3) and concluded
that the EC structure is the result of the inversion of a Jurassic-Cretaceous basin. They also suggested
that the Cusiana-Tmara eastern frontal thrust fault (Fig. 5.3) system originated as a Late Cretaceous
normal fault and continued to be episodically active as a normal fault until mid-Miocene, in order to
accommodate flexural loading produced by compression and surface-uplift of the Central Cordillera
and the EC. They also recognized shallow thin-skinned thrusting.

186
NW Fault bounded block section length after deformation
SE
L0 = 219 km
Fault bounded block in the
deformed belt
Magdalena

Arcabuco Ant.
Pesca F

Velez Ant.
Guicaramo F
Yopal F
Llanos Basin
Valley Neogene

La Salina F
0
Palaeogene
Km 14
Cretaceous
Early Mesozoic

187
Fault bounded block section length before deformation
Chapter 5

Basement
L1 = 219 km + 105 km

Lf 0 100 km
0
Km 8

Figure 5.4. Balanced cross-section of the EC illustrating the double vergence geometry of the mountain range generated by inversion of a
Mesozoic extensional basin. (from Colletta et al., 1990). L0: Pre-deformation length between pin lines. L1: Present-day length between pin
lines. Location of the section in Figure 5.8.
Map View Restauration of Transpressional Basin Inversion in the Eastern Cordillera
Jones (1995) presented a radical alternative hypothesis for the structure of the EC. He proposed
that both the EC and the Sierra de Perij (Fig. 5.3) form elements of a single southeast-verging thrust
sheet with minor imbrications. A single basement-involved fault-bend fold is interpreted to underlie the
whole mountain range in this model, with a huge triangle zone at the eastern front. The west-verging
structures in the Magdalena Valley foothills are interpreted as large gravitational slides developed over
a west-dipping ramp. The total shortening according to this model is 185 km.
Linares (1996) based on surface geology and limited seismic data interpreted the structure of
the EC as the result of inversion of an extensional Mesozoic basin. Reactivation of basement normal
faults resulted in both thick and thin-skinned structures related to a single tectonic event. Linares
(1996) estimated a minimum of 50 km of shortening at a regional cross-section through the
southernmost part of the Sabana de Bogot area (20 % for the whole Cordillera). A similar
interpretation was proposed by Fajardo-Pea (1998) for the Tunja axial region of the EC. However, he
proposes five cycles of accumulation and deformation between late Maastrichtian and Holocene times
with three peaks of deformation during the Early-Middle Eocene, the Early-Middle Miocene and Late
Miocene-Early Pliocene. Inversion of the Magdalena-Tablazo graben (between the La Salina-Landzuri
and Boyac palaeonormal faults) and the Cocuy half graben (between the Soapaga Fault and the
Guaicramo palaeonormal fault) generated the western and eastern flanks of the EC. The Floresta less
subsiding block during Mesozoic time correlates with the Tunja axial region where Neogene
sedimentary record has been preserved. Tertiary deformation began by inversion of the Magdalena-
Tablazo graben producing folding within a regional pop-up structure. The Soapaga Fault is interpreted
as a footwall shortcut of the Boyac Fault (Fajardo-Pea, 1998). This author recognized thin-skinned
thrusting in the Tunja and Sogamoso regions with detachments in Lower and Upper cretaceous shales
and possible positive flower structures in the northern part of the Sogamoso region.
Based on seismic data in the Middle Magdalena Valley foothills, Kovas et al. (1982) suggested
that hanging-wall anticlines (Paya-Provincia and La Salina oil fields; Fig. 5.3) were formed above
west-directed en chelon thrust faults on the western side of the Nuevo Mundo Syncline (Fig. 5.3).
They suggested that the thrust faults are probably flexural-slip features associated with the folding of
the syncline. They also showed progressive transfer of displacement from the Paya to the La Salina
faults (Fig. 5.3) arranged in chelon. In the hanging-walls of these faults they also described
backthrusts. To the southward, Steuer et al. (1997) described the Opn structure (gas field, Fig. 5.3 for
location) as a large antiform located along the La Salina Fault (Fig. 5.3), comprising a wedge bounded
by upper and lower detachments with opposite vergence. They interpreted it as the result of a complex
history with three episodes: (1) Post-Paleocene to pre-late Eocene compressional deformation which
exhumated and eroded the marine Cretaceous rocks; (2) Deposition of the Upper Eocene to Middle
Miocene dominantly fluvial detrital rocks; (3) Folding and thrusting beginning at approximately 10.5
Ma ago.
Even further southward in the Magdalena Valley foothills area and western flank of the
Cordillera, Restrepo-Pace et al. (1999a,b) interpreted 3 events in the west-verging fold-and-thrust belt,
which consists of a complex thin-skinned thrust system resulted from polyphase deformation: (1) West-
verging thin-skinned closely spaced imbricate thrusts involving Paleocene sediments, concealed below
the Eocene unconformity; (2) Reactivation of these structures in a break-back sequence that began in
mid-Late Miocene time as indicated by syn-kinematic deposits and onlap relations. That event
generated frontal intercutaneous wedges and hinterland west-verging closer spaced thrusts; (3) Major
inversion of fundamental hinterland faults creating the present structural relief.
In the Llanos foothills, Rathke and Coral (1997) described the Cupiagua condensate- gas field
(Fig. 5.3) as a tight asymmetric thrust-controlled anticline with a frontal very steep to overturned and
highly imbricated limb. The back limb is broken by several backthrusts. A major tear fault associated
with a lateral thrust ramp crosses the structure from west to east. They interpreted that the early
development of the structure was simultaneous with deposition of the lower part of Carbonera Fm
(Oligocene).
De Freitas et al. (1997), Branquet et al. (1999b) and Branquet (1999) have emphasized the
importance of strike-slip faults in the EC. Branquet (1999) and Branquet et al. (1999b) identified en-
echelon folds in the hanging-wall of major basement-involved faults, while basement pop-up structures

188
Chapter 5
limited to the east by convex-upward reverse faults were interpreted by them as positive flower
structures. They proposed that these structures formed as the result of dextral transpressional inversion
of an Early Cretaceous half-graben during the Andean orogeny. They also emphasize that Andean
inversion of normal faults would be impossible without involving an important strike-slip component.
Based on outcrop structural data Kammer (1993b, 1996) proposed the existence of high-angle
normal and reverse faults affecting the basement of the EC (e.g. Santander and Floresta Massifs, Fig.
2.1). According to this author the Santander Massif is a broad basement uplift limited on both flanks by
major reverse faults. He also recognized reverse faults in the eastern border of the Floresta Massif. He
described more internal faults as high-angle normal faults limiting half-grabens. He interpreted the
Neogene fault pattern of the Santander Massif as inherited from a Jurassic extensional event. In
contrast Kammer and Mojica (1995, 1996) recognized that deformation of the sedimentary cover is
dominated by folding, implying significant shortening, and interpreted them as having been produced
by gravitational sliding of the sedimentary cover detached from the basement. For the folded
Cundinamarca sub-basin area, Kammer (1999a) suggests that there is a lack of a deformational contrast
between sedimentary cover and basement. Mora and Kammer (1999) have presented an illustrative
example of the basement-sedimentary cover relationships in the eastern flank of the EC between the
Quetame Massif and the Sabana de Bogot area. In the Quetame Massif (Chingaza dome) deformation
affects both basement and cover and precludes, by its homogeneous nature, a detachment within the
Lower Cretaceous shales. The Chingaza dome displays open second-order folds, associated to an axial
plane cleavage with pytgmatic folds indicative of local shortening values up to 40%. In contrast the
Sabana de Bogot area folding occurred by flexural slip, layer parallel strain is absent, and geometry of
the folds suggests detachment horizons within the sedimentary cover. In the western flank of the EC
west of Bogota similar observations have been described by Cardozo-Puentes (1989).
Based on characteristic relay patterns of marginal folds at both borders of the EC: right-stepped
fold pairs along NW- trending borders and left stepped structural relays along N-S trending
deformation fronts Kammer (1999b) suggested a transpressional deformation with a strain direction of
about 300. He compared these patterns with outcrop fault slickenside measurements mostly indicative
of strike-slip deformations, their populations define, pseudo-conjugate sets, whose compression or
extension directions are indicated by reverse or normal fault pairs. In the Santander Massif compression
directions are oriented both perpendicular to the NNW striking Bucaramanga Fault and to N-S
trending, faults. To the west of the Bucaramanga Fault and affecting the EC the striated faults suggest a
clockwise departure from the perpendicular to the structural trend. In contrast, joints evidence a pre-
folding origin. This author proposed that the model of simple transpression apply to the EC while in the
Santander Massif simple transpression is partitioned into a strike-slip deformation along the
Bucaramanga Fault and a homogeneous pure shear in the internal part of the massif.
Casero et al. (1995, 1997) proposed that the structure of the EC is the result of reactivation of
previous structures. They suggested several tectonic episodes to explain the build up of the EC: (1)
Strong compressive deformation of Lower Palaeozoic (Lower Ordovician) rocks and subsequent
peneplanation; (2) Late Cretaceous flexuration followed by a strong orogenic deformation predating
deposition of Paleocene regressive sands; (3) Late Eocene compressional deformation and subsequent
moderate erosion; (4) Oligocene-Early Miocene flexuration of the EC foreland followed by moderate
compressional deformation; (5) Main Andean orogeny until Miocene-Pliocene times and is still active.
They suggest that the EC corresponds to the Lower Cretaceous rift area and represents a mega-
inversion and shortening of the same.
Based on surface geology and limited seismic data Schelling (1994) interpreted the structure of
the EC as a large basement rooted tectonic pop-up block. They made six regional balanced cross-
sections across the Cordillera (one in the Upper MV). They proposed that the LLA and Middle
Magdalena fold-and-thrust belts (Fig. 5.3) are divergent, thin-skinned deformational systems, that
define the borders of the Cordillera, while the other internal structures are characterized by basement-
rooted thrust faults and associated basement cored, hanging-wall structural uplifts. They interpreted the
Cordillera as a series of inverted graben and half-graben structures. They estimated tectonic shortening
between 45 to 65 km, or 18 to 26%.

189
Map View Restauration of Transpressional Basin Inversion in the Eastern Cordillera

4. REGIONAL MAP VIEW RESTORATION OF THE NW CORNER OF SOUTH AMERICA

4.1. METHOD

Our aim is to restore the Andean deformation of the EC. In order to establish some boundary
conditions for the restoration, I first carried out a regional map view restoration of the NW corner of
South America, which defines the tectonic framework for our more local restoration of the EC. I
applied a methodology similar to that of Laubsher (1987).
I have manually restored in map view a mosaic of regional fault-bounded blocks using all the
kinematic constraints available in literature. I followed these steps:

(1) I compiled a regional tectonic map, including tectonic plates, deformation belts in the North
Andes and major structures (Fig. 5.1). Between the major tectonic plates, i.e. South America,
Nazca and Caribbean, I identified the northern Andes and the Panama Block as deformed belts.
The north Andes deformed belt consists of (1) a deformed margin of the South American Plate
bounded to the west by the Romeral suture zone and (2) the accreted oceanic terranes of western
Colombia and northern Venezuela (Fig. 5.2).
(2) I selected a number of tectonic blocks in the Eastern Andes. In order to simplify the model I have
used the minimum number of blocks (Fig. 5.5). Because I am interested in restoring the
deformation of the EC (including the Santander Massif) and its two northern prolongations, the
Sierra de Perij and the Mrida Andes, I assume them to be deformed belts located between rigid
blocks (Fig. 5.5). The stable South American Plate is the larger stationary reference block (Fig.
5.5) and displacements of all other blocks are related to it. To restore the deformation of the
deformed belts (EC, Sierra de Perij and Mrida Andes) I assumed the Maracaibo, Central
Colombia (Central Cordillera and Magdalena Valley) and Santa Marta blocks to be rigid (Fig.
5.5). This assumption is based on the observation that during Neogene time their internal
deformation is small, most deformation in these areas appears to have occurred during the
Palaeogene.
(3) I compiled all the kinematic constraints available in literature such as strike-slip displacement
estimates and amounts of shortening from available balanced cross-sections (Figs. 5.5 and 5.6,
Table 5.1).
(4) I manually restored the mosaic of major tectonic blocks according to the kinematic constraints
(Fig. 5.7) by means of rigid body translations and rotations. This regional restoration is similar to
that done by Laubsher (1987) and by Villamil and Pindell (1998) in northern South America,
although their work did not apply to the EC.

A: BALANCED CROSS-SECTIONS USED IN THE MAP VIEW RESTORATION


No. Reference
1 to 4 Findlay (1988)
5 Roeder and Chamberlain (1995)
6, 7 Colletta et al. (1997)
8 Kellogg (1984)
9 Namson et al. (1994); Linares (1996)
10 Amzquita and Montes (1994)
11 Amaya and Santamara (1994)
12 Etayo-Serna and Florez (1994)
13 Laubscher (1990); Dengo and Covey (1993)
14 to 16 Covey and Dengo (1987); Dengo and Covey (1993)
17, 18 Toro (1998)
19 Laubscher (1990); Namson et al. (1994)
20 Butler (1983); Ressetar and Schamel (1986); Butler and Schamel (1988); ESRI-HOCOL (1989); Laubscher (1990)

Table 5.1 A and B: Structural cross-sections used in the map view restoration of the Eastern Cordillera. Location of
the sections is shown in Fig. 5.6.

190
Chapter 5

No. Reference
21 Kellogg and Duque (1994); Kellogg et al. (1995)
22 Dengo and Covey (1993)
23 to 25 Chigne and Rojas (1997)
26, 27 Cooper et al. (1995)
28 Ecopetrol et al. (1995)
29 to 31 Naar and Coral (1993)
32 Colletta et al. (1990)
33 Butler and Schamel (1988)
34, 35 Butler (1983); Ressetar and Schamel (1986); Butler and Schamel (1988); ESRI-HOCOL (1989)
36 to 41 Schelling (1994)
42 to 45 Covey and Dengo (1987)
46 Laubscher (1990); Galindo et al. (1994)
47, 48 Galindo et al. (1994)
49 Geotec (1994)
50 to 64 Peel and Hossack (1987)
65 to 70 British Petroleum (1989)

B: NOT BALANCED CROSS-SECTION WITH APPROXIMATE BED LENGTH PRESERVATION, USED IN


THE MAP VIEW RESTORATION.
No. Reference
71 Casero et al. (1995, 1997)
72, 73 Linares (1996)
74 Ecopetrol et al. (1994)
75 Steuer et al. (1997)
76 Ecopetrol et al. (1995); Ecopetrol-ICP (1995)
77 Ecopetrol-ICP (1995)
78, 79 Butler (1983); Ressetar and Schamel (1986); Butler and Schamel (1988); ESRI-HOCOL (1989)
80 to 82 Ecopetrol Informe No 2133 Tunja area
83 to 102 Ecopetrol and Beicip (1995)
102, 103 Ecopetrol and Beicip (1995); Linares (1996)
104, 105 Ecopetrol and Beicip (1995)
106 Casero et al. (1995, 1997)
107 Corredor (1997)
108 British Petroleum (1987, 1988, 1989)
109 Butler and Schamel (1988)
110, 111 Laubscher (1990)

Table 5.1 A and B: Continued

4.2. RESULTS

According to the regional map restoration of NW South America (Fig. 5.7), convergence and
clockwise rotation of the Central Colombia micro-plate (Central Cordillera) relative to South America
resulted in transpresion with northward increasing shortening in the EC of Colombia. In Ecuador the
Andean Block was mainly displaced by right-lateral strike-slip (50 km) with only minor shortening. In
contrast the maximum shortening of the EC (from 324 km to 210 km, i.e. 114 km) and clock-wise
rotation (+6.5) of the Central Colombia Block (Central Cordillera) occurred in the north (north of
latitude 6 N) at the latitude of the Choc-Panam Block. Although the restoration does not explain the
cause of Andean deformation, this result is in accordance with the hypothesis that the Andean
deformation resulted from the collision of the Choc-Panam Block during Neogene time as suggested
in literature (e.g. Duque-Caro, 1990). However, for the EC the maximum amount of shortening (from
324 km to 210 km, i.e. 114 km) occurred between 6 and 7 N, related to the inversion and thrusting of
the Mesozoic Cocuy and Tablazo sub-basins. Although the amount of E-W shortening between the
Central Colombia Block and the South American increases northward, it has partially been absorbed by
the right-lateral strike-slip displacement of the Bocon Fault (60 km; Fig. 5.7). The restoration

191
Map View Restauration of Transpressional Basin Inversion in the Eastern Cordillera

C. COMPLEMENTARY NOT BALANCED STRUCTURAL SECTIONS USED TO CHECK CONSISTENCY OF


THE MAP VIEW RESTORATION.
NORTHERN VENEZUELA LOWER MAGDALENA VALLEY SANTANDER MASSIF
Av Lallemant (1997) Breen (1989) Boinet (1985)
Duval and Valds (1995) Duque-Caro (1979, 1984) Boinet et al. (1985)
MARACAIBO BASIN MIDDLE MAGDALENA VALLEY Kammer (1993b)
Colletta et al. (1997) Ecopetrol et al. (1994) Laubscher (1987)
Laubscher (1987) Kovas et al. (1982) EASTERN CORDILLERA
Roure et al. (1997) Mojica and Franco (1990) British Petroleum (1987-1989)
SIERRA DE PERIJ Namson et al. (1994) Calvache and Muoz (1984)
Araujo (1997) ESRI - HOCOL (1989) Cardozo-Puentes (1989)
Kellogg (1984) UPPER MAGDALENA VALLEY Campbell and Burgl (1965)
Laubscher (1987) Butler and Schamel (1988) Casero et al. (1995, 1997)
Testamarck et al. (1994) Galindo et al. (1994) Fabre (1987)
MRIDA ANDES Mojica and Franco (1990) Fajardo and Chamorro (1994)
Laubscher (1987) Ojeda and Pea (1994) Irving (1971)
Meier et al. (1987) Ressetar and Schamel (1986) Julivert (1970)
LLANOS ORIENTALES Toro (1992) Kammer (1993b)
Ecopetrol and Beicip (1995) Van der Wiel (1991) Kammer (1999a)
Kluth et al. (1997) GARZON MASSIF AND FOOTHILLS Kammer and Mojica (1995)
McCollough and Carver (1992) Butler and Schamel (1988) Kammer and Mojica (1996)
LLANOS FOOTHILLS Casero et al. (1995, 1997) Keeley and Sarez (1997)
British Petroleum (1987-1989) Kammer and Mojica (1995) Laubscher (1990)
Casero et al. (1995, 1997) Laubscher (1990) Linares (1996)
Ecopetrol-ICP (1995) Van der Wiel (1991) McLaughlin (1972)
Linares (1996) PUTUMAYO-ORIENTE BASIN Occidental et al. (1996a,b)
Naar and Coral (1993) Baby et al. (1998) Toro (1990)
Rathke and Coral (1997) Dashwood and Abbotts (1990) Torres (1996)
Ureta and Du Toit (1997) Ujueta (1993)

Table 5.1 C: Continued

(compare Figs. 5.5 and 5.7) suggests that the total amount of shortening perpendicular to the deformed
belts in the north (10 km of shortening has been estimated in the Sierra de Perij and 50 to 60 km of
shortening in the Mrida Andes) is less than the maximum shortening (114 km) of the EC between 6
and 7 N.
Convergence (from 100 to 65 km, i.e. 35 km) and left-lateral strike-slip movement (100 km) of
Central Colombia (Central Cordillera and MV) relative to Maracaibo produced the northern part of the
EC in the Santander Massif region. Clock-wise rotation of the Central Colombia Block relative to
Maracaibo is +3.5 (Figs. 5.5 and 5.7).
Convergence (from 45 km to 35 km, i.e. 10 km) of the Santa Marta Block relative to Maracaibo
generated the Perij Mountain Range. Restoration suggests also that clock-wise rotation of the Santa
Marta Block relative to the Maracaibo Block (+7) took place. Finally convergence (50 to 60 km) and
dextral strike-slip motion (60-km) of the Maracaibo Block relative to South America generated the
Mrida Andes of Venezuela (Figs. 5.5 and 5.7).

5. MAP VIEW RESTORATION OF THE EASTERN CORDILLERA

5.1. METHOD

I carried out a more detailed map view restoration of the EC of Colombia (Figs. 5.1, 5.2 and
5.3), using balanced cross-sections from the literature (Figs. 5.4, 5.6 and Table 5.1). The aim is to
reconstruct the undeformed initial state. I used a modified version of the method applied by Bourgeois
et al. (1997) to restore thrust sheets in map view in the Tajik Depression, central Asia.

192
Chapter 5
15 N

14 N

13 N
CARIBBEAN PLATE
GUAJIRA BOCK
12 N

R
11 N SANTA

ain
MARTA

unt
BLOCK

Mo
10 N

Buca
MARACAIBO

rija
BLOCK

Pe
rama
9N

Santa
nga F

nder
s
8N de

ault
An

Mass
r ida
e
7N M

if
k
Bloc

a
ler
6N

il
ord
mbia

NAZCA PLATE C
Colo

rn
5N
st e
Ea

4N
ES

tral
AN

Cen
RR

3N
TE

t
ul
Fa
ED

2N ira
ET

tam
Al
RR

Detail in Fig. 5.8


AC

1N
SOUTH AMERICAN PLATE
IC
AN
CE

0 Continental plate and


tectonic blocks
-O
O
AE

1S Deformed belts
L
PA

2S Kinematic constraints and location


of selected balanced cross
sections from literature

3S
Shortening
Strike-slip displacement
4S
0 500 km
5S
83 W 82 W 81 W 80 W 79 W 78 W 77 W 76 W 75 W 74 W 73 W 72 W 71 W

Figure 5.5. Tectonic blocks used in the regional map view restoration of the NW corner of South America.
Kinematic constraints include: (1) Regional balanced cross-sections across the Sierra de Perij (Kellogg, 1984), the
Mrida Andes (Colletta et al., 1997), and the EC (Colletta et al., 1990; Dengo and Covey, 1993; Schelling, 1994;
Cooper et al., 1995; Roeder and Chamberlain, 1995; Linares, 1996); (2) Strike-slip displacement estimates along the
El Pilar-Oca fault system (Pindell, 1993; Villamil and Pindell, 1998), Bocon Fault (Schubert, 1982; Henneberg,
1983; Pindell, 1993; Pindell and Erikson, 1993; Colletta et al., 1997), Bucramanga-SantaMarta Fault (Campbell,
1968; Irving, 1971) and Altamira Fault (Casero et al., 1995, 1997). In the restoration I assumed the South American,
Central Colombia, Maracaibo, Santa Marta and Guajira blocks were rigid and their relative Neogene movement
generated deformed belts of the EC (including the Santander Massif) and its northern prolongations the Sierra de
Perij and the Mrida Andes. I did not include in the restoration oceanic accreted terranes, the Choc-Panam Block
or the Nazca and Caribbean plates.

193
Map View Restauration of Transpressional Basin Inversion in the Eastern Cordillera

0 100 km 8

Buc
aram
7

anga

Sierra
fault
s

de Pe
de
An
a
rid

rija
Me

s
uca
an L

San
Cucuta 4
eS

tand
ia d

er M
ran

Barinas
2

assif
Ser

Bucaramanga 1
6
Arauca

88

Medellin 52
25
83
89
ra
Cordille

Llanos Orientales
30
Basin (LLA)
Tunja

Yopal
Manizales 19

22
Bogota
Ibague

21
85
Villavicencio
47
ult
tral

Fa Structural cross-sections
ira
Cen

am
Alt used in the map view
restoration
La Macarena
Serrania de

35
Neiva

6 Structural cross-sections where


bed shortening was measured
71
(see Table 5.1 A and B).

Complementary not balanced


sif
as structural sections used to check
M
n Guyana Shield consistency of the map view
zo
ar
G restoration (see Table 5.1 C).

Figure 5.6: Location of the structural cross-sections used in the map view restoration of the Eastern Cordillera.
Numbers refer to Table 5.1.

194
Chapter 5

0 500 km

GUAJIRA BLOCK

ija
SANTA

Per
Bucaram
MARTA MARACAIBO

de
BLOCK BLOCK
s

n ia
de

BLOCK
An ault

r ra
anga Fa
9N
a

Se
F
rid no
LOMBIA Me Boco
ult
CENTRAL CO 8N

7N

ra
dille
Cor
6N

5N
tern
Eas

4N

3N

t
ul 2N
Fa
ira Detail in Fig. 5.10
m
ta 1N
Al

SOUTH AMERICAN PLATE 0

1S

Continental plate and


tectonic blocks

Deformed belts

Kinematic constraints
and location of selected
balanced cross sections
from literature
79 W 78 W 77 W 76 W 75 W 74 W 73 W 72 W 71 W 70 W

Figure 5.7. Palaeogene restoration of the major tectonic blocks of the NW corner of South America before the
Neogene Andean deformation. The co-ordinate grid represents the present day geographic grid restored to its
location before Neogene deformation.

They reconstructed the undeformed state of a single stratigraphic layer, currently folded and offset by
reverse faults. First they represent in map view the present, deformed state of the stratigraphic surface
to be restored as a mosaic of folded blocks bounded by faults. On this mosaic, blocks may be
contiguous, separated by gaps, or overlapping each other, according to the nature of the faults (strike-
slip, normal or reverse, respectively). The width of the gap or overlap between blocks is proportional to
the heave of the associated fault. Second, they unfold the blocks. Unfolding changes the shape of each
block and the width of overlaps at their boundaries. Third, they packed the unfolded blocks using rigid
translations and rotations to minimise the total area of gaps and overlaps. This yields a restored mosaic.
Because I lack structure contour maps at particular stratigraphic horizons in the EC, their method
cannot be applied directly.

195
Map View Restauration of Transpressional Basin Inversion in the Eastern Cordillera
Additionally, in the EC the erosion level is highly variable with the result that any particular
stratigraphic horizon may lie deep (without outcrop control) in areas of shallow erosion level or may be
completely eroded in areas of deep erosion (Fig. 5.9). Clearly in these cases there is no structural
control. Structural control for most of the sections is limited to surface geological maps and outcrop
data. Bed length preservation can only be applied to deep stratigraphic horizons in areas of deep
erosion level and to shallow stratigraphic horizons in areas of shallow erosion. In addition the location
of faults is known only from surface geological maps. In order to deal with these limitations, instead of
measuring bed lengths at a particular stratigraphic horizon, I measured both the undeformed and
deformed bed lengths, for each block, referred to the present-day topography surface and its restored
geometry before deformation (Fig. 5.9, see explanation below).

I used the following procedure:


(1) I compiled a detailed structural map of the EC (Fig. 5.3).
(2) I compiled as many as possible balanced structural cross-sections of the EC from available
literature (Fig. 5.6 and Table 5.1) and I checked their consistency in order to discard a few
number of clearly inconsistent sections (c.f. Roeder and Chamberlain, 1995 cross-section; see
discussion in section 6.2 of this chapter).
(3) I defined a number of blocks limited by faults (Fig. 5.8). I assumed that it is reasonable to divide
the region to be restored into a mosaic of blocks completely bounded by faults. Since natural
faults arrays are seldom totally connected, the mosaic has to be completed by introducing
artificial block boundaries. I extrapolated each fault trace until it meets another fault trace. To
allow some bending within blocks that are far from equant, I subdivided them into smaller ones,
using new artificial block boundaries. These artificial block boundaries were drawn on the basis
of shape of the blocks and where possible following minor faults or tight folds. Blocks so defined
are assumed to consist of folded layers. This assumption is reasonable if surface strains are
negligible.
(4) For each block I measured along each balanced cross-section the present-day deformed length
and the undeformed (pre-Andean) restored length (Figs. 5.4 and 5.9). I also measured the
horizontal displacement for each fault bounding a block (Figs. 5.4 and 5.9). Because there are no
accurate structural contour maps of particular stratigraphic reference levels, and the location of
faults is only known from surface geological maps, I measured both the undeformed and
deformed bed lengths, for each block, as the bed length between the location of the surface trace
of the faults limiting the block (Figs. 5.4 and 5.9). In the restored undeformed state I also
measured the horizontal separation of the present-day topographic surface, and its restored
geometry before deformation, across each boundary fault, and I distributed this length (Lf , see
Fig. 5.9) in two equal parts that were added to the restored length of the two blocks limited by
the fault (Figs. 5.4 and 5.9). In this way the restoration was done not at a particular stratigraphic
surface but at the present-day topographic surface. Although this method can produce small
differences as compared to a restoration at a particular stratigraphic horizon, the total amount of
shortening between pin lines is the same. Restorations made along different stratigraphic
horizons can differ but the total amount of shortening between pin lines will remain essentially
the same (Fig. 5.9).
(5) I unfolded each block separately (Fig. 5.9). I assumed that the stratigraphic horizons before
deformation were planar. This assumption is reasonable, as sedimentary irregularities are small
in comparison with the horizontal scale of the restoration. Obviously this restoration takes into
account all the deformation since deposition. In general, for non-cylindrical folds, it is
convenient to use automatic unfolding methods (Gratier et al., 1991). For the EC, where folds are
nearly cylindrical, I have performed this task manually. On a series of cross-sections, drawn
perpendicular to the main structures, I have conserved bed lengths. This process yielded an
extension of the blocks in map view parallel to the directions of the cross-sections. After each
block has been unfolded, it was assumed to be rigid.

196
Chapter 5

(6) I manually restored the relative location of each fault-bounded block to the pre-deformation state
fitting in map view the original undeformed lengths and minimising the distances across fault
separations (Figs. 5.9 and 5.10). I manually packed the blocks using rigid body translations and
rotations, visually minimising the total area of gaps or overlaps. This yielded a restored block
mosaic. Boundary conditions on displacements were conveniently introduced by keeping one
block stationary. I choosed the cratonic stable South American Plate as the stationary block: All
the displacements were referred to this.
(7) By comparing the material present-day co-ordinate grid and its location in the restored pre-
deformed (pre-Andean) state, I obtained an approximate field of finite horizontal displacements
(Fig. 5.11).

5.2. RESULTS

Figure 5.10 shows the map view restoration of the EC. The overall fit of the restored fault block map is
visually satisfactory. According to the restoration the amount of shortening of the EC is approximately
one half of the present-day width of the Cordillera. Both width and amount of shortening perpendicular
to the mountain range increase northward (Table 5.2). This is the result of clock-wise rotation (+6) of
the Central Colombia Block (Central Cordillera and Magdalena valley) relative to the stable South
American Plate. In addition to shortening normal to the axis of the Cordillera there are important strike-
slip components along the Bucaramanga Fault (left-lateral 100 km) in the north and the Altamira Fault
(right-lateral 30 km; Fig. 5.10).

ORIGINAL DEFORMED
SHORTENING L1/LO
LOCATION LENGTH L0 LENGTH
L0-L1 (KM) (%)
(KM) L1 (KM)
Northern Magdalena-Tablazo and Cocuy sub-basins 360 240 120 66.67
Southern Magdalena-Tablazo and Cocuy sub-basins 340 230 110 67.65
Northern Cundinamarca sub-basin 300 210 90 70.00
Southern Cundinamarca sub-basin 200 140 60 70.00
Neiva sub-basin (Upper Magdalena Valley-
100 90 10 90.00
Southern EC)

Table 5.2. Amount of shortening obtained from the map view restoration along several cross-sections perpendicular
to the EC. See Figures 2.1 and 5.8 for location of regions mentioned.

The Upper Magdalena Valley and south-eastern part of the Cordillera were displaced
transpressionally toward the NE with respect to the stable Llanos area (South American Plate) as
proposed by Branquet (1999) and Branquet et al. (1999b). Left-lateral displacement of the
Bucaramanga-Santa Marta Fault is accommodated by shortening and relative left-lateral displacement
of blocks in the western flank of the EC (Magdalena-Tablazo sub-basin, Fig. 5.10). Space constraints
from the restoration also suggest N-S shortening in the western flank of the Cordillera. In this
restoration I assumed that the Garzn Massif (Fig. 5.10) was displaced northeastward relative to stable
South American Plate, as suggested by Casero et al. (1995, 1997). The Serrana de La Macarena Block
(Fig. 5.10) represents the leading edge of the Garzn Massif systems of blocks. Results of the map
view restoration suggest that the SE flank of the EC (Cocuy and eastern Cundinamarca sub-basin, Fig.
5.10) was right-lateral transpressively deformed, but that the NW flank of the EC (Magdalena-Tablazo
sub-basin, Fig. 5.10) was left-lateral transpressively deformed.

197
Map View Restauration of Transpressional Basin Inversion in the Eastern Cordillera
1700

0 100 km

Buca
1600

rama

ija
d e Pe r
Maracaibo

n
Basin

ga F
1500

Sierra
ault
s
de

OCK
Serrania
A n
de Perija a
1400 id
er
A BL
ucas
M

Serran
MBI Cucuta

an L
de S
O

ia Mas
COL

Barinas
1300 basin
ania

Bucaramanga
L
Serr
TRA

Arauca

sif
CE N

1200
sin
ba
b-

Medellin
Su
lera

o
laz
Cordil

Llanos Orientales
Ta

ba cuy

1100 Basin (LLA)


a

Tunja
len

sin
Su d C
da

te
g

b-
er
Ma

Yopal
Inv
ed
ert

Manizales
Inv

1000
sin

Bogota
Ba
ca

Ibague
ar
m
ina
nd
Cu

900 Villavicencio
rn
ste
Ea
al
Centr

Continental plate and


ult tectonic blocks
Fa
800 Neiva ira
tam
a

Deformed belts
en

Al
La Macarena
al

Serrania de
gd
Va Ma

Kinematic constraints and


y
er
lle

location of balanced cross


pp
U

sections fron literature


700 ssif
Ma
n Colletta et al. (1990)
rzo
Ga Guyana Shield section (Figure 5.4)

800 900 1000 1100 1200 1300 1400


Figure 5.8. Present day mosaic of fault limited blocks used in the map view restoration of the EC and location of
balanced cross-sections from literature used in the restoration. Black line represents the balanced cross-section by
Colletta et al. (1990) shown in Figure 5.4.

198
Chapter 5

Fault bounded block section length after deformation


WNW
ESE

Portones Anticline Pin


Magdalena Pin

Landazuri Thrust
La Salina Thrust
Cimitarra Thrust

Armas Thrust

Horta Thrust
Kpt s
Ktr
Ku Kpt s Kpt s
Tm Tm
Tr Tr h Kl Kl Jg
Tr Tc ho Kpt s Jg
Tchu
Tchu u Tc Kpt s
T c ho Ku Ktr
Kl Jg
T c ho Ktr Jg
Kpt s Jg
Ku B
Ktr
Jg
B B
B B B

Fault bounded block section length before deformation

Portones Anticline Pin


Magdalena Pin

Lf
Cimitarra Thrust

Landazuri Thrust
La Salina Thrust

2
Armas Thrust

Horta Thrust
Lf
Tm
Tm Tr
Tr
Tc hu Tc hu
T c ho T c ho
Ku Ku Ku
Ku
Kl Kl
Kpt s Kpt s Kpt s
LEGEND B Ktr Ktr Ktr
Jg
Jg
Tm Mesa Fm. B Jg Jg

Tr Real Fm. B

Tchu Chuspas Gp. B


B B
Tcho Chorro Gp. 10 km
Ku Umir Fm. Horizontal and vertical scale
Kl La Luna Fm.
Kpts Paja, Tablazo and Simiti Fms.
Ktr Tambor and Rosablanca Fms.
Jg Giron Gp.
B Basament

Figure 5.9. Method used in the map view restoration of the EC. This example is a fragment of the western end of the
structural cross-section BB from Schelling (1994). Location of this section is shown in figure 5.8.

199
Map View Restauration of Transpressional Basin Inversion in the Eastern Cordillera

1 60 0

San
ta
Ma
rta
- B

Sierra de Perija
1500

uca
ram

Catatumbo
Sub-basin
s
de

ang
An

a
ida
er

Fau
14 00 M

lt
A BLOCK

Se
rra
nia

13 00
COLOMBI

Ma
sin

ssi
b-ba

f
lley

o Su

ra
Va

dille
laz
CENTRAL

1 20 0
Tab

Cor
len
gda

s in
tern
Ma

-b a
-
na

Eas
ale

Sub
dle

gd
Mid

Ma

ills
y

11 00
cu
ed

th
ert

o
Co

fo
Inv

os
ed

an
ert

Ll
Inv

1 00 0
s in
Su rn
ba
rca ste
b-

0 100 km
ma Ea

ult
ina ted

Fa
nd ver

e
Cu In

gu
Iba
a
V a l gdalen
l ey
Ma

900
er
Upp

Continental plate and


La Macarena

tectocnic block
Serrania de

800 Deformed belts


Kinematic constrains and
if location of balanced cross
ss
Ma sections from literature.
n
700 zo
Ga
800 900 1000

Figure 5.10. Palaeogene restoration of the mosaic of fault limited blocks before Neogene deformation. Grid
represents the present day plane co-ordinate grid restored to its location before Neogene deformation. Thick black
broken line represents the Colletta et al. (1990) balanced cross-section shown in figure 5.4.

200
Chapter 5
5.3. DISPLACEMENT OF BLOCKS RELATIVE TO STABLE SOUTH AMERICA AND
ROTATIONS ABOUT VERTICAL AXES

By comparing the material present-day co-ordinate grid and its location in the restored pre-
Andean state, I obtained an approximate field of finite horizontal displacements (Fig. 5.11). The
displacement vectors increase in length from east to west, because they are referred to the eastern
cratonic South American Block, which is stationary. Several regions can be distinguished according to
their displacement vectors (Fig. 5.11)

(1) The Garzn Massif displacement vectors are toward NE and their magnitude decreases from
about 30 km in the SW to almost zero in the Serrana de la Macarena. Rotation of these blocks
about a vertical axis is practically negligible. This region is separated from the southern EC and
Upper Magdalena Valley by the Altamira Fault.
(2) The Upper Magdalena Valley-southeastern flank of the EC (eastern Cundinamarca sub-basin and
Cocuy sub-basin) and southern Central Cordillera (south of the Ibagu Fault) displacement
vectors are toward NNE in the Upper Magdalena Valley and gradually change toward E in the
Cocuy area. Their magnitude gradually increases NW to a maximum of 90 km in the Girardot
sub-basin of the Upper Magdalena Valley. Blocks in this region show clockwise rotation about a
vertical axis. These rotations decrease toward NE to a minimum in the Cocuy area.
(3) The western flank of the EC (Magdalena Tablazo sub-basin and western part of Cundinamarca
sub-basin) and northern part of Central Cordillera (north of the Ibagu Fault) displacement vector
azimuths are ESE but gradually change toward SE in the Magdalena-Tablazo sub-basin close to
the Bucaramanga Fault. Displacements in this region increase northward as this region has
rotated in a clockwise sense about a vertical axis close to its southern tip. Displacement also
increases westward. Between the regions 2 and 3 I assume a left-lateral shear zone.
(4) The Santander MassifSierra de Perij region displacement vectors are toward E in its southern
part, but change to ENE in its northern part. The displacement magnitude increases northward as
this region has been rotated over a small angle about a pole close to its southern tip. Strong
changes in the displacement field between regions 3 and 4 are accommodated by the mostly left-
lateral strike-slip relative displacement of the Santa Marta-Bucaramanga Fault.
(5) The Mrida Andes-Maracaibo Basin has been displaced toward the east. Displacement increased
westward within the Mrida Andes. Rotation of blocks about a vertical ax is negligible in this
region. A complex zone is located between the Cocuy sub-basin, the Santander Massif and the
Mrida Andes.

6. DISCUSION

6.1. ADVANTAGES OF THE MAP VIEW RESTORATION

As discussed in the results, the method I applied is useful to constrain shortening and strike-slip
displacement estimates as well as to detect strike-slip motions and rotations about vertical axes not
revealed by balancing cross-sections. Obviously results should be checked with other independent
methods.

6.2. COMPARISON WITH PREVIOUS SHORTENING ESTIMATES

Table 5.3 shows a comparison between some shortening estimates from literature and those
obtained in the map view restoration. The amount of shortening perpendicular to the EC obtained in the
map view restoration is greater than the shortening estimates from Cooper et al. (1995), Linares (1996)
and most of the balanced cross-sections of Schelling (1994). Map view shortening estimates are,
however, smaller than those estimates from Colletta et al. (1990), Dengo and Covey (1993) and Roeder
and Chamberlain (1995). The smallest difference 9% is for the Colletta et al. (1990) section. The
shortening figure of Dengo and Covey (1993) seems to be overestimated. The balanced-cross-section

201
Map View Restauration of Transpressional Basin Inversion in the Eastern Cordillera

16 0 0

Sierra de Perija
1500

5
s
de
An
a
rid
1 40 0 Me
K
BLOC

13 00

Santa
4
MBIA

nder
COLO

Mas

12 0 0
sif
RAL
CENT

as na-
su gdale
in
b-b
Ma

in
Ta rted
zo

as
11 00
e
bla

b-b
Inv

su
y
cu
Co
ed
ra
ille

ert

1 0 00
rd

Inv
Co
rn
ste
Ea

ult
Fa 2
lley

e
gu
Iba
Va

900
a
len
gda
Ma
per

La Macarena
Up

Sierra de

800

1
if
700 ss 00 100 Km
100 km
Ma
n
rzo
Ga
800 900 1000 1100 1200 1300

Figure 5.11. Displacement field during Neogene deformation of the EC. Each vector connects the present day co-
ordinate grid with its position before Neogene deformation. Numbers represent regions discussed in section 5.3 of
this chapter.

202
Chapter 5
interpretation they propose contains a hypothetical low-angle mid crustal detachment that extends
westward to the Central Cordillera up to the western edge of the South American plate. According to
them, during Miocene to Pliocene shortening in the EC, the Central Cordillera was carried passively up
this mid-crustal detachment, decapitating and thrusting a Jurassic graben below the Magdalena Valley
(Fig. 5.3) and generating regional uplift of the Central Cordillera and erosion of its sedimentary cover.
A decapitated graben below the basement rocks of the Magdalena Valley seems to be very speculative
and it is not supported by evidence. Roeder and Chamberlain (1995) also overestimated the amount of
shortening. Their shortening estimation seems to be unreal. Their cross-section interpretation is based
on the questionable assumption that the whole Cretaceous sedimentary record represents a thermal sag
basin fill after Jurassic rifting (ignoring earlier work by Fabre, 1987, who demonstrated the existence of
an Early Cretaceous extensional basin).

Literature This work Shortening


Reference %
Shortening Shortening Difference
L1/l0 L1/L0 (km) *
Differen
(km) (km) #
70.31
Colletta et al. (1990) 105 66% 95 -10.00 -9.52
%
69.57
Dengo and Covey (1993) 150 40% 105 -45.00 -30.00
%
72.58
Cooper et al. (1995) 68 80% 85 17.00 25.00
%
Roeder and Chamberlain 69.57
230 58% 105 -125.00 -54.35
(1995) %
69.57
Jones (1995) 185 58% 105 -80.00 -43.24
%
71.43
Linares (1996) 50 20% 80 30.00 60.00
%
18 to 69.6 to 11.11 to
Schelling et al. (1997) 45 to 65 50 to 108 5 to 43
26 % 78.2 % 66.15
L0: Restored undeformed length of the section; L1: Present day deformed length of the section
* Difference = shortening this work shortening literature; # % Difference relative to shortening literature as
100%

Table 5.3. Comparison of amount of shortening obtained from the map view restoration and some balanced cross-
sections from literature.

6.3. COMPARISON WITH OUTCROP STRUCTURAL STUDIES, STRESS INFERRED


FORM BOREHOLE BREAKOUT DATA AND PLATE MOTIONS FROM
GEOPHYSICAL DATA

6.3.1. Shortening perpendicular to the regional structural grain of the EC

A general shortening perpendicular to the regional structural grain of the EC recognized in the
map view restoration (compare Figs. 5.8 and 5.10) is supported by structural kinematic fault and fault
striae data from the EC (Figs. 5.12 to 5.16, Cobbold et al., 1988; Cardozo and Ziga, 1995; Daz and
Sotelo, 1995; Kammer, 1996, 1999b), by the orientation of systematic conjugate joints normal to
bedding planes interpreted by Kammer (1999b, 2000) as originated before folding (Fig. 5.15) and by
the orientation of stylolites and vertical joints of recent origin (Fig. 5.16, Mojica and Sheidegger, 1981;
1984; Mojica 1983 in Torres, 1992; and Mojica, 1985). Also a general shortening perpendicular to the
regional structural grain of the EC is supported, the present stress orientation as inferred from borehole
breakout data (Fig. 5.17, Castillo and Mojica, 1990; Ocha and Ponguta, 1991 and Torres, 1992) as
well as from the orientation of compressional axis of focal mechanisms solutions of earthquakes in the
upper 50 km of the crust (Fig. 5.18, Daniels, 1991). Although fault striae data from the EC
approximately suggest shortening perpendicular to the mountain range in detail a clockwise departure
from the perpendicular to the structural trend (Fig. 5.14) may be interpreted as a result of a collision of
the western block of the Bucaramanga Fault with the eastern flank of the EC (Kammer, 1999b).

203
Map View Restauration of Transpressional Basin Inversion in the Eastern Cordillera

Mojica and Sheidegger (1981) measured the orientation of recent vertical joints with smooth
surfaces in outcrops (Fig. 5.16). Their data shows large dispersion. They found two sets: N 1-8 E and
N 91-100 E forming angles close to 90. The bisectrix of these sets corresponds theoretically to
principal stress directions with azimuths N 45 54 E (approximate parallel to the EC) and N135
144 E (approximately perpendicular to the EC). The identification of the maximum and minimum
stress is uncertain from the joint measurements alone, their results show consistent orientation of the
principal stress axis, but different identification of the maximum stress axis at neighbour locations (Fig.
5.16). This inconsistency could be eliminated if all orientations perpendicular to the EC are
reinterpreted (c in Fig. 5.16) as those of maximum stress according to inferences from horizontal
stylolites and other structural data. Mojica (1983 in Torres, 1992) and Mojica and Scheidegger (1984)
studied the orientations of horizontal stylolites, fold axes, deformed fossils, recent joints and fault plane
solutions throughout the Colombian Andes and found NW-SE maximum horizontal compression
(approximately 140) with local variations of E-W and WNW-ESE. Mojica (1985) measured tectonic
stylolites in the EC between SW of the Neiva City and the southern termination of the Bucaramanga
Fault. Stylolites are oriented NW-SE with subordinate NE-SW and E-W directions at few outcrops.
These structural data suggest an approximate shortening perpendicular to the structural trend of the EC.
From borehole breakout analysis of wells in the main sedimentary basins of Colombia Torres
(1992) suggested a present-day stress regime with the maximum horizontal stress oriented WNW-ESE
to NW-SE (Fig. 5.17). She distinguished a northern (above 6 latitude, Guajira, Maracaibo, Catatumbo
and Lower and Middle Magdalena basins; Fig. 1.1), central (between 4 and 6 latitude, Llanos and
Upper Magdalena basins) and southern (between 0 and 1 latitude, Putumayo Basin; Fig. 1.1) sub-sets
with mean maximum horizontal stress orientations of 117, 111 and 133 respectively. Although this
reflects small changes in the horizontal stress orientation with latitude, the horizontal stress orientation
trend perpendicular to the EC is regionally preserved. Borehole breakouts indicate a horizontal
principal stress of 138 for the Llanos wells and 112 for the Middle Magdalena Valley (Castillo and
Mojica, 1990) and NW-SE compression in the Upper Magdalena Valley (Ocha and Ponguta, 1991,
Fig. 5.17).
Daniels (1991) compiled the orientation of compressional axes from focal mechanism solutions
of earthquakes in the upper 50 km of the crust (Fig. 5.18). His results indicate ENE-SWW compression
in the northern part of the EC and E-W to WSW-ENE in the southern part of the EC and Ecuador.
As mentioned before, all these data support the shortening approximately perpendicular to the
EC suggested by the map view restoration.

6.3.2. Conjugate (?) left-lateral and right-lateral strike-slip faults

Left- and right-lateral strike-slip fault zones suggested by the map view restoration are
supported by outcrop structural data.
Left-lateral strike-slip motion of the Bucaramanga Fault (Fig. 5.3) is supported by kinematic
analysis of lineation data on the fault (Figs. 5.12 and 5.13). Acosta et al. (2000) interpreted from these
data that the Bucaramanga and Suarez faults (Fig. 5.3) are steep dipping left-lateral strike-slip faults.
Two more set of planes were measured in the Bucaramanga Fault: one set trends sub-parallel to the
main structure and is interpreted as riedels. A second set is formed by steep-dipping right-lateral faults;
these are interpreted as antithetic riedel shears. A similar pattern was found by Meier et al. (1987) in
the Tchira area, which is the link between the EC of Colombia and the Mrida Andes of Venezuela
(Fig. 5.3).

204
Chapter 5

Serrania
de Perija
0 100 km
Maracaibo
Basin

s
de
An
a
e rid
M

Lucas
de San Cucuta

7
Barinas
Serrania

basin
Bucaramanga

1 Arauca
V) a
(M alen
d
lle ag
Va M
y

Medellin
le
idd

2
illera

Llanos Orientales
4 3 Basin (LLA)
Cord

Tunja
Yopal
)C
(E
A
ER

Manizales
LL

5
DI
R

Bogota
CO

Ibague
6
N
ER
ST
EA

Villavicencio
8 Guyana Shield
al

alley

OUTCROP STRUCTURAL DATA


Centr

(Kinematic indicators)
a V

Ne iva
alen

1. Cobbold et al.(1988)
La Macarena
sif

Serrania de
agd

Mas

2. Acosta (2000)
er M

3. Kammer (1996); Cardozo and Zuniga (1995)


zon
U pp

4. Diaz and Sotelo (1995)


Gar

5. Cardozo-Puentes (1989)
6. Kammer and Mojica (1995)
7. Meier et al. (1987)
8. Kammer (2000)

Figure 5.12. Location of detailed outcrop structural studies of the EC, including kinematic indicators and fault striae
from literature used to compare with the results of the map view restoration.

205
Map View Restauration of Transpressional Basin Inversion in the Eastern Cordillera

0 100 km Serrania
de Perij a
Marac aibo
Basin

s
de
An
a
id
er
s
L uc a

Cucuta M

Bu c
n
e Sa

Barinas

a ra
ma
nia d

basin
n ga
Bucaramanga
S e r ra

F
Arauca
F
na

F
ali
os F

rez
S

Sua
La
ma n

Medellin

2
H er
ra
i ll e

4 3
Do s
rd

Tunja
Co

aF
nd

aF
Ho

Yopal
uim
Bi t

Manizales

Bogota
ue F
Ibag
Ibague

N A
R
S TE L E R Villavicencio
L Llanos Orientales
EA DI
l
ra

R Basin (LLA)
CO
nt
Ce

l le y
Va

La Macarena
na

Serrania de
if
a le

ss

Neiva
gd

Ma

Guyana Shield
Ma

n
r zo
per

OUTCROP STRUCTURAL DATA


Ga
Up

2. Acosta et al.(2000)
3. Kammer (1996); Cardozo and Zuniga (1995)

4. Diaz and Sotelo (1995)


|
Figure 5.13. Kinematic indicators and other outcrop structural data in the EC from literature.

206
Chapter 5

Serrania
de Perija Maracaibo
Basin

s
de
An
ida
er
M
0 100 km

alley
Cucuta

V
Barinas

lena
basin

a
agd
le M

Bucaramanga
Arauca
Midd

A
illera

Medellin
ER
ILL
Cord

RD
CO

Tunja
RN

Yopal
STE

Manizales
EA

Bogota

Ibague

Villavicencio LLANOS ORIENTALES


BASIN (LLA)
ey

Guyana Shield
Vall
l
ntra

La Macarena
Serrania de
alen
Ce

Neiva OUTCROP STRUCTURAL DATA


agd

(Kinematic indicators)
er M

sif 8. Kammer (2000)


as
Upp

M
n
r zo Reverse fault
Ga
Normal fault

Strike-slip fault

Figure 5.14. Kinematic indicators from fault striae in the EC. At each locality solid arrows indicate shortening
direction, open arrows indicate extension direction and lines intermediate direction. Sterographic plots (equal area
lower hemisphere) indicate also the bedding/foliation orientation. Relative fault movement is indicated at the
localities where most of measured fault planes with that sense predominate (modified from Kammer, 1999b)

207
Map View Restauration of Transpressional Basin Inversion in the Eastern Cordillera

Serrania
de Perija Maracaibo
Basin

ucas
s
de

an L
An
a
rid

eS
0 100 km Me

d
ania

lley
Serr
Cucuta Barinas

a Va
basin

alen
ag d

Bucaramanga
le M

Arauca
Midd
illera

Medellin
A
ER
Cord

ILL
RD
CO

Tunja
RN

Yopal
STE

Manizales
EA

Ibague
Bogota
LLANOS ORIENTALES
Villavicencio BASIN (LLA)
ey

Guyana Shield
Vall
l
ntra

La Macarena
Serrania de
alen

sif
Ce

M as

Neiva
ag d

OUTCROP STRUCTURAL DATA


er M

zon

(Joint analysis)
Gar
Upp

8. Kammer (2000)
Reverse fault

Normal fault

Strike-slip fault

Figure 5.15. Inferred shortening and extension directions from joints measured at outcrops. At each locality solid
arrows indicate shortening direction, open arrows indicate extension direction and lines intermediate direction (from
Kammer, 1999b).

208
Chapter 5

ORIENTATION OF RECENT
VERTICAL JOINTS AND
INFERRED STRESS ORIENTATION

de Perija
Serrania
(from Mojica and Sheidegger, 1981)
Maracaibo
Basin
Orientation of vertical joints.

Inferred compressive stress axis.


Average of the region closed by a rectangle.

Orientation corrected from the Mojica and Scheidegger


c (1981) figure because that maximum stress is inconsistent s
de
with horizontal stylolites reported by Mojica and An
Scheidegger (1984) ida
er
M

ucas
Barinas
basin
Cucuta
an L
c
de S

c
ania

c
Bucaramanga Arauca
y
Serr

alle

c
aV

c
len
illera

gda

Medellin
Ma

A
Cord

c
ER
dle

ILL
Mid

RD
CO

c Tunja
Yopal
RN

Manizales c
STE
EA

c
Ibague Bogota
LLANOS ORIENTALES
Villavicencio BASIN (LLA)
y
alle

Guyana Shield
V
l

len
ntra

La Macarena
Serrania de
gda
Ce

Neiva c
Ma
er

si f
Upp

as
M
n
c r zo
Ga 0 100 km

Figure 5.16. Orientation of recent vertical joints and inferred stress orientation (modified from Mojica and
Sheidegger, 1981).

209
Map View Restauration of Transpressional Basin Inversion in the Eastern Cordillera

Serrania Maracaibo
de Perija Basin

s
de
An
a
id
s

er
Luca

M
Cucuta
an
de S

Barinas
Bucaramanga basin
ania
Serr

lley

Arauca
Va
na
ale

)
C
(E
gd

Medellin
A
Ma

R
LE
ra

L
DI
le
d il l e

dd

R
O
Mi

C LLANOS ORIENTALES
Cor

N Tunja
R BASIN (LLA)
S TE
Yopal
EA
Manizales

Bogota

Ibague

Villavicencio Guyana Shield

0 100 km
tral
Cen

La Macarena
Serrania de
if
ss
Ma

Neiva
n
rzo

STRESS FIELD ORIENTATION ( 1 )


Ga

INFERRED FROM BOREHOLE BREAKOUT


PUTUMAYO BASIN DATA
Decreassing quality rank
A
B Torres (1992)
C
D Castillo and Mojica (1990)
Ochoa and Ponguta (1991)

Figure 5.17. Orientation of present day stress field inferred from borehole breakout studies.

210
Chapter 5
In this area Meier et al. (1987) recognized a complex block mosaic with: (1) NE-trending
dextral strike-slip faults; (2) SE-trending sinistral strike-slip faults; and (3) N-S compressive thrust
faults. In the southern part of the study area right-lateral strike-slip motion of the Altamira Fault and the
southern part of the EC is supported by the Cobbold et al. (1988) and Casero et al. (1995, 1997)
interpretations that the southern EC is a zone of thrusting with minor right-lateral wrenching, mainly
upon synthetic faults and right stepping folds and thrusts. In the Upper Magdalena Valley and south-
eastern part of the Cordillera transpressional displacement toward the NE with respect to the stable
LLA area, suggested by the map view restoration, is supported by outcrop and seismic data from
Branquet (1999) and Branquet et al. (1999b), as well as by geophysical data by Pennington (1981, Fig.
6.6 Chapter 6), Daniels (1991, Fig. 5.18) and Mora (1995, in Acosta et al., 2000, Figs. 5.2).
Using field and seismic data Branquet (1999) and Branquet et al. (1999b) interpreted a section
of the eastern flank of the Cordillera. They identified en-chelon folds in the hanging-wall of major
basement-involved faults and they interpret basement pop-up structures limited to the east by convex-
upward reverse faults as positive flower structures. They proposed that these structures resulted from
dextral transpressional inversion of an Early Cretaceous half-graben during the Andean orogeny. They
also mentioned shallow-focus earthquakes, shifting rivers and the morphology of the Llanos foothills
(Fig. 2.1 for location) as evidence of present-day dextral transpression. They also interpreted the
Quetame Massif (Fig. 2.1) as an en-chelon basement pop-up structure uplifted during tansspressional
Andean deformation. This interpretation is also supported by earthquake focal mechanism solutions
compiled by Pennington (1981, Fig. 6.6 Chapter 6) and Daniels (1991, Fig. 5.18). Global Positioning
System data suggest a NE displacement of the Eastern Andes Block relative to the stable South
American Plate (Llanos; Mora, 1995, in Acosta et al., 2000, Fig. 5.2).
Shortening and relative left-lateral displacement of blocks in the western flank of the EC
(Magdalena-Tablazo sub-basin) recognized in the map view restoration are supported by kinematic
analysis of lineation data on the main faults on the western EC foothills and the Middle MV carried out
by Acosta et al. (2000, Figs, 5.12 and 5.13). These authors suggested that: (1) the Bituima-La Salina
and Chucur faults (Fig. 5.3) are represented by a series of steep dipping left-lateral to oblique faults
with a reverse sense; (2) The Cambras and Dos Hermanos faults (Fig. 5.3) are oblique faults with left-
lateral and reverse sense that form a duplex. Cobbold et al. (1988) also suggested that the northern EC
is a zone of convergent left-lateral wrenching.
The existence of two conjugate (?) or pseudo-conjugate (?) sets of NNW-SSE to NW-SE left-
lateral strike-slip faults and NE-SW to E-W right-lateral strike-slip faults, is also confirmed from
kinematic indicator fault striae measurements and other outcrop structural observations in the EC (Figs.
5.12 to 5.15, Cobbold et al., 1988; Cardozo-Puentes, 1989; Cardozo and Ziga, 1995; Kammer,
1999b) and the Tchira area (Fig. 5.3; Meier et al., 1987). In the Santander Massif (along the
Bucaramanga-Ccuta road) Cobbold, et al. (1988) determined the Bucaramanga Fault (Fig. 5.3) and
other prominent faults with the same trend (340) as reactivated Precambrian structures. These faults
have oblique striae indicating dominantly left-lateral wrenching as well as minor thrusting. Another
fault family striking 020 to 060 shows right-lateral thrusting. They hypothesised that the northern EC
is a zone of convergent left-lateral wrenching. According to these authors the southern EC is a zone of
thrusting with minor right-lateral wrenching, mainly upon synthetic faults and right stepping folds and
thrusts. They speculated that the entire EC is at the restraining intersection of right- and left-lateral
wrenches.

6.3.3. Clockwise rotation of the Central Cordillera, Magdalena Valley and western flank of the
Eastern Cordillera

The clock-wise rotation of the Central Colombia Block (Central Cordillera), Magdalena Valley
and western flank of the EC: Magdalena Tablazo sub-basin and western part of Cundinamarca sub-
basin relative to the stable South American Plate inferred from the map view restoration is in
agreement with Acosta et al. (2000) interpretation. Acosta et al. (2000) using kinematic analysis of
lineation data from the main faults on the western EC foothills and the Middle MV as well as regional
interpretation of structures in the northern Central Cordillera and MV interpreted a left-lateral shear

211
Map View Restauration of Transpressional Basin Inversion in the Eastern Cordillera

0 500km

South Caribbean deformed Belt


CARIBBEAN
PLATE

PANAMA
BLOCK Venezuela

A
CHOCO

ER
BLOCK
ILL
RD
CO
IA

ER
MB

Colombia
ST
LO

EA

NAZCA
CO

PLATE
RN
TE
ES
SW
A IN

SOUTH AMERICAN PLATE


RR

Brazil
TE

STRESS FIELD (P-AXES) INFERRED


NIC

FROM FOCAL MECHANISMS IN THE


EA

UPPER 50 KM OF THE
OC

Ecuador
CRUST
Straight lines indicate the trend of compression axes and length
inversely proportional to the dip of the axis (from Daniels, 1991)

Peru ESTRUCTURES
Thrust fault
Strike-slip fault
Normal fault

Figure 5.18. P-axes from earthquake focal mechanisms in the upper 50 km of the crust. Lines indicate the trend of
the axis and the length is inversely proportional to the dip of the axis away from horizontal (from Daniels, 1991).

zone between the Bucaramanga and Uramita faults formed as a result of indentation of the Choc
Block (Figs. 5.1 and 5.2), resulting in transpression and clockwise block rotation (Acosta et al., 2000).

6.3.4. Deformation of the EC by transpressive basin inversion

Results of the map view restoration suggest that the SE flank of the EC (Cocuy and eastern
Cundinamarca sub-basin) was right-lateral transpressively deformed but the NW flank of the EC

212
Chapter 5

(Magdalena-Tablazo sub-bassin) was left-lateral transpressively deformed (see for example thick
broken black line Collettas section in Fig. 5.10). These results are in agreement with De Freitas et al.
(1997) who interpreted that the EC comprises: (1) A northern segment produced by the inversion of the
Tablazo and Cocuy grabens separated by the Santander-Floresta palaeo-high. They recognized the
Boyac and Soapaga faults as inverted normal faults (Fig. 2.28 Chapter 2); (2) A southern segment
(Cundinamarca sub-basin) where the junction of the two northern depocenters implies accommodation
of en chelon normal faults separated by strike-slip faults. They suggest that during basin inversion
the western part of the Cundinamarca sub-basin was affected by N-S sinistral transpression, while the
eastern part was affected by E-W dextral transpression.

6.3.5. Minor structures recognized in outcrops, but not in the map view restoration

Three structural features recognized from outcrop studies but not recognizable from the map
view restoration are (Figs. 5.13, 5.14 and 5.19): (1) Local NW-SE to N-S extension associated with
minor normal faults crossing obliquely to perpendicularly the structural grain of the EC. These local
minor features have been recognized by Cardozo and Ziga (1995) and Kammer (1996, 1999b) in the
eastern border of the Floresta Massif and by Cardozo-Puentes (1989) in the western part of the
Cundinamarca sub-basin; (2) E-W to WNW-ESE-trending minor left-lateral strike-slip faults cross
cutting the structural grain of the EC and thus of later (recent) origin. These faults have been
recognized by Kammer (1999b), by Acosta et al. (2000) in the western flank of the EC and Middle
MV, and by Cardozo-Puentes (1989) in the Sabana de Bogot region; (3) E-W -trending right-lateral
minor strike-slip faults cross cutting the major structures recognized by Cardozo and Ziga (1995) in
the eastern flank of the Floresta Massif. Probably these structures are small enough to be detected in the
regional map view restoration.

6.4. COMPARISON OF DISPLACEMENT OF BLOCKS RELATIVE TO STABLE SOUTH


AMERICA WITH PLATE VELOCITY VECTORS FROM GEOPHYSICAL DATA

Displacement of blocks relative to the stable cratonic South American Plate obtained from the
map view restoration (Fig. 5.11) can be directly compared with velocity vectors determined from
Global Positioning System measurements relative to South American plate (Kellogg et al.,1985;
Freymueller et al., 1993; Mora, 1995, in Acosta et al., 2000; Fig. 5.2).
Northeastward displacement vectors of the Garzn Massif relative to the stable South American
Plate suggested by the map view restoration (Fig. 5.11) are in notable agreement with velocity vectors
obtained from GPS measurements (Kellogg et al., 1985; Freymueller et al., 1993; Mora, 1995, in
Acosta et al., 2000, Fig. 5.2).
Northeastward displacement vectors of the Upper Magdalena Valley-southeastern flank of the
EC (eastern Cundinamarca sub-basin and Cocuy sub-basin) and southern Central Cordillera (south of
the Ibagu Fault, Fig. 5.11) are in agreement with velocity vectors relative to stable South America
from GPS measurements (Kellogg et al., 1985; Freymueller et al., 1993; Mora, 1995, in Acosta et al.,
2000, Fig. 5.2). However, the gradual change toward E and decrease in magnitude in the Cocuy area
(Fig. 5.11) are in disagreement with the velocity vectors from GPS data (Fig. 5.2). The WNW-ESE to
E-W orientation of the regional stress field inferred from recent joints and stylolites (Mojica and
Scheidegger, 1981; Mojica and Scheidegger, 1984; Mojica, 1985; Mojica, 1992, in Torres, 1992; Fig.
5.16), borehole breakout studies (Castillo and Mojica, 1990; Ocha and Ponguta, 1991 and Torres,
1992, Fig. 5.17) and focal mechanisms solutions of earthquakes in the Andes Block (Fig. 5.18) support
an internal deformation of the EC. Such an internal deformation of the EC is in agreement with the
gradual change toward the east in the displacement vectors obtained from the map view restoration
(Fig. 5.11). In contrast, earthquake focal mechanism solutions for the easternmost Andes boundary
faulted margin indicate an important dextral strike-slip component (Pennington, 1981, Fig. 6.6 Chapter
6) which is in agreement with the Northeastward velocity vectors obtained from GPS data (Kellogg et
al., 1985; Freymueller et al., 1993; Mora, 1995, in Acosta et al., 2000, Fig. 5.2). In general for the

213
Map View Restauration of Transpressional Basin Inversion in the Eastern Cordillera

0 100 km Serrania
de Perija

Maracaibo
Basin

s
de
An
a
id
er
M

Buca
Cucuta

ram
cas

Barinas
San Lu

anga
basin

F
nia de

Bucaramanga
Serra

Arauca
F
rez
Sua
F

aF

Medellin
nos

lin
Sa
ma

La
Her
ra
l le

Dos
rdi

Llanos Orientales
Co

F
da

Tunja Basin (LLA)


Hon

Yopal
RA
F
a

LE
im
tu

IL

Manizales
RD
Bi

CO
E RN

ue F
ST

Ibag
EA

Bogota
Ibague

Villavicencio
l a
nt r

Guyana Shield
Ce

y
alle
aV

La Macarena

Neiva
Serrania de
alen

if
ass
agd

nM
er M

rzo
Ga
Upp

OUTCROP STRUCTURAL DATA


2. Acosta et al. (2000)

Figure 5.19. Kinematic indicators of northwest-trending left-lateral strike-slip faults (from Acosta et al., 2000).

214
Chapter 5

study area, the conflict between the displacement vectors from the map restoration (Fig. 5.11) and the
velocity vectors from GPS data (Figs. 5.2), can be resolved if we consider that the displacement vectors
from the map view restoration represent mostly the Neogene Andean deformation and the GPS velocity
vectors represent the present-day movement of the Andes relative to stable South American Plate.
Probably the Andes Block as a whole is at present moving with a dominant right-lateral strike-slip
component along the faults in its easternmost boundary. This strike-slip component was probably much
less important during the whole period of Neogene Andean deformation as suggested by the map
restoration.

6.5. ANDEAN NEOGENE DEFORMATION, BASIN INVERSION OF THE EASTERN


CORDILLERA AND LITHOSPHERE RHEOLOGY

Andean deformation of the EC and its northern prolongations, the Sierra de Perij and the
Mrida Andes, coincide with the location of Mesozoic extensional basins (Figs. 2.13, 2.15 and 2.17;
Chapter 2). These extensional basins were more intensely deformed as mountain belts than the more
rigid blocks with Neogene sedimentary basins surrounding them (Figs. 1.1, 5.5 and 5.7). Young rifts
are characterized by weak lithosphere (e.g. Ebinger et al., 1989, in Burov and Diament, 1995) due to
thermal weakening and necking of the lithosphere. Weak thermal destabilized lithosphere in the area of
the former extensional basins (Jurassic to Early Cretaceous rifts) was prone to deformation.
Transpressive tectonic stresses generated inversion of these former extensional basins to generate these
mountain ranges. Under transpressive tectonic stresses the major normal fault systems originally
delimiting the extensional basins were inverted (Fig. 5.4). In the external western and eastern foothills
of the EC short-cuts within the sedimentary cover generated thrust systems (Fig. 5.4).

7. CONCLUSIONS

The Northern Andes of Colombia and Venezuela are located in a broad zone of deformation
resulting from the interaction of the South American, Nazca and Caribbean plates. Such interaction has
fragmented the NW corner of the South American Plate into a number of micro-plates or tectonic
blocks: Guajira, Santa Marta, Maracaibo and Central Colombia (Colombian Central Cordillera).
Relative movements of the tectonic blocks resulted in deformed belts with a combination of
compressional thick/thin-skinned thrusting, folding and strike-slip faulting which form several
mountain ranges or strike-slip fault zones in the Northern Andes. I have manually restored in map view
a mosaic of fault-bounded blocks using all the kinematic constraints available in literature, such as
strike-slip displacement estimates and amounts of shortening from available balanced cross-sections. In
order to simplify the model I have used a limited number of blocks. This technique is useful to
constrain shortening and strike-slip displacement estimates as well as to detect strike-slip motions and
rotations about vertical axes not revealed by balancing cross-sections. According to this regional map
restoration, the following relative movements of regional blocks leading to deformation belts were
interpreted: (1) Eastward convergence and clockwise rotation of the Central Colombia tectonic block
relative to South America resulted in transpression of the EC, with northward increasing shortening in
this mountain range; (2) Eastward convergence and left-lateral strike-slip movement of Central
Colombia relative to Maracaibo produced the northern part of the EC in the Santander Massif region;
(3) Northeastward convergence of the Santa Marta Block relative to Maracaibo generated the Perij
Mountain Range; and (4) Southeastward convergence and right-lateral strike-slip motion of the
Maracaibo Block relative to South America generated the Mrida Andes of Venezuela.
A more detailed map view restoration of the EC of Colombia was done using balanced cross-
sections from the literature. According to the restoration the amount of shortening during Andean
deformation is approximately one half of the present-day width of the Cordillera. Both width and
amount of shortening increase northward. Results of the map view restoration are in general supported
by kinematic indicators (fault striae) and other outcrop structural data, as well as by the orientation of
stress field inferred from recent vertical joints, stylolites, borehole breakout data and the focal
mechanism solutions of upper crustal earthquakes. These results indicate: (1) ENE-WSW shortening

215
Map View Restauration of Transpressional Basin Inversion in the Eastern Cordillera

perpendicular to the regional structural grain of the EC; (2) Conjugate (?) or pseudo.conjugate (?) left-
lateral and right-lateral strike-slip faults; (3) Clockwise rotation of the Central Cordillera, Magdalena
Valley and western flank of the EC; (4) The SE flank of the EC (Cocuy and eastern Cundinamarca sub-
basin) was right-lateral transpressively deformed; (5) The NW flank of the EC (Magdalena-Tablazo
sub-basin) was left-lateral transpressively deformed; (6) Andean deformation generated the EC through
transpressive inversion of Mesozoic extensional basins. The western part of the EC was affected by N-
S sinistral transpression, while the eastern part was affected by E-W dextral transpression (see for
example thick broken black line Collettas section in Fig. 5.10).
Comparison of displacement of blocks relative to stable South America obtained from the map
view restoration, with plate velocity vectors from geophysical data, suggest that the Andes Block as a
whole is at present time moving with a dominant right-lateral strike-slip component along the faults at
its easternmost boundary. This strike-slip component probably was less important during the whole
period of Neogene Andean deformation, as suggested by the map view restoration.

216
CHAPTER 6

RHEOLOGICAL EVOLUTION OF THE LITHOSPHERE OF THE EASTERN CORDILLERA


AND HYPOTHESES ABOUT ITS DEEP STRUCTURE

1. INTRODUCTION

Two major controls on lithosphere dynamics of basin formation and inversion are: (1)
lithosphere rheology and (2) stresses affecting the lithosphere (Ziegler et al., 1995, 1998). The aim of
this chapter is to constrain the Meso-Cenozoic lithospheric scale tectonic evolution of the EC in terms
of the interaction of rheology of the lithosphere and stresses affecting it. Application of this concept
helps to understand the present-day structure of the EC in terms of its Meso-Cenozoic tectonic
evolution, and allows one to propose some hypotheses about its deep structure. In the first part of the
chapter I summarise the available geophysical data used to constrain the deep structure of the EC. Then
I review rheological properties of the continental lithosphere and present rheological models of the EC
lithosphere along a regional cross-section through its Mesozoic extensional basin formation and
Cenozoic basin inversion history. Using the results of these rheological models and previous flexural
models I discuss the rheological evolution of the EC lithosphere during Meso-Cenozoic time.
Subsequently I discuss the probable plate-tectonic-related palaeo-stress field evolution that affected the
EC lithosphere and the interaction with the EC lithosphere throughout the Meso-Cenozoic tectonic
evolution of the EC. Finally, I compare the EC with similar mountain belts, and with analogue and
numerical model experiments from literature in order to propose some hypotheses about the deep
structure of the EC.

2. GEOPHYSICAL DATA CONSTRAINING THE DEEP STRUCTURE OF THE EASTERN


CORDILLERA

Data constraining the present-day deep structure of the EC are limited to seismic refraction, gravity and
earthquake data.

2.1. SEISMIC VELOCITY MODEL BASED ON REFRACTION IN THE SOUTHWEST OF


COLOMBIA

Ocola et al. (1975) presented a model of seismic velocity in the crust and upper mantle, for a
refraction line going from La Cocha Lake in the southern part of Colombia to Bogota (Fig. 6.1). Leeds
(1977) used this seismic refraction model and with additional surface wave measurements using a
resolution analysis (Wiggins, 1972) calculated a shear wave velocity model for the crust and mantle in
the Colombia-Ecuador region (Fig. 6.1). According to this model there is a zone of low seismic
velocity at a depth of 110-200 km. For simplicity Coral (1985) assumed that the low velocity zone
extends below all of the EC.

2.2. GRAVITY AND MOHO DISCONTINUITY DEPTH

I used gravity data from the map of Colombia, Eastern Panama and adjacent marine areas
published by Kellogg et al. (1991).
The depth to the Moho in Colombia (Fig. 6.2) has been calculated by Salvador (1991) using the
following method: (1) Application of free air, Bouguer and topographic corrections to the gravity data;
(2) Two-dimensional gravity modelling along three seismic refraction profiles, in the west and
southwest part of Colombia (Ramrez and Aldrich, 1973; Ocola et al., 1975; Money et al., 1979; Fleh
et al., 1981), in order to isolate the gravity effect of the Moho. He found the frequency effect of the
Moho fell in the 250-580 km wavelength band; (3) The gravity field due to the Moho was recovered
from the total field using a two-dimensional band pass filter with these corner frequencies. The
estimated depth to the Moho was consistent with the values inferred from refraction seismic. Also
Salvador (1991) applied flexural models considering only the topographic load of the EC and assuming
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

A. Velocity (km/sec)
3 4 5 6 7 8 9

50
Depth (km)

100

Vp

150

200
Vs

B. 5o
N

BOGOTA
A
E

4o BUENAVENTURA
C
O

3o
C

COLOMBIA
I

2o
F
I
C
A

PASTO
P

1o LA COCHA

0o
QUITO
P E R U

1o
E C U A D O R

RIO BAMBA
2o
GUAYAQUIL

81o 80o 79o 78o 77o 76o 75o 74o 73o

Figure 6.1: A. Seismic velocity model of the crust and upper mantle along a refraction line going from La Cocha
Lake in the southern part of Colombia to Bogota in the EC. Vp: compressional wave velocity (from Ocola et al,
1975); Vs shear wave velocity, the arrows represent the uncertainties for the two mantle layers (from Leeds, 1977).
B. Location of the refraction line (from Ocola et al, 1975).

218
Chapter 6
33 33
Serrania Maracaibo 41
34 Basin 40
34 de Perija

cuas
35
36
1400 37 s
38 de Barinas
An

nL
39 Cucuta
40 a
35 rid

e Sa
41 basin
42 Me 39 8 37 36
43 38 39 3 35 3 4

nia d
44

40
33
36 45

Serra
46
1300

41
37 Bucaramanga
38
45 Arauca
39
44

45
44

44
40 43

43
42
41
42 42
1200 43
Medellin 44

r a
ille
44

rd
Co
lera

45 41

rn
Tunja
C ordil

st e
46
1100 Ea
47
Yopal 40 Llanos Orientales
Manizales 48
Basin (LLA)
39

38

1000 Bogota 37
33
34

47

Ibague 36
35

35
36

46

34
37

Villavicencio
45
38

44
39

43

40 33 0 100 200 Km
900
tr a l

32
C en

41

41 42
42 42 MOHO
43 Neiva
DEPTH
30

28
29

44
27
37

800 45
46 San Jose del Guaviare (km) from gravity
47
31
36
S

35
34
46

41

48
44
45

40

39

38
43

31
32
42
47

33

800 900 1000 1100 1200 1300 1400 1500

Figure 6.2: Moho discontinuity depth under the study area estimated from gravity data (modified after Salvador,
1991). Major strike-slip faults and suggested subcrustal mantle lithosphere subduction under the EC are also shown
(see discussion in section 6.3, 6.4 and 6.5).

a laterally constant effective elastic thickness. He obtained values of 25 km which is in excellent


agreement with the values for most of the EC presented here, which were calculated assuming lateral
changes in elastic thickness. Salvador (1991) interpreted the presence of a volcanic chain in the Central
Cordillera as the cause for a weak lithosphere, and he also found that the gravity effect of the deflection
is enough to compensate the topography.

219
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

2.3. SEISMICITY

Figure 6.3 shows seismicity distribution of the EC, obtained from the United States Geological
Survey, National Earthquake Information Center. Three profiles of seismicity across the Colombian
Andes are shown in figure 6.4 (from Pennington, 1981) and two profiles of seismicity across the EC are
shown in figure 6.5 (from Taboada et al., 1999; 2000).

Serrania
de Perija Maracaibo
Basin

CucutaB
8
Barinas
basin
Bucaramanga
Arauca

7 A

Medellin

6 Paipa
Tunja Iza

Yopal

Manizales A
5 B
Bogota
Llanos Orientales
Ibague
Basin (LLA)
Villavicencio
4

Guyana Shield
3 Neiva

0 100 200 Km
San Jose del Guaviare

-76 -75 -74 -73 -72 -71 -70


Figure 6.3: Seismicity in the study area. Triangles: shallow earthquakes (< 10 km). Lozenges: intermediate-depth
earthquakes (10 to 60 km). Circles: Deep earthquakes (>60 km). (modified from United States Geological Survey,
National Earthquake Information Center. World Data Center for Seismology,
http://wwwneic.cr.usgs.gov/neis/epic/epic.html). Note the local presence of volcanic rocks at Paipa and Iza and the
Bucaramanga earthquake nest south of Bucaramanga. AA and BB represent the location of seismicity cross-
sections shown in figure 6.5.

220
Chapter 6

NW SE

20

100
Ca
(P e r i b b e
r ez a
et a n Plat
l., 1 e
D

200
a 997
)
BUCARAMANGA SEGMENT
100 KM
300

NW SE NE SW

35 35

100
(T
ab N a
oa z c
da a P
D

e t la t 200
b al e c
. ,2
CAUCA SEGMENT 0 00 ECUADOR SEGMENT
)
300

Figure 6.4: Cross-sections of the seismicity of the three segments of subducted oceanic lithosphere, including the
seismicity of the overlying continental lithosphere. No vertical or horizontal exaggeration. Larger circles are
excellent and high quality locations. The positions of volcanoes projected onto the line are shown at the top of each
diagram, where they exist. The Bucaramanga nest is shown as a solid symbol in the top diagram. Location of the
sections is in Figure 6.8 (from Pennington, 1981 and after Prez et al., 1997 and Taboada et al., 2000). Note the
subduction angle (20) of the Caribbean Plate north of Colombia in (a) and the steeper angle of the Nazca Plate
(35) in (b).

On the base of well located seismic events (Fig. 6.4) and focal mechanism solutions (Fig. 6.6)
Penington (1981) concluded (Fig. 6.7) that: (1) The Panama block is accommodating east-west
compression along a series of thrust faults striking NW to NE; (2) The Andean Block (Ecuador,
Colombia and Venezuela Andes) is moving as a block NNE relative to South America, as indicated by
right-lateral strike-slip and thrusting focal mechanism solutions (Fig. 6.6), along a system of faults
following the eastern front of the EC; (3) The subducted portions of the old Farallon Plate under the
Colombian Andes have been segmented into three parts (Fig. 6.7): (a) A Bucaramanga segment
continuous with the Caribbean plate NW of Colombia; (b) A Cauca segment continuous with the Nazca
Plate being subducted west of Colombia-Ecuador; and (c) An Ecuador segment at the northern end of
the subducted lithospheric plate which is dipping at a small angle to the east beneath northern Peru.
However, according to Isacks and Molnar (1971) subduction is dipping to the east beneath Ecuador and
part of Colombia and to the south-east beneath northern Colombia. Pennington (1981) also recognised
that the Bucaramanga and Cauca segments can alternatively be interpreted as one continuous contorted
segment, of shallower dip to the north and with some aseismic regions within it. A particular feature of
the seismicity of the EC is a small zone (<65 km in diameter) of intense seismicity centred at a depth of
160 km, known as the Bucaramanga earthquake nest (Schneider et al., 1987).

221
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

Figure 6.5: Seismicity section across the EC. Location of the section in Fig.6.3. No vertical or horizontal
exaggeration. Vertical exaggeration for topography is 3, maximum topographic elevation is 4 km. (a) Section
orthogonal to the EC. It is approximately the same of the structural balanced cross-section of the EC by Colletta et
al. (1990; Figs. 5.4 and 5.7 Chapter 5, and Figs. 6.13 and 6.27). (b) Section parallel to the axis of the EC.
Earthquakes have been projected from a distance up to 150 km from the section (a) and up to 100 km from the
section (b). Note the seismicity below the EC suggests a subduction angle about 35-45 greater than the Caribbean
subduction angle below northern Colombia (see Fig. 6.4 and 6.8). (from Taboada et al., 2000).

222
Chapter 6

15o N

Caribbean Plate

FFP

10o

FB
Bloc andin

Cocos South American


Plate Plate
5o
U
FG
GR
ch

FD
tr en
ian

Nazca 0o
omb

A
FG

Plate
Co l

FDGR : Romeral Fault


FGA : Garzn-Altamira Fault
FGU : Guaicaramo Fault
FB : Bocono Fault 5 oS
FEP : El Pilar Fault

85o W 80 o 75 o 70o 65 o

Figure 6.6: Focal mechanism solutions along the eastern border of the Andes (Andean Block; after Pennington,
1981, Sarez et al., 1983, Cisternas et al., 1984 and Guillande, 1988).

Based on new data from the National Seismological Network of Colombia (NSNC), during the
period 1993 to 1996, Taboada et al. (1999, 2000) concluded that seismicity of the EC has the following
features: (1) Crustal earthquakes concentrated along the eastern and western foothills of the Cordilleras
are aligned with the main faults (Fig. 6.5). Shallow seismicity (< 25 km), occurs in the brittle domain of
the upper crust. Taboada et al. (1999) interpreted the scarce seismicity between 25 and 55 km depth as
coming from the lower crust (Fig. 6.5).; (2) Intermediate seismicity shows a NNE trending subduction
segment beneath the EC with a well-defined cluster along the NW margin of the Cordillera. (Fig. 6.5)
This subduction segment was not well defined before the installation of the NSNC. Taboada et al.
(2000) suggested that this slab corresponds to a remnant of the Palaeo-Caribbean Plate attached to the

223
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

Baud-Panam Block which displays a shallow subduction beneath the EC (Southern part of the
Bucaramanga segment of Pennington, 1981, Fig. 6.7).

15o N

Caribbean Plate

50
10o

0
Caribbean Plate

15
northern Bucaramanga
0
10 segment
1 00
BN a
+
Palaeo-Caribbean Plate
200
o
southern Bucaramanga 5
18 0
14 0

Cocos segment
0+

Plate
20
1 4010 0

Nazca Plate Nazca Plate


0

b
18

Cauca segment

0o
c
20 Ecuador segment
0+
18
1 0
10 40
0 4 oS
o 80 o 75o 70o 65o
85 W

Figure 6.7: Contours to the Benioff zones beneath north-western South America (labelled in kilometres). Lines a, b
and c show the location of seismicity cross-sections shown in figure 6.4. These three segments of subducted
lithosphere: Bucaramanga segment, Cauca segment and Ecuador were identified by Pennington (1981). A fourth
fragment of the Palaeo-Caribbean Plate attached to Panam under the EC has been recently demonstrated by
Taboada et al., (2000, see Fig. 6.5). Note that the strike (N 30 E) and subduction angle ( 35 to 45 SEE) of the
Palaeo-Caribbean Plate fragment beneath the EC (from Pennington, 1981 and Taboada et al., 2000) are different
from the strike (N 40-45 E) and subduction angle (20 SE) of the Caribbean Plate beneath northern Colombia (from
Pennington, 1981, and Prez et al., 1997). The different geometry of these slabs fragments and the different velocity
vectors suggest that they are converging obliquely along a SEE boundary where the Bucaramanga earthquake nest
(BN) is located (see also Fig. 6.9)

The intermediate seismicity slab beneath the EC is limited southwards, along a roughly E-W
trending boundary approximately located at 5.2 N (Fig. 6.7). This boundary aligns with the southern
termination of the Baud Range, and Taboada et al. (1999, 2000) suggested that this shear zone defined

224
Chapter 6
by the Itsmina and Ibagu Faults and the Santamarta-Bucaramanga Fault are the boundaries of a
moving continental wedge associated with the most active segment of this subduction (Figs. 6.7 and
6.8). The formation of this subducted slab probably is linked with the accretion of the Choc Block, the
indentation of the continental wedge between the Itamina-Ibagu Faults and the Santamarta-
Bucaramanga Fault and the uplift of the EC (Taboada et al., 2000). Accretion of the Choc Block at 12

20 Ma.
P a la e o -Cari b b e an P l ate

P a l a e o-Cocos
P l ate
o
a

P a l ae o -Nazc a P l ate

a a

Present 80 W

C a ri b b e a n P l a te

10

b
Cocos P l a t e
5
b

b Nazca P l a te
5 0 0 km

Figure 6.8: Tectonic plate reconstruction of the Northern Andes.


(a) Reconstruction at 20 Ma (modified from Taboada et al., 2000).
(b) Reconstruction at present time (modified from Taboada et al., 2000).
The collision of the Baud-Panam island arc (BPA oblique line pattern) favouring the uplift of the EC began at 12
Ma. Black circles indicate the location of active volcanism. Shaded areas indicate mountain ranges. At 20 Ma the
Central Cordillera (CC) and low topography mountains produced by incipient inversion of Mesozoic extensional
basins (TS Tablazo inverted sub-basin, CS Cocuy inverted sub-basin, SM Santander Massif, MA Mrida Andes,
GM Garzn-Quetame Massif were uplifted areas. EC Eastern Cordillera, WC western Cordillera, UMV Upper
Magdalena Valley, MMV Middle Magdalena Valley, LMV Lower Magdalena Valley, SSJ Sin-San Jacinto fault
and fold belts, AR abandoned ridge.

225
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

Figure 6.8: Continued


(c) Schematic tectonic cross-sections of the Northern Andes at 20 (section aa) and at present time (section bb).
NP Nazca Plate, PCP Palaeo-Caribbean Plate, RFS Romeral Fault System, CML, continental lithospheric mantle,
WC Western Cordillera, CC Central Cordillera, MV Magdalena Valley, EC Eastern Cordillera (from Taboada et al.,
2000).
80 W

10 N
Cari bbean Plate
Panama 70 W
SJ J SN SM
5 N
Coi ba SP
LM V
Micro-plate MA
Nazc a
Pl ate MM V
CC EC
WC LLA
South Amer i can
Pl ate

BN
k
1

Figure 6.9: Schematic block diagram indicating the 3D geometry and kinematics of the lithospheric plates
beneath the EC and north-western Colombia. Intracontinental subductions are defined beneath the EC and the
Mrida Andes. The Bucaramanga nest (BN) is located at the oblique collision between the Palaeo-Caribbean
subduction segment beneath the EC connected to the Panam Block and the Caribbean Plate subducting beneath
northern Colombia. Very deep overturning of the Nazca Plate has bee interpreted by Taboada et al. (2000) from
mantle tomographic sections. the inflected zone where the EC subduction joints the Santamarta-Bucaramanga Fault
(from Taboada et al., 1999). EC Eastern Cordillera, LLA Llanos Basin, MMV Middle Magdalena Valley, LMV
Lower Magdalena Valley, CC Central Cordillera, WC Western Cordillera, MA Mrida Andes, SP Sierra de Perij,
SNSM Sierra Nevada de Santa Marta (modified after Pennington, 1981; Prez et al., 1997; Taboada et al., 1999,
2000).

226
Chapter6
Ma blocked normal oceanic subduction of the Caribbean Plate beneath NW South America (Fig. 6.8;
Duque-Caro, 1990; Taboada et al., 2000). Taboada et al., (2000) interpreted that the northern part of
the Nazca Plate (Cauca segment according to Pennington, 1981; Fig. 6.7) subducts beneath the Choc
Block overlapping the southern part of the Palaeo-Caribbean Plate fragment recognized beneath the EC
(southern part of Bucaramanga segment of Pennington, 1981; Figs. 6.7 to 6.9). The convergence rate
along the oceanic trench decreased and active deformation shifted toward weak zones of continental
lithosphere. Shortening localised along Mesozoic extensional basins, creating tectonic inversion (Fig.
6.9). New data from the NSCN suggests that the Bucaramanga earthquake nest is not an isolated
phenomenon as previously though. Taboada et al., (2000) suggested that it is related to an inflection of
the subducting surface. However the subduction angle (~35 45) of the Palaeo-Caribbean plate-
fragment below the EC is stepper than the Caribbean plate subduction beneath northern Colombia
(about 20 according to Pennington, 1981 and Prez et al., 1997). (see Figs. 6.4, 6.5, and 6.7).

2.4. GEOTHERMAL REGIME

A compilation of geothermal gradient and heat flow estimates from literature is presented in
figures 6.10 and 6.11. Puerto (1982), Bachu et al. (1995) and ICP-GEX-DYA (1995) studied the
geothermal regime of the Magdalena Valley, Llanos Orientales and Neiva sub-basin (Upper Magdalena
Valley) respectively. According to these authors these data should be considered preliminary because
bottom hole temperatures are incomplete and contain errors, and thermal conductivity has not been
measured. The few estimates for the EC should also be considered preliminary. These estimates have
been calculated from organic matter maturity models using bottom hole temperatures and vitrinite
reflectance data by Mora (1996, 2000) and Lance-Le Cornec (1998). Local lateral changes in
geothermal gradients and heat flow estimates may result from convective heat transport due to water
circulation within the sedimentary rocks (Bachu et al., 1995; ICP-GEX-DYA, 1995) and evidenced by
numerous hot springs in the EC (Forero, 1958). Hydrothermal activity (300C) has also been
interpreted during emerald formation in the emerald mines (western and eastern flanks of the EC) by
Cheilletz et al., 1993). Such hydrothermal activity related to emerald crystallization occurred during
Palaeogene (38-32 Ma) according to Cheilletz et al. (1993) based on 40Ar/39Ar and K/Ar data, or close
to the Cretaceous-Paleocene boundary (67-61 Ma) according to Romero et al. (2000) based on 87Sr/86Sr
data.

3. RHEOLOGICAL EVOLUTION OF THE LITHOSPHERE OF THE EASTERN


CORDILLERA

3.1. RHEOLOGY OF THE CONTINENTAL LITHOSPHERE

The assessment of lithospheric rheology is an important tool in understanding lithospheric


dynamics (Ranalli, 1995). In this section I review the rheology of the continental lithosphere and I
present rheological models of the lithosphere of the EC along a regional cross-section through
Mesozoic and Cenozoic time.

3.1.1. Rheology of the continental lithosphere from rock mechanics data

The following summary about the rheology of the lithosphere has been extracted from
Beekman (1994) and Van Wees (1994). The rheological behaviour of rocks that constitute the

227
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

continental lithosphere is derived from laboratory experiments and subsequently extrapolated to


geologically relevant times.
At low confining pressures and low temperatures, such as those of the upper crust, brittle
fracture is predominant (strain-rate independent deformation). Fracturing is described by a Mohr-
Coulomb criterion, which, following Byerlee (1978), can be expressed in terms of effective principal
stress difference, lithostatic overburden pressure, and pore fluid pressure (Ranalli, 1995):

Serrania
Maracaibo
de Perija
GEOTHERMAL Basin
1400
GRADIENT Cucuta
Barinas
(F/100 feet) basin
After Puerto (1982)
Bachu et al. (1995) 0.75 0 100 200 Km
1300 ICP-GEX-DYA (1995) Bucaramanga
Arauca
Lance-Le Cornec (1998)

1200 Medellin

2.90
1.25 2.00
Tunja 3.51
3.34 3.12
1100 Yopal
3.51

Manizales

2.52

1000 Bogota
3.62
Ibague

1.75 2.25
3.95
1.25 1.25 Villavicencio
3.45 3.40
1.50 3.29 3.73
2.76
3.01
900

2.50 2.75

Neiva 3.00
800
San Jose del Guaviare Guyana Shield
800 900 1000 1100 1200 1300 1400 150

Figure 6.10: Compilation of present day geothermal gradient estimates. Estimates for the Llanos Basin from Bachu
et al. (1995), for the Middle and Upper Magdalena Valley from Puerto (1982), some estimates for the Upper
Magdalena Valley from ICP-GEX-DYA (1995) and for the EC from well data and organic matter maturity models
from Mora (1996, 2000) and Lance-Le Cornec (1998).

228
Chapter 6

Serrania 35 .0
Maracaibo
THE R MAL FLO W de Perija
Basin
1 40 0 (mW /m 2 ) s
de Barinas
Afte r B a c h u e t al. (1 9 9 5 )
C ucu ta
An
a bas in
ICP-GEX-D YA (1 9 9 5 ) rid
L a n c e - Le C o r n e c ( 1 9 9 8) Me
Mora (1996, 2 0 0 0 ) 0 100 200 Km
1 30 0 Buca ram anga
50 .0 Arau ca
iller

1 20 0 M edellin
C o r da

47 .0
Tunj a 50 .0
60 .0
1 10 0 39 .0
Yop al
47 .0
35 .4
29 .2 26 .8
M anizales 30 .0 37 .2
51 .9
35 .4 38 .1
28 .0
38 .6
Bog ot a 38 .8
1 00 0 33 .0 48 .5
Ib ag ue
33 .6
33 .0 41 .4
58 .5
Villavicencio 62 .7
42 .0
34 .0 55 .0
36 .0
al

34 .0
68 .3
9 00 19 .0
ntr

54 .4
Ce

57 .6 57 .9
53 .9
65 .3
26 2. 1 54 .7
N eiva
8 00
31 3. 7
San Jose del Gua viar e Guyana Shield
62 .3 35 .0

8 00 9 00 1 00 0 1 10 0 1 20 0 1 30 0 1 40 0 1 50 0

Figure 6.11: Compilation of present day geothermal heat flow estimates. Estimates for the Llanos Basin from
Bachu et al. (1995), for the Middle and Upper Magdalena Valley from Puerto (1982), some estimates for the Upper
Magdalena Valley from ICP-GEX-DYA (1995) and for the EC from well data and organic matter maturity models
from Mora (1996, 2000) and Lance-Le Cornec (1998).

brittle = 1 3 = gz (1 ) (6.1)
where
R 1 For normal faulting
R

= R 1 For thrust faulting
R 1
1 + (R 1) For strike-slip faulting

229
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

where 1 and 3 are maximum and minimal principal stresses, is density, g is gravitation acceleration,
z is depth, and where

P1
= = w 0.35 is the hydrostatic pore fluid factor
gz

0 < < 1 denoting magnitude of intermediate stress 2 = 3 + (1 - 3)

R= [ 1 + ] 2

is the (static) sliding friction coefficient. According to Ranalli (1995) a common value is = 0.75

Steady-state creep of a wide variety of rocks, as believed to occur in the lower crust, is described by a
ductile flow law. Experiments with rock-forming minerals show that the critical principal differential
stress necessary to maintain a given steady-state strain rate, is a function of a power of the strain rate
and varies strongly with mineral composition and temperature. The ductile flow
obeys the so called power-law creep (Kirby, 1983):

1
E
n
creep = 1 3 = exp p (6.2a)
Ap nRT

Where & is strain rate AP, n and EP are flow parameters, T is absolute temperature and R is the gas
constant. The above power-law creep indicates that critical differential principal stresses grow
exponentially with decreasing temperature. Flow parameters for various rock-forming minerals show
that power-law creep flow parameters are strongly controlled by silica content (e.g. Carter and Tsenn,
1987). Felsic rocks (e.g. granite) show low critical differential stress values compared to mafic rocks
(e.g. dunites), under similar conditions of strain rate and temperature. The flow parameters are also
very sensitive to the abundance of water. Wet rock samples show significantly lower critical
differential stress values than dry ones. However, according to the experiments, above a critical stress,
the power-law creep breaks down into the so-called low-temperature plasticity, which is characterised
by a nearly linear increase in stress with decreasing temperature. For olivine, the main constituent of
the subcrustal lithosphere, Goetze and Evans (1979) found that at differential stresses higher than 200
Mpa (=2 kbar) low-temperature plasticity is described by the so-called Dorn law:

1

RT 2
creep = 1 3 = D 1 ln
E D AD
(6.2b)

Where AD, ED, and D are flow parameters for low-temperature plasticity.

230
Chapter 6

3.1.2. Rheological profiles and integrated strength of a stratified lithosphere

For a given tectonic environment (thrusting, normal faulting or strike-slip faulting), at any
depth, flow properties, temperature and strain rate, the lowest of the brittle and creep principal stress
difference given by equations (6.1) and (6.2a/b) respectively, gives a rheological strength (called the
yield strength). For stress differences below this yield strength, the imposed strain rate will not occur.
The dependence of the principal stress differences in equations (6.1) and (6.2a/b) on external conditions
(pressure and temperature) and material properties (mineralogical composition), both varying with
depth within the lithosphere, makes the yield strength also vary with depth, thus constituting a strength
envelope.
Continental crust, with a thickness in the order of several tens of kilometres, is layered, with
large variations between different tectonic provinces. A simple and representative crustal mineralogical
model is a two layered crust with felsic rocks dominating the upper crust and mafic rocks dominating
the lower crust, as indicated by seismic velocities. These crustal layers overly ultramafic upper mantle
material composed primarily of olivine. For a typical continental lithosphere composed of a quartzite
upper crust, a diorite lower crust, and an olivine mantle, the associated rheological layering is shown in
Figure 6.12. Composite strength envelopes are shown for compressive (thrust faulting) and tensile
(normal faulting) stress regimes at several stages of thermal cooling. Lithosphere strength envelopes
typical for Phanerozoic extension and inversion settings show a marked layering of a relatively strong
(mostly brittle) upper crust, a weak (mostly ductile) lower crust, and a strong (brittle and ductile)
subcrustal lithosphere. This layering, predicted from extrapolation of rock mechanics, agrees quite well
with interpretation of geophysical and geological data. This rheological layering is confirmed by the
depth distribution of seismic activity, which is approximately confined to the crustal layers (e.g.
Cloetingh and Banda, 1992), as well as by the presence of minima in seismic wave velocity and
electrical resistivity coinciding with the ductile layers (Ranalli and Murphy, 1987). The rheological
stratification may significantly affect the thermo-mechanical response of the continental lithosphere to
stresses. For example, large normal faults in the upper crust tend to flatten out towards a detachment
zone in the lower weak part of the upper crust (Braun and Beaumont, 1987). Another possible effect of
the strong/weak intracrustal alternation may be mutual mechanical decoupling of the layers under
substantial loading along intra-lithospheric shear zones leading to breakdown in coherent deformation
of the multi-layered continental lithosphere (Ord and Hobbs, 1989).
A scalar measure for the total strength of a (multi-layered) lithosphere with a depth varying
rheology can be obtained by vertically integrating the yield envelope:

h
L = y (z )dz (6.3)
0

Where L represents the total lithospheric strength, h is the thickness of the mechanically strong
lithosphere, defined, for example, by the depth where the rheological yield stress has decreased to a
level of 10 Mpa and with no further rheological discontinuities at larger depths.

231
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

Strength (MPa) / Temperature


-1000

1000
-500

500
Strength evolution of rifted lithosphere
0
0
ext Sediments
comp
Crust
20

40 Mantle
Lithosphere
60
Km 100 Ma 30 Ma
90 Ma 60 Ma 0 Ma
Stretching Coolng

Age (Ma)
100 80 60 40 20 0
Integrated strength (TN/m)

25
Compression
20

15

10
Extension
5

Figure 6.12: Depth-dependent rheological models for the evolution of a stretched lithosphere (from Van Wees,
2000).

3.1.3. Thermal structure of continental lithosphere

Temperature plays an important role in the thermo-mechanical response of the lithosphere to


stresses. Of particular importance is the variation of temperature with depth on the ductile strength of
lithospheric rocks. Conductive cooling of the lithosphere over a relatively short period of a few million
years after the last heating event rapidly leads to a significant increase in the mechanical strength of the
lithosphere. Over longer periods cooling also leads to thickening of the lithosphere, and of its
mechanically strong sub-layers. For both oceanic and continental lithosphere the derived geotherms are
solutions of the Fouriers law of heat conduction

T
c p = .(kT ) + A (6.4)
t
where T is absolute temperature, t is time, and with the material parameters: density, Cp heat capacity
at constant pressure, k the coefficient of heat conduction, and A the radioactive heat production per
volume unit. Material parameters are assumed to stay constant within the major crustal and subcrustal
layers, except for the heat production A, for which in some cases an exponential decay with depth is
adopted. Initial geotherm (temperature-depth curve) were assumed to be equal to those of a steady-state
geotherm (Chapman, 1986).

232
Chapter 6

There are several processes that may affect the temperature distribution locally or regionally,
including erosion and sedimentation (thermal blanketing), among others. In general, it is the most
recent thermal event that controls the thermal structure of the continental lithosphere.

3.2. RHEOLOGY MODELS OF THE EASTERN CORDILLERA THROUGH TIME

3.2.1. Rheology models

Sarmiento et al. (in prep) forward modelled the evolution of rheological 1-D profiles during the
Mesozoic extension basin formation and Cenozoic basin inversion along a cross section of the EC (Fig.
6.13, see also Fig. 3.1.), using the stratigraphic record to calculate tectonic subsidence curves and
lithospheric crustal and subcrustal stretching events (see Chapter 2). The forward modelling approach
is based on lithospheric stretching assumptions (McKenzie, 1978; Royden and Keen, 1980). For the
Cenozoic basin inversion episodes we used the present-day topography and the Palaeogene palaeo-
topography estimated from flexural models (see Chapter 3, Table 3.2a/b), as well as the Neogene
erosion as observed (see Chapter 4), in order to estimate compressional lithospheric factors
(stretching crustal and subcrustal lithospheric factors < 1, < 1). Compressional lithospheric factors
also were independently calculated from: (1) Comparison of present-day crustal thickness from gravity
data (Salvador 1991,Fig. 6.2) and values estimated from basin lithosphere stretching models at the end
of Cretaceous (see Chapter 2, Fig. 2.35); (2) Crustal shortening amounts due to Neogene compressional
deformation calculated from regional structural balanced cross-sections of the EC (see Chapter 5).
For the continental lithosphere a model with a layer of sediments, an upper crust and a lower
crust overlying a half-space mantle was adopted. The surface and base of the lithosphere are treated as
fixed temperature boundaries during the basin extension/inversion history. Initial temperatures were
derived from a steady-state geotherm, assuming a surface temperature of 0C (although a real surface
temperature value is ~ 25C, the error of this assumption is <1% and affects only the ductile
temperature dependent lithosphere layers), and a melting point temperature of the lithosphere of 1333
C (these are standard values used in literature). Sarmiento et al. (in prep) applied a simple one-
dimensional kinematic forward model, which calculates the thermal evolution of the lithosphere using
the crustal and subcrustal stretching lithospheric factors constrained by tectonic subsidence data (see
Chapter 2). This model can be used to calculate the evolving geotherm resulting from thermal
equilibration after stretching events and to calculate rheological strength evolutions (Van Wees and
Beekman, 2000). For the inversion episodes we applied lithospheric compression by means of
stretching events values less than 1 as already described. Material parameters used are listed in table
6.1. Sarmiento et al. (in prep) adopted the values of material properties from Van Wees (1994) and
calculated rheological profile evolution using three different material parameters (listed by Okaya et
al., 1996) for power-law creep rheologies taking into account a decrease in silica content with depth:
(1) Dry quartzite for the upper crust, wet quartz-diorite for the lower crust and dry dunite for the
mantle; (2) Wet quartzite for the upper crust, wet quartz diorite for the lower crust and wet dunite for
the mantle, and (3) Dry granite for the upper crust, dry diabase for the lower crust and dry olivine for
the mantle. However the results presented are only for the first rheology. Dry olivine rheology is
combined with low-temperature plasticity (eq. 6.2b). We used a strain rate of 10-15 s-1, corresponding to
the deformation rates induced from the forward basin modelling. For the relatively fast Neogene
Andean deformation we used a strain rate of 10-14 s-1. For rheological strength calculations in the brittle
domain we assumed hydrostatic pore fluid pressure ( 0.4) and = 0.75 (Ranalli, 1995). For the
initial crustal thickness we adopted a value of 35 km (which is the value of the stable cratonic South
American Moho depth in the eastern LLA area according to Salvador, 1991) and an initial lithospheric
thickness value of 120 km. For the numerical integration of the heat conduction equation in one
dimension a finite difference technique was used. The grid spacing of the 1-D vertical mesh is about
250 m. The model allows the incorporation of finite and multiple stretching/compression phases and
includes thermal effects of heat production and sediment fill.

233
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

10 Shortening-Thickening Maracaibo
ratio Serrania Basin 12
de Perija 11 s
1400
Basin
Length initial/lenght de Barinas
final Cucuta An
compartment a
1.38 Paleogene rid basin
1.58 Neogene 7
1.00 Me
1.37
Villa de Leiva Rheology model 2 1.35 0 100 200 Km
1.00
Bucaramanga
1300
Tablazo Arauca
A Casabe-199
Infantas-1613
Cascajales-1

1 Cocuy
1.01
1200 Medellin 9
Cimitarra 1.04 1.72
3 1.36
i lle ra

Velez
Arcabuco Medios
C o rd

Tibasosa 1.00
1 Yopal 10
1100 Tunja Pajarito 1.00
Inverted Cocuy Aguazul
1.03 sub-basin
Manizales 4 1.44
Suesca-1
La Maria

Quipile Planas-1

1000 Apulo
Ibague
1.03
5 1.59 Caqueza
INVERTED BASIN COMPARTMENTS
Q El Cobre 8 1.01 1. CENTRAL CORDILLERA
1.38 1.37 2. MIDDLE MAGDALENA VALLEY
6 1.58 Villavicencio 3. TABLAZO SUB-BASIN
Q Olini 4. CUNDINAMARCA SUB-BASIN
al

900 Prado
5. SOUTHERN CUNDINAMARCA SUB-BASIN
nt r

6. UPPER MAGDALENA VALLEY


Ce

7. SANTANDER FLORESTA MASSIF


8. S CUNDINAMARCA AND W COCUY SUB-BASINS
Neiva 9. COCUY SUB-BASIN
La Macarena

10. LLANOS ORIENTALES


Serrania de

800
11. MARACAIBO (CATATUMBO) SUB-BASIN
12. MERIDA ANDES Guyana Shield

800 900 1000 1100 1200 1300 1400 1500

Figure 6.13: Location map of the regional cross-section of the EC and the lithosphere rheological models shown in
figures 6.14, 6.15 and 6.18. Note the shortening-thickening ratios calculated for each inverted basin compartment
from rheological models for the Palaeogene, and from structural balanced cross-sections for the Neogene.

Thermal
Initial Specific Heat Rheology 1
Density Conductivity expansion
Layer thickness heat production (rock
(Kg m-3) (W m-1 oC-1) coefficient
(km) (J kg-1 oC-1) (W m-3) analogue)
Quartzite
Sediments 0 2500 2.6 900 1.66 3.4x10-5
(dry)
Quartzite
Upper crust 17.5 2800 2.6 1190 1.66 3.4x10-5
(dry)
Lower Quartzdiorite
17.5 2800 2.6 1190 0.5 3.4x10-5
crust (wet)
Mantle 3300 3.1 1190 0.0 3.4x10-5 Olivine (dry)

Table 6.1. Parameters used to calculate the rheological models of the lithosphere of the EC.

234
Chapter 6

3.2.2. Local isostasy vs regional isostasy effects

In order to study local isostasy vs regional isostasy effects Sarmiento et al. (in prep) performed
the following steps:
(1) We modelled in 1-D the lithosphere stretching/shortening evolution at each location based on the
observed sedimentary record (see Chapters 2 to 4) and topography history inferred from flexural
models (see Chapters 3 and 4).
(2) We compared the stretching/shortening calculated with the models with independent
stretching/shortening estimations based on crustal thickness from gravity (Salvador, 1991; Fig.
6.2) and structural balanced cross-section interpretations (see Chapter 5). The 1-D model results
suggest excessive thickening of the crust in disagreement with gravity interpretations. The
difference between the 1-D model results (which imply local isostatic compensation) and
independent observations can be explained as the result of the flexural behavior of the lithosphere
(see Chapters 3 and 4), which imply regional isostatic compensation.
(3) Although we did not apply a true flexural correction, we used the observed Moho depth (from
Salvador, 1991) and the observed present-day topography to correct the models, the new
stretching/shortening results were more in agreement with independent stretching/shortening
estimations. The rheological models presented below include this correction.

3.2.3. Model results

Figures 6.14, 6.15 and 6.16 show the modelled depth-dependent strength of the lithosphere, the
integrated lithosphere strength and their evolution throughout the Mesozoic basin extension and
Cenozoic basin inversion history of the EC. Figure 6.17 shows the modelled heat flow at the top of the
basement resulting from such tectonic history. This evolution was the result of thermal re-equilibration
after basin extension and inversion. For these calculations we neglected the effect of gravitational
buoyancy forces arising from lateral variations in crustal thickness and temperature (Buck, 1991),
because they contribute to no more than 0.5 T/m (0.5x1012 N/m). The integrated strength can be taken
as a measure to overcome observed lithospheric deformation. In this case decrease or increase of
strength values relative to initial values (representing the unaffected margins: La Maria well in the
unaffected LLA area in figures 6.14 and 6.15) are indicative for localization, locking/widening or
deformation activity. This approach is valid if deformation is homogeneously distributed in the
lithosphere on the scale of the basin. Because at each location the model is 1-D we assumed local
strength estimates for lithospheric deformation for the entire extension and inversion history.

3.2.4. Comparison of lithosphere strength estimates from rheological models and flexural models

Results of rheological models of the EC are in partial agreement with lithosphere strength
estimates from flexural modelling. Both types of models indicate crust-mantle decoupling with a weak
lower crust (see Fig 4.15).
In agreement with flexural models calculated for Palaeogene time, rheological models at the
end of Cretaceous and Palaeogene time indicated a weak lithosphere in the area of the Mesozoic
extensional basin. The latter is characterised by integrated strength values lower than 12 TN/m for
extension and lower than 19 TN/m for compression, and effective elastic thickness values between 1
and 5 km. Additionally, both models indicate a stronger lithosphere in the stable LLA area. Such a
stronger stable lithosphere, not affected by Mesozoic stretching, is characterised by integrated strength
values of 12 TN/m for extension, 19 TN/m for compression and EET values between 50 and 60 km.
However, results of rheological models and flexural models for the present-day are in
disagreement. Rheological models suggest that the lithosphere of the EC is stronger (integrated strength
values for compression L > 22 TN/m) than that of the stable South American lithosphere in the
eastern LLA area (L 22 TN/m). Flexural models suggest an opposite trend with a weaker lithosphere

235
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

NW END JURASSIC (142 Ma) SE


Paleo-Magdalena Magdalena-Tablazo Santander-Floresta
Valley sub-basin paleo-high

Casc ajales-1

La M aria
Strength

Arcabuco

Pajarito
Tibasosa
Medios
scale (MPa)

Velez
-1000 -500 0 500 1000 1500

Tension Compression 0

20
Moho

Depth (km)
14 40
12
Integrated 10 60
strength 8
(TN/m) 6
4 80
a 2
0
100

END CRETACEOUS (65 Ma)


Paleo-Magdalena Magdalena-Tablazo Santander-Floresta Cocuy Llanos
Valley sub-basin paleo-high sub-basin Orientales
Cascajales-1

A rcabuco

Tibasosa

La Maria
P ajarito
M edios
Velez

Moho

14
12
Integrated 10
strength 8
(TN/m) 6
4
b 2
0

0 100 km

Approximate horizontal and vertical scale

Figure 6.14: Depth dependent rheological models of the EC calculated for Mesozoic time: (a) End of Jurassic (142
Ma); (b) End of Cretaceous (65 Ma).

in the EC (EET=25 km, assuming a continuous elastic plate, or locally EET = 0 km assuming a broken
plate under the EC) and a stronger lithosphere in the stable South American eastern LLA area (EET=50
km). Flexural models calculated for present time indicate a weak lithosphere under the EC; in contrast
rheological models suggest a strong lithosphere under the EC. Two important effects not considered in
the standard rheological models of the EC may explain the difference:

236
Chapter 6

NW EN D P ALEO G E NE (16.4 M a) SE
P a le o - M a gd a le n a Ma g d a l e na -Ta bl a zo T un ja A x ia l Co cu y L la n os
V a ll e y s u b -b a sin R e g i on su b- b a s in O ri e n ta le s
C a sc ajale s-1

Arca bu co

Tib as osa

P aja rit o
Str eng th sc ale (M Pa )

Me dio s
P a le o - Ce n tra l C or d il le r a

Ve le z
-100 0 -500 0 500 1000 150 0

0
C omp res sion
T ens ion

20

Mo h o

D epth (km )
45 40
40
35
30
Int eg rated 25
60

s trength 20
15
(T N /m ) 80
a 10
5
0
10 0

P RE SE N T D AY
M a g da le n a In ve rte d M a g d ale n a - Tu n ja A x ia l Inv er te d C o cu y L l an o s
V a ll ey Ta b la z o su b -b a si n R e gi o n s u b - ba sin O r ie n ta le s
C as cajales -1

Arc ab uco

Tib aso sa

Pla na s -1
L a Ma ria
Me dio s

P aja rito
V e lez

Moho

45
40
35
Int eg rated 30
s trength 25 .
20 ..
(T N /m ) 15
10
b 5
0
0 1 0 0 km

A p p rox im ate h or iz o ntal an d v ertic al s ca le

Figure 6.15: Depth dependent rheological models of the EC calculated for Cenozoic time: (a) End of Palaeogene
(end of early Miocene, 16.4 Ma); (b) Present time (0 Ma). The present time models illustrated here were calculated
assuming local isostasy. Theoretical Moho geometry inferred from these models may suggest mantle lithosphere
subduction. Compare with the present time models calculated with the observed Moho depth (from Salvador,
1991) interpreted to represent flexural isostasy.

237
(
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

i
C om p res s ion
45

40

35

30

25

20
h

15

10
L

Age (Ma)
0
a 30 0 25 0 20 0 15 0 10 0 50 0

Exte n s io n
40

35

30 C ascajale s-1 M iddl e Mag dalen a Val ley MV

Velez M agda lena -Tab lazo


Arca buco inver ted s ub-ba sin
25 Los M edio s EC
Tibas osa Foresta M assif
Pajari to C ocuy inver ted s ub-ba sin
20 LLA
La M ari a -1 Llan os B asin
h

15

10
L

0
30 0 25 0 20 0 15 0 10 0 50 0

b Age (M a )

Figure 6.16: Modelled lithosphere integrated strength evolution at several locations along the regional cross-section
of the EC shown in Fig. 6.13. (a) Lithosphere strength to compression; (b) Lithosphere strength to extension.

238
Chapter 6

(1) The presence of crustal discontinuities that weaken the lithosphere and play a crucial role in strain
localisation (Van Wees, 1994; Van Wees and Stephenson, 1995; Ziegler et al., 1995). Deep
reaching pre-existing crustal discontinuities, such as lithological inhomogeneities associated with
ancient suture zones and faults, cause significant weakening of the upper crust. Such
discontinuities are apparently characterised by a reduced friction angle particularly in the presence
of fluids (Sibson, 1995; Van Wees, 1994; Ziegler et al., 1995; Sibson, 1995). During the Neogene
the complete basin inversion of the EC probably linked weak upper crust faults with the weak
lower crust, reducing the integrated lithosphere strength. However, predictions from rheological
models of the EC suggest that present-day deformation would be restricted mainly close to the
borders of the EC, which is in agreement with shallow upper crustal seismicity, which is
concentrated along the eastern and western borders of the EC (Figs. 6.5 and 6.18). Upper crustal
seismicity is a general feature observed in many other areas (e.g. European lithosphere, Cloetingh
and Burov, 1996). Intraplate seismicity is essentially restricted to the upper crust providing
additional support to the notion of upper crust-lithospheric mantle decoupling, as suggested by
Cloetingh and Burov (1996) for the European lithosphere.

80

70

60

50

40
300 250 200 150 100 50 0
Age (Ma )
C ascajale s-1 M iddl e Mag dalen a Val ley MV

Velez M agda lena -Tab lazo


Arca buco
inver ted s ub-ba sin
Los M edio s EC
Tibas osa Foresta M assif
Pajari to C ocuy inver ted s ub-ba sin
La M ari a -1 Llan os B asin LLA

Figure 6.17: Modelled heat flow at the top of the basement at several locations along the regional cross-section of
the EC shown in Fig. 6.13. Input data were the tectonic subsidence curves shown in Fig. 3.16. The heat conduction
from the asthenosphere, heat generation from the crust and sediments blanketing effects, but it does not include heat
generation from the sedimentary rocks filling the basin or convective heat transport. Differences between these
modelled results and actual measurements may in part be due to these two effects, as well as additional effects
discussed in the text.

(2) Anomalous heat input to the lithosphere additional to the thermal effects included in the
rheological models which are: (a) Thermal conduction; (b) Heat radioactive generation in the
upper crust; and (c) Thermal blanketing effect of the sedimentary cover. The local presence of
Neogene volcanic rocks in the northern part of the Tunja axial region of the EC and southwest
Floresta Massif (Renzoni et al., 1967; Romero and Rincn, 1990, Figs. 6.3 and 6.4), elevated
geothermal gradients (Fig. 6.10) as well as several hot springs in the central and eastern parts of
the EC (Forero, 1958) suggest that such additional heat input exists. Modelled strength profiles
are inconsistent with the observed seismicity (Fig. 6.18). The strong mantle lithosphere predicted
by the rheological models is characterised by a practical absence of seismicity. Standard
rheological models of the EC for the present time are inconsistent with flexural modelling

239
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure
Magdalena Llanos
Valley Basin Eastern Cordillera
Paipa - Iza vo lcanic rocks Basin
NW SE

1 0 0 km

2 0 0 km
a 0 1 0 0 km 2 0 0 km

Aquitaine Basin Pyrenees Ebro Basin


LATE Y P R E S I A N
A rize NPF B i x ol s Montsec

Nogueres

Om Rialp

c
50 K m 0
Subducte d lower crust

Magdalena Llanos
Valley Basin Eastern Cordillera
Basin

Figure 6. 18: (a) Depth dependent rheological models of the EC at present time (0 Ma) along a regional section of
the EC showing the structure (after Colletta et al., 1990) and observed seismicity (after Taboada et al., 1999, 2000).
The schematic deep structure of the EC is the hypothesis proposed here. (b) Numerical geodynamic model of a
collisional orogen (from Beaumont et al., 2000). (c) Crustal cross-section of the Pyrenees (from Beaumont et al.,
2000). Compare the hypothetical structure of the EC (a) with the geodynamic numerical model of a collision orogen
(b) and with the structure of the Pyrenees. See section 6 for discussion.

240
Chapter 6
results, local volcanic rocks and hot springs in the EC and the observed intermediate seismicity. This
point will be discussed in section 6.6 of this chapter.

3.2.5. Models of evolution of lithosphere rheology and basin extension and inversion in the EC

At the end of the Jurassic the reduced integrated strength values for extension (L < 12 TN/m)
indicate a weak lithosphere under the Magdalena-Tablazo sub-basin. Clearly this weak lithosphere was
the result of heating and thinning of the lithosphere as a result of Triassic and Jurassic rifting in this
area. Models predict that further deformation and strain localization would occur in this region. In
contrast, the models suggest that the unaffected lithosphere east of the Magdalena-Tablazo sub-basin
(L 12 TN/m) would not be prone to further deformation. However, Early Cretaceous rifting did
occur in this eastern region, generating the Cocuy sub-basin (see Chapter 2). In the Magdalena-Tablazo
sub-basin further rifting occurred during Early Cretaceous time (see Chapter 2) in agreement with
model predictions, though in an opposite sense to the predicted rifting localization in the western
Magdalena-Tablazo sub-basin. Evidently, the lithospheric strength models are not completely in
agreement with observed deformation during Early Cretaceous time. Because the model includes
thermal blanketing effects due to sedimentation, this effect cannot be used to explain the difference
between model prediction and observation. Following Van Wees (1994) the difference may be due to
weak fault/suture zones with considerable lower strength values than those predicted from bulk
properties of the lithosphere. Development of the Cocuy sub-basin during the Early Cretaceous
probably was due to reactivation of a weak normal fault zone bounding a Palaeozoic rift (Hossack et
al., 1999). This weak zone also represents the Palaeozoic (Toussaint, 1993) or Precambrian (Etayo-
Serna et al., 1983) Guaicramo frontier along which the Chibcha Terrane was accreted (Etayo-Serna et
al, 1983; Toussaint, 1993).
During Early Cretaceous a wide rift system developed with two major rift sub-basins, he
Magdalena-Tablazo and the Cocuy sub-basins, with the less subsiding Santander-Floresta palaeo-high
between them (see Chapter 2). Models suggest that at the end of Cretaceous thinning and heating of the
lithosphere produced by rifting clearly resulted in reduced integrated strength values for extension (L
< 12 TN/m) in the wide rift system. The two sub-basins were characterized by strength values for
extension lower than 10 TN/m. The Santander-Floresta palaeo-high was characterized by strength
values for extension intermediate between those of unaffected lithosphere of the LLA and those of the
rifted sub-basins (10 TN/m <L < 12 TN/m). In accordance to model predictions Palaeogene
deformation mainly affected the weakest lithosphere of the two-rifted sub-basins that were slightly
inverted, generating some local palaeo-topography (see Chapter 3). Local sedimentary basins were
preserved in the MV, Tunja Axial region (former Santander-Floresta palaeo-high) and LLA (see
Chapter 3). Rheological models suggest that at the end of the Palaeogene (end of Early Miocene)
lithosphere strength in the former extensional basin had increased due to thermal re-equilibration
cooling and slight basin inversion. However integrated strength values for compression in that region
(L < 19 TN/m) still were lower that those of unaffected lithosphere (L 19 TN/m). In accordance
with these model predictions further compressional deformation affected weak lithosphere (L < 19
TN/m).
During the Neogene the original extensional basin was inverted (see Chapter 4) in accordance
with rheological model predictions. Deformation affected a wide belt of the EC. Rheological models
calculated at the present time suggest a strong lithosphere in the area of the EC characterized by
integrated strength values for compression (L > 22 TN/m and even L > 35 TN/m) greater than those
of unaffected LLA lithosphere (L 22 TN/m). Only the deepest proximal depocenters of the Neogene
MV and LLA foreland basins are characterized by integrated strength values lower that those of
unaffected lithosphere (L < 19 TN/m). This result probably is related to thermal blanketing effect due
to thick sediments in these proximal foreland depocenters. However flexural models suggest a weak
lithosphere in the EC (see Chapter 4). As mentioned before probably the presence of crustal
discontinuities and anomalous heat input explain the difference. Shallow seismicity suggests that at the
present-day, brittle deformation is mainly active at its frontal eastern and western borders. Probably
most of the present-day brittle deformation is concentrated by transpressional reactivation of the

241
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

external fault systems that bounded the Mesozoic extensional basin and their short-cuts (see Chapters 4
and 5). If these rheological models were reliable they would allow us to predict that further deformation
of the EC is locked. Therefore further deformation would rather be restricted to the proximal
depocenters of the MV and LLA foreland basins close to the borders of the EC. These predictions seem
to be in agreement with present-day shallow upper crustal seismicity, which mainly is concentrated
along the eastern and western fronts of the EC (Figs. 6.3 and 6.5). During Palaeogene (see Chapter 3)
and Neogene (see Chapter 4) time, deformation affected a wide belt including most of the EC.
Localisation of deformation in the EC can be attributed in part to the occurrence of rheologically
permanent weak zones and additional heat input not considered in the standard rheological models.

4. EVOLUTION OF STRESSES AFFECTING THE LITHOSPHERE OF THE EASTERN


CORDILLERA

In addition to lithosphere rheology the second key parameter controlling lithosphere dynamics
is stress. I review here the possible plate-tectonic-related stress evolution that affected the lithosphere
of the study area.

4.1. MESOZOIC

The inferred Mesozoic stretching events (see Chapter 2) probably were initially produced by
tensional/transtensional stresses associated with the separation of South America and North America
during early Mesozoic time, and later by backarc extension (Figs. 2.3 and 2.4) as suggested by their
temporal correlation with gaps in magmatic activity or reduced activity in the Central Cordillera (Fig.
2.25). If the plutonic belts of the Central Cordillera were developed as subduction -related magmatic
arcs during Mesozoic times, as suggested by Aspden et al., (1987), the extensional basins behind them
may be interpreted as backarc basins. Where the rate of subduction exceeds the rate of convergence,
extensional deformation occurs in the overriding plate. Gravitational forces acting on the subducting
slab give rise to seaward retreat of the subduction boundary (Royden, 1993a, b). Trench suction forces
trigger extension in the overriding plate which subsequently fills the space created by the retreating slab
(Shemenda, 1993). Extension in the upper plate may also be produced by horizontal mantle flow acting
on top of the subducted plate, or to pulling out of the subducted plate from under the overriding plate
(Shemenda, 1993). Backarc extension is primarily driven by a combination of plate boundary forces
that cause subduction and body forces that act on the sinking slab. Important strike-slip components
have also been suggested by Aspden et al. (1987). According to Aspden et al. (1987) the Triassic
magmatic belt was controlled by strike-slip faults. Aspden et al., (1987) suggested that oblique
convergence and an offset in the subduction zone along a major NE-SW transform fault could account
for the notable absence of Cretaceous plutonism in southern Colombia and Ecuador.

4.2. CENOZOIC

Build up of intraplate compressional/transpressional stresses in the NW margin of South


America during latest Cretaceous-Palaeogene (see Chapter 3) could have been favoured by a
subduction impediment caused by the arrival of more buoyant oceanic crust, such as an oceanic plateau
(c.f. Nivia, 1987; Kerr et al., 1997; Sinton et al., 1998). In Colombia all plate-tectonic interpretations
(e.g. Pindell and Barret, 1990; Pindell, 1993; Pindell and Erikson, 1993; Pindell and Tabut, 1995)
propose collision of the Caribbean with northwestern South America. Collision was oblique and
diachronous, becoming younger northward (Pindell and Erikson, 1993; Pindell and Tabut, 1995).
Therefore plate-tectonic history suggests that Palaeogene basin inversion was collision-related,
probably associated with the right-lateral transpressional deformation that led to pre-Andean orogeny in
the Central Cordillera. Transpressionally deformed grabens are also associated with zones of major

242
Chapter 6
wrench faulting (Ziegler et al., 1995). In Colombia some right-lateral strike-slip faults (e.g. Palestina
Fault, Irving, 1971) could have been active during the Palaeogene.
Compression in the region behind a magmatic arc is associated with Andean type orogens and
occurs during periods of increased convergence rates between the subducting and overriding plates
(Ziegler et al., 1998). Inversion of rift tensional hanging-wall basins located behind a magmatic arc is
the result of acceleration of convergence rates between the colliding plates, their increased mechanical
coupling and the transmission of compressional stresses into the backarc domain of the overriding plate
(Uyeda and McCabe, 1983; Ziegler, 1993 in Ziegler et al., 1998). Rates of the Caribbean-South
America plate convergence changed during the Cenozoic. Periods of development of compressional
structures seem to correlate with times of high convergence rate, particularly during the Eocene (Daly,
1989) and during Late Miocene-Pliocene Andean orogeny (Cooper et al., 1995; see Chapter 4). During
Late Oligocene-Early Miocene development of compressional/transpressional structures was probably
associated to the rupture of the Farallon Plate into the Cocos and Nazca plates about 25 Ma ago (Wortel
and Cloetingh, 1981; Duncan and Hardgraves, 1984). In addition Neogene basin inversion and Andean
orogeny has been correlated to the collision of the Choc Block (Cuna Terrane according to
terminology of Toussaint, 1993, 1995a, b) with the northwestern margin of South America (Duque-
Caro, 1990, Cooper et al., 1995; Taboada et al., 1999, 2000; Fig. 6.8).

5. MESOZOIC-CENOZOIC TECTONIC EVOLUTION OF THE EASTERN CORDILLERA

5.1. MESOZOIC EXTENSIONAL BASIN FORMATION

During the Triassic and Jurassic (see Chapter 2) tensional/transtensional stresses probably
initially related to the break-up of Pangea, and later to backarc extension, resulted in lithosphere
stretching that generated narrow rifts (< 150 km wide), located in the places of the present-day MV and
the western flank of the EC (Magdalena-Tablazo sub-basin). Repeated stretching events in the same
area suggest that strain localisation affected a weak lithosphere. Thermal heating associated with
stretching and reactivation of crustal discontinuities could have contributed to such a strain localisation.
During the Early Cretaceous (see Chapter 2) tensional/transtensional stresses probably related to
backarc extension produced new episodes of lithosphere stretching, and generated a wide (> 180 km
wide) system of asymmetric half-rift basins. These stretching events generated an asymmetric half-rift
basin with a depocenter located along the palaeo-eastern flank of the EC and a major normal fault
system in its eastern border (Cocuy sub-basin). A system of horst blocks was also located in the area of
the Santander and Floresta massifs. A less developed second order half-rift developed at the
southwestern flank of the Cordillera (Magdalena-Tablazo sub-basin). Stretching in the Magdalena-
Tablazo sub-basin can be related to previously weakened lithosphere, but development of the new
Cocuy cub-basin only can be explained by reactivation of older crustal discontinuities, probably the
pre-Mesozoic Guaicramo palaeo-fault. Two layered stretching models with stretching factors of up to
1.66 for the crust and up to 3.49 for the subcrustal lithosphere suggest that some decoupling occurred
between the crust and subcrustal lithosphere, or that an increased thermal thinning affected the mantle
lithosphere. In places of maximum crustal stretching (greater than 1.4), small mafic intrusions were
emplaced during the Cretaceous (Fabre and Delaloye, 1983; Moreno and Concha, 1993). During the
Late Cretaceous, subsidence was produced by thermal relaxation of the lithosphere.

5.2. CENOZOIC TRANSPRESSIONAL INVERSION OF MESOZOIC EXTENSIONAL


BASINS

During the Palaeogene (see Chapter 3) compressional/transpresional stresses, probably


initially related to collision of the South American plate margin with a relatively buoyant oceanic
plateau and later related to acceleration of rates of plate convergence and collision to the Choc Block
during Neogene time, affected a weakened lithosphere below the former extensional basins. This
resulted in inversion episodes. Incipient basin inversion affected the weak lithosphere of the former
Magdalena-Tablazo and Cocuy sub-basins. Later, during the Neogene (see Chapter 4), the EC resulted
from the inversion of the Mesozoic extensional basins (e.g. Colletta et al., 1990; Cooper et al., 1995;

243
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

Casero et al., 1995, 1997) favoured by a previously weakened lithosphere. Basin inversion explains the
coincidence in location of the former extensional basins and the EC. Results of tectonic subsidence
analysis (see Chapter 2) and flexural modelling (see Chapters 3 and 4) agree with this interpretation.
These results suggest that contractional reactivation of Mesozoic extensional faults was initiated during
Palaeogene times, leading to an initial slight basin inversion; then during Neogene time the
compressional deformation increased in rate and magnitude leading to a complete inversion of the
original extensional basin. Scarce fission-track data and the results of flexural models suggest that
moderate (< 1000 m but mostly 200 m) and local surface-uplift and exhumation of blocks, mostly
associated to the master normal fault systems delimiting the Mesozoic extensional basins, was initiated
during Palaeogene times (see Chapter 3). However, only during Neogene times, increased surface-
uplift (> 1000 m) affected the former Lower Cretaceous depocenters adjacent to the master normal fault
systems delimiting the basin (see Chapter 4). Complete inversion of the master normal fault systems
delimiting the former extensional basins probably lead to extrusion of the sedimentary fill of half-
graben basins that now form the eastern and western flanks of the EC. Regional structural balanced
cross-section interpretations (e.g. Colletta et al., 1990, Fig. 6.5; Cooper et al., 1995, Fig. 3.37) suggest
that maximum dip-slip displacement and shortening occurred on the thrust faults connected with
inverted Mesozoic normal faults that now approximately delimit the Lower Cretaceous outcrops of
both flanks of the EC. In these uplifted flanks, Lower Cretaceous or older exposed rocks indicate that
all Late Cretaceous and younger sediments have been eroded. However only in the Santander, Quetame
and Garzn massifs all the sedimentary cover has been removed by erosion. Proximity of these massifs
to important strike-slip faults suggests that inversion was associated with important strike-slip motions.
Inversion of normal faults is only possible if there is a strike-slip component (Mandl, 1988).
Preservation of the Palaeogene sedimentary record in the axial Bogot-Tunja zone of the EC, as well as
of the Neogene sedimentary record of the Sabana de Bogot area suggest that this axial region
remained low during all Miocene time and was surface-uplifted only during Pliocene-Pleistocene time.
Flexural model results also support this interpretation. The exceptional palynological record of the
Neogene sediments of the Sabana de Bogot area documents this uplift history which occurred between
5 and 3 Ma b.p. (Helmens, 1990; Andriessen et al., 1993).

6. COMPARISON OF THE EASTERN CORDILLERA WITH SIMILAR MOUNTAIN BELTS,


ANALOGUE AND NUMERICAL MODELLING EXPERIMENTS AND HYPOTHESES
ABOUT THE DEEP STRUCTURE OF THE EASTERN CORDILLERA

In this section I highlight some structural features of the EC that are common to other mountain
belts, and I make comparisons with similar mountain belts, analogue and numerical models from
literature in order to propose some hypotheses about the deep structure of the EC.

6.1. ASYMMETRY OF THE EASTERN CORDILLERA

Opposite vergence of the EC is not symmetric (Colletta et al., 1990). Regional balanced cross-
sections perpendicular to the Cordillera at the latitude of Tunja (e.g. Colletta et al., 1990 Figs 5.4 and
6.18; Cooper et al., 1995, Fig. 3.37) show:
(1) A wide zone of east-verging thrusts located in the Tunja region and the eastern flank of the
Cordillera. A large amount of eastward-directed shortening occurred in this zone.
(2) A narrow zone of west-verging faults at the western front of the mountain range. A moderated
amount of westward-directed shortening occurred in this zone.
(3) A central zone of shallow pop-up basement (Vlez-Arcabuco anticlinorium) bordered by
opposite vergence reverse faults located west of the axial Tunja region (Colletta et al., 1990).

244
Chapter 6
6.2. COMPARISON AND SIMILARITIES WITH THE PYRENEES AND THE MRIDA
ANDES OF VENEZUELA

It is interesting to note some remarkable similarities between the EC (Figs 3.1, 6.14 and 6.15)
and the Central Pyrenees (Fig. 6.18 and 6.19). In the following comparison information about the
Pyrenees have been extracted from Choukroune et al. (1989), Roure et al. (1989) and Beamount et al.
(2000). Both mountain ranges show the following features (compare Figs. 6.5, 6.18 and 6.19):
(1) An asymmetrical tectonic double-wedge structural geometry in cross-section. The more
developed wedge (eastern flank of the EC and southern flank of the Pyrenees) is wider than the
other (western flank of the EC and northern flank of the Pyrenees). Correspondingly,
displacement and cumulative shortening is also greater in this wedge. Thrust transport direction
in each wedge is toward the external part of the mountain range. Between the two external
wedges there is an antiformal structure with shallow basement. In the EC this antiformal
structure is the antiformal pop up in the Vlez-Arcabuco anticlinorium. In the Pyrenees the
corresponding is an antiformal stack where basement rocks crop out.
(2) Areas of outcropping basement occur at places where the sedimentary cover has been eroded.
These areas are bounded by strike-slip faults. In the EC the Bucaramanga-Santa Marta and
Altamira faults bound the Santander Massif and Garzn-Quetame Massif, respectively. In the
Pyrenees the basement antiformal stack is bounded by the North Pyrenean Fault.
(3) Two foreland basins are preserved one in each side of the double-wedge. The wider foreland
basin is located adjacent to the more developed wedge. In the EC close to the more evolved
eastern wedge there is the wider LLA Basin. In the Pyrenees close to the more evolved southern
wedge there is the wider Ebro Basin.
(4) A similar geological history: In both areas, extensional basins were developed during the Early
Cretaceous related to extensional faults presently located at both sides of the orogenic double-
wedge. Marine sediments accumulated in the hanging-wall strongly subsiding blocks of these
extensional faults. There was a first stage of convergence and inversion that started to develop
foreland basins. During Paleocene time sedimentation changed from marine to continental in
these foreland basins (although in the EC case this is only applicable to the MV).

Comparison of the regional cross-sections of the EC published by Colletta et al. (1990; Figs.
5.4 and 6.18) and Cooper et al. (1995, Fig. 3.37) with a series of balanced cross-sections of the
Pyrenees showing its evolution through time (Fig 6.19; Choukroune et al., 1989; Choukroune and
ECORS team, 1989; Roure et al., 1989; Beaumont et al, 2000) indicates that the closest similarity is
between the present-day structure of the EC and the Pyrenees during late Ypresian time (early Eocene).
The most remarkable similarity between the two mountain ranges is their asymmetry. However the
comparison should not be extended beyond because in other aspects the EC and the Pyrenees seem to
be different. Similarities between the EC and the Pyrenees (a collision orogen) indicate that the EC has
some features of a collision orogen. However the EC was not formed during collision of two different
continental plates as defined for collision orogens, it was formed by inversion of a Mesozoic
extensional basin. True collision orogens, such as the Alps, have ophiolite suites associated with a
suture formed during the closing of an oceanic basin. However because the EC was not formed by the
closure of an oceanic basin it lacks such ophiolitic suture zones.
The lithospheric structure of the Pyrenees has been deduced by a range of geophysical
techniques (deep reflection and refraction seismic profiles, gravity, magnetotellurics, magnetic
anomalies, tomography and heat flow). Based on these data and geological data, Roure et al. (1989),
Roure (1998), Beaumont et al. (2000) suggested that the orogenic double wedge involves mainly upper
crustal rocks.

245
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

PRABONIAN Continental
piggy-back basins salt sedimentation
Tertiary
Arize Bixoils Montsec 36Ma 50Ma 60Ma 80Ma
foreland basin NPF

Nogueres
90 km
Orri
Rialp

EARLY OLIGOCENE

Arize NPF Bixoils Montsec Serres Marginals 30Ma 36Ma 50Ma 60Ma 80Ma

Nogueres Orri
120 km
Rialp

PRESENT Watershed devide


Serres
Aquitaine basin Arize NPF Nogueres Bixoils Montsec Marginals Ebro basin 20Ma 30Ma 36Ma 50Ma 60Ma 80Ma

Orri Rialp
135 km

50 Km 0

N S
CENOMANIAN
Arize NPF Bixols Montsec Serres Marginals

Nogueres Om Rialp

Pre-collision Mesozoic
PALEOCENE NPF
Mesozoic
Arize Bixols foreland basin 60Ma 80Ma

Noguere s Rialp 35 km
Om

LATE YPRESIAN
Arize NPF Bixols Montsec Serres Marginals 50Ma 60Ma 80Ma

Nogueres
60 km
Om Rialp

50 Km 0

Subducted lower crust

Figure 6.19: Tectonic evolution of the Pyrenees illustrated a series of balanced retrodeformable sections
representing the same regional section at different times. North is in the left side of the sections (from Beaumont et
al., 2000).

246
Chapter 6
According to Beaumont et al. (2000) the crust was decoupled and the lower crust, below the
upper crustal double wedge, was subducted together with the lithospheric mantle into the mantle.
However Roure et al. (1989) and Roure (1998) based on the presence in the tectonic wedge of lower
crustal (Castilln and St. Barthelemy granulites) and upper mantle material (lherzolite of the Lherz
lobe, Roure et al., 1989 and Roure, personal communication) have interpreted that the lower crust was
not subducted due to its buoyancy. According to Roure (1998) crustal balance in the Pyrenees, the Alps
and similar orogenic belts does not require subduction of lower crust. Instead this author proposed that
lower crust has deformed in duplex with triangle geometry. A northward subduction of the Iberian
mantle lithosphere under the European lithosphere has been interpreted from deep seismic crustal
reflection interpretation (Roure, 1998; Beaumont et al., 2000). In this case the subduction vergence
coincides with the vergence of the thrust faults in the more developed wedge.

In contrast data on crustal structure of the EC is limited to gravity and seismicity observations.
In the Andes horizontal shortening observed in the brittle crust has been balanced by a progressive
stacking and thickening of the ductile lower crust in the internal part of the orogen, as indicated by the
configuration of the Moho beneath the orogen (Roure, 1998). This is applicable to the Colombian EC
characterised by a thick crust (Fig. 6.2). A progressive rise of isotherms beneath the Central Andes
(Colombian Central Cordillera with an active magmatic arc) accounts for progressive thinning of the
continental mantle lithosphere, and the anomalous high elevation of the Central Cordillera. In the EC
the presence of small areas of Neogene volcanic rocks at Paipa and Iza (Romero and Rincn, 1990),
elevated geothermal gradients (Fig. 6.10), Tertiary hydrothermalism (Branquet, 1999) and hot springs
(Forero, 1958) may suggest a similar relatively thin mantle lithosphere. Interpretation of flexural
models suggests decoupling between the upper crust and lithospheric mantle. Colletta et al. (1990)
estimated a minimum of 105 km of shortening along the Tunja regional cross-section.
According to these authors this shortening is mostly due to the two thrust fronts. They assumed
also that the upper crust deformed independently from the lithospheric mantle. While the upper crust
formed a double wedge, the mantle lithosphere was probably subducted (Colletta et al., 1990). Taking
the top of the basement as a reference they calculated the cross sectional area of structural relief (1500
km2). With this value and their shortening estimate, they calculated a dcollement at a depth of 14 km
below the top of the basement. Assuming 5 to 8 km of pre-Miocene sedimentary cover, these authors
estimated the depth of dcollement was about 19 to 22 km before deformation. The similarities of the
EC with the Pyrenees may suggest that the EC has a similar deep structure to the Pyrenees. The
conceptual model of the crustal architecture of a collision orogen proposed by Roure (1998, Fig. 6.20)
might be applicable to the deep structure of the EC.
A comparison of the structure of the EC with the Mrida Andes of Venezuela shows
similarities but also important differences:
(1) Both mountain ranges show a double wedge structural geometry in cross-section. The more
developed wedge in the EC related to the LLA flexural basin can be compared to the northern
wedge of the Mrida Andes related to the Neogene flexural basin in the Maracaibo Block
(Colletta et al., 1997). However, the southern wedge of the Mrida Andes is dominated by both
north and south-verging basement-involving structures and the southern Barinas Basin hardly
compares with a flexural basin (Colletta et al., 1997). Between the external wedges there is an
antiformal pop-up structure with shallow basement. This antiformal pop-up structure is the
Vlez-Arcabuco anticlinorium in the EC and the pop-up structures associated with the Bocon
Fault in the Mrida Andes.
(2) Both mountain ranges show internal areas of outcropping basement in places where the
sedimentary cover has been eroded. These areas are bounded by strike-slip faults. In the Mrida
Andes of Venezuela the main strike-slip fault is the Bocon Fault.
(3) Both mountain ranges have a geological history of Mesozoic extensional basin formation
followed by Cenozoic inversion.
(4) Both mountain ranges show an asymmetric structure. The more developed flexural basin is
associated with a strong negative gravity Bouguer anomaly. (Neogene LLA foreland basin for
the EC and Neogene Maracaibo foreland basin for the Mrida Andes). Based on
reflection/refraction seismics and potential field geophysical data Colletta et al. (1997)

247
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

interpreted a south-dipping subduction of the infracontinental lithosphere mantle of the


Maracaibo Block under the Mrida Andes. If the analogy were valid, probably it would imply a
west-dipping subduction of the infracontinental lithosphere mantle of the LLA basin under the
EC.

lt,
au
nF
n ea
P y re
orth ult )
e , N a Fa
p c lin n g
o st ik e sli s ub ri r am a Fronta l thrusts
a lm e s t r . I n B uc a
pu
r ( i.e Mo lasse
Conjugate or
backthrust foreland
b asins
(i.e. Ebro Basin, Llanos Basin)
Foreland basin
(i.e. Po Plain or Aquita ine,
M agdalen a Valley)

Stack lower
crustal duplexes
T
CR US
L OW ER
LO W ER CR
UST

Mantle
Buffer Subduction of the mantle lithosphere

Figure. 6.20: 3D crustal architecture of a collisional orogen: oblique collision and strain partitioning. Figure shows
some similarities of the EC with some collision orogens (modified from Roure, 1998).

6.3. FLEXURAL MODELLING OF BROKEN PLATE AND SUBDUCTION OF THE


MANTLE LITHOSPHERE

Results of flexural models suggest that the LLA lithosphere is stronger than the EC lithosphere.
Taking into account the geological history of the LLA compared to that of the EC, the weaker EC
lithosphere is probably hotter than the LLA lithosphere. If a continental subduction started to develop
between these two lithospheres, probably the colder South American (LLA) lithosphere would be
subducted under the hotter, more buoyant Andean (EC) lithosphere. Hbrard (1985) suggested that
underthrusting of the South American crust was occurring under the EC. In the Alps the more stabilised
and colder European mantle lithosphere subducts below the hotter Apulian mantle lithosphere (c.f.
Roure, 1998).

6.4. COMPARISON WITH ANALOGUE MODEL EXPERIMENTS

In order to infer the deep structure of the EC, Colletta et al. (1990) compared the structure of
the Cordillera with analogue models developed by Malavielle (1984) and Ballard et al. (1987; Fig.
6.21). In both small-scale analogue model experiments, the asymmetry is caused by a velocity
discontinuity along the basal dcollement. The thrust distribution with respect to the velocity
discontinuity is almost the same.
In the Malavielle (1984) experiment the velocity discontinuity corresponds to a west-dipping
subduction. The west-vergent thrusts are permanent faults, the throw of which increases with increasing
shortening. The east-verging thrusts are migrating backthrusts with smaller displacement, which is
opposite to the sense of maximum displacement in the EC (east-verging thrusts). These faults were
successively activated during deformation and their throw is small in comparison to the permanent
thrust zone. The entire model was made up of sand; thus the model has a brittle rheology with no
ductile layers, which is not realistic.

248
Chapter 6
According to Colletta et al. (1990), velocity discontinuity in the Ballard et al. (1987)
experiment represents an east-dipping subduction. In this case, the east-verging thrusts are migrating
major faults, and the west-verging thrusts are only backthrusts with smaller displacement, which is
similar to the EC. The model has interlayered silicone putty, representing the lower ductile crust, which
is a better representation of the lithosphere rheology. In this model the sand layer below the silicone
putty, representing the upper lithospheric mantle developed a west-dipping thrust (Fig. 6.21). Because
the lowermost layer of the model is sand, it did not represent the asthenosphere, which is not realistic,
the velocity discontinuity at the bottom of the model representing an east-dipping subduction is also not
realistic. The west-dipping thrust developed in the lowermost sand layer probably suggest a west-
dipping subduction zone developed on the uppermost mantle contrary to the interpretation of Colletta et
al. (1990). However, the absence of an analogue representation of the asthenosphere in the model
makes this interpretation hypothetical.

a
1 2 3
4 10 cm
5
6 7

S
Malavieille (1984)

1- 7 Order of thrust S Singular point Direction of simulated


development subduction

b
10 cm

Backthrusts Forethruats

Ballard (1987) S
Sand Silicone putty

Figure 6.21: Small-scale analogue models where strong asymmetry in thrust distribution has been simulated by
velocity discontinuity at or close to the basal dcollement: (A) Malavielles model (1984) simulating a west-dipping
subduction. The forethrust zone is reduced but absorbs most of the throw. The backthrust zone is more widely
distributed but the throw of each thrust is slight; (B) Ballard et al. (1987) model, simulating an east-dipping
subduction. Due to the silicone putty, decollement forethrusts can develop and propagate, whereas backthrusts have
small throws (from Colletta et al, 1990).

249
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

6.5. COMPARISON WITH NUMERICAL MODELS OF BEAUMONT ET AL. (2000) AND


ELLIS AND BEAUMONT (1999)

Beaumont et al. (2000) applied to the Pyrenean mountain belt a geodynamical numerical model
that calculates the deformation of a crustal layer controlled by a combination of subduction of the sub-
orogenic lithosphere and surface denudation. Results of the model are comparable to a crustal-scale
restored cross-section of the Pyrenees (Beaumont et al., 2000). The purpose of the models is to
investigate the relative importance of factors controlling the development of the mountain range rather
than to reproduce the details. As Beaumont et al. (2000) pointed out, although the models have been
designed to investigate the evolution of the Pyrenees, the main conclusions are of general interest
because the factors analysed are common to other orogens. I will summarise and discuss the
applicability of these models to the Colombian EC.

6.5.1. Model features

The plain-strain vertical section model crust is deformed under kinematic basal boundary
conditions in which the velocity represents subduction of the mantle lithosphere and part of the lower
crust (Fig. 6.22). In the models mantle detaches from the crust and subducts beneath the stationary
retro-mantle lithosphere (Beaumont et al., 2000). Rheology of the crust in the model is rigid-perfect
plastic with a brittle non-cohesive Coulomb yield criterion (Beaumont et al., 2000). At high
temperatures the deformation mechanism changes from frictional plastic-flow to thermally activated
power law creep (Beaumont et al., 2000). The brittle ductile transition is determined dynamically as
part of the calculation (Beaumont et al., 2000). Beaumont et al. (2000) assume that the minerals which
control ductile creep are wet quartz and wet feldspar. In their models initial crustal thickness is 31
km with upper and lower crustal thicknesses of 16 and 15 km respectively. Flexural isostatic
compensation of the thickened crust is modelled by the bending of two, semi-infinite elastic beams
which support the crust and float on an inviscid asthenosphere (Beaumont et al., 2000). Crust and
mantle densities are typical values. Denudation of the upper surface of the model is calculated at a rate
that depends on the surface elevation. Sediments are deposited in model basins in the same way with a
rate proportional to depth of the basin (Beaumont et al., 2000). Because the parameters of the models
(e.g. crustal thickness, mantle and crustal densities) and tectonic histories of the Pyrenees are similar to
the EC, results of this models can also be applied to the EC.

6.5.2. Results

Figures 6.22 to 6.26 show the results. Beaumont et al. (2000) models P1 and P2 (Figs. 6.22 and
6.23) with laterally constant crustal thickness and strength produce results that are very different to the
Pyrenees and the EC structure.
Their models P3 and P4 (Figs. 6.24 and 6.25) include an inherited mid-crustal weak zone that is
embedded in a strong upper crust. This localised weak zone represents the zone of the former Mesozoic
crustal extension (Beaumont et al., 2000). Although precursory extensional deformation of the crust
and lithosphere is not included in the model and the initial crust has a constant thickness, results of the
models, particularly P3, has many similarities with the Pyrenees and the EC structure. In both (P3 and
P4) models the upper crust is deformed in some way independently of the lower crust: shortening in the
upper crust is consumed by deformation while shortening in the lower crust is consumed by
subduction. However, Roure (based on the surface occurrence of lower crustal rocks in the Pyrenees,

personal communication) believes that subduction of lower continental crust is not realistic. This is also
the case for the Canadian Rockies and the Andes, where all the crustal material necessary to balance
the cover shortening has been stacked at the base of the crust and account for the local roots (Roure,
1998). Model P3 represents a higher degree of upper crust-lower crust-mantle coupling than model P4
with weak coupling between upper and lower crust. The assumption of strong coupling between upper

250
Chapter 6
and lower crust in some of the models (e.g. P3, P5) is debatable in the Pyrenees: From deep seismic
reflection data the lower crust and Moho have been interpreted as important decoupling horizons, with
total detachment of the subcrustal mantle and stacking of ductile lower crust, resulting in uplift of
overlying upper crustal units (Roure et al, 1989; Roure, personal communication). Although this aspect
remains debatable, model P3, particularly in the prediction after 90-120 km of shortening, resembles
the structure of the EC (Fig. 6.5). This is in agreement with the minimum estimate of shortening of 105
km estimated by Colletta et al. (1990). The step-up shear zones developed in the model, equivalent to
major thrust systems at the borders of the EC, are located just at the borders of the weak zone between
them. Clearly if major normal fault systems delimiting the former extensional basin were located in this
position, they would be compressionally reactivated as predicted by the model P3. It is interesting to
note that the location of the shear-up zones predicted by the model coincides with major thrust systems
even when the model does not include the effect of pre-existing crustal discontinuities. Beaumont, et al.
(2000) also emphasised the importance of pre-existing discontinuities in the evolution of the Pyrenees.
Model prediction shows the important effect of an inherited weakness zone. Also model results indicate
that the shear-up zones may signify lateral changes in crustal properties and not the subduction point
(Beaumont et al., 2000). Similar to the analogue models of Malavielle (1984) and Ballard et al. (1987),
the step-shear-up zone dipping in the same direction as the subducted slab is abandoned as it is
advected past the subduction point, and the retro step-up shear zone, representing a backthrust, is more
permanent. In model P3, after 90-120 km of shortening, the shear-up zone with the same vergence as
the subduction zone has the maximum displacement and is adjacent to the wider foreland basin. This
is comparable to the eastern thrust system of the EC adjacent to the wider LLA Foreland. The retro-
shear-up zone in the model is located adjacent to the narrower foreland, comparable to the MV in the
case of the EC. In both the model P3 and the EC the flanks of the orogen are characterised by antiforms
that have been uplifted and eroded and the middle axial region has a synform geometry less uplifted
and eroded. Also in both the model P3 and the EC the topographic profile is asymmetric with
maximum elevation in the wedge underthrusted by the subducted slab (eastern flank in the case of the
EC). The temporal evolution of model P3 also probably is similar to the evolution of the EC.
In model P5 (Fig. 6.26), is similar to P3 except that a subduction load acts to flex the broken
beams beneath the model downward. This load represents the negative buoyancy of the subducted
mantle lithosphere and consequently was increased linearly with the total amount of shortening
(Beaumont et al., 2000). The effect of the subsurface load is to decrease the uplift, which preserves a
basin in the axial region. It is possible to assume that a subsurface subduction load is responsible to the
preservation of the Sabana de Bogot Basin during Neogene time. If such hypothesis is true the palaeo-
topography predicted by flexural models without any subsurface loads is overestimated. It is interesting
to note that some of the geodynamical numerical models with a mid crustal shear zone conducted by
Ellis and Beaumont (1999), predict a local thickening of the lower crust. As already mentioned the
assumption of strong coupling between upper and lower crust of this model is debatable: total
detachment of the subcrustal mantle and ductile lower crust has been interpreted from deep seismic
profiles (Roure et al., 1989; Roure, personal communication).
Further Beaumont et al. (2000) models applied to the Pyrenees explore the role of evaporite
layers to explain some particular features of the Pyrenees not observed in the EC. According to the
results of these models the role of evaporite layers is to create detachment levels. Shear zones
developed on the tips of the inherited weakness zone connect these evaporite detachment levels,
facilitating preservation of piggyback basins. Probably the evaporite levels in the Sabana de Bogot
region facilitated the development of detachment levels that transmitted deformation to the flanks of
the Cordillera, preventing or retarding the surface-uplift of the axial Sabana de Bogota region. This
alternative hypothesis may also explain the late preservation of the Neogene Sabana de Bogot Basin.

251
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

Figure 6.22: In this figure and the following figures, geodynamical models of the deformation of a crustal layer are
controlled by a combination of subduction of the sub-orogenic lithosphere and surface denudation represented by
uniform erosion at a constant rate. In these figures x represents the amount of shortening applied to each model.
Model P1: strong coupling between upper crust and lower crust and between lower crust and mantle (from
Beaumont et al., 2000).

An interesting feature that geodynamical numerical models reveal is that denudation of the
crust at rates approaching the rock-uplift rate can cause rapid exhumation of mid and even lower crustal
rocks at the surface, and create focused shear zones such as the Insubric Line in the Alps (Ellis and
Beaumont, 1999). Probably this effect partly is responsible for the observed association of major strike-
slip faults and basement uplifts (Santander, Garzn-Quetame massifs) in the EC. Also (Ellis and
Beaumont, 1999) models suggest that denudation can maintain the focussed shear deformation. Shear
thrusting deformation in the EC seems to be concentrated in the two major thrust systems in both flanks
of the mountain range (rather than in the central axial zone) where denudation has been large.
Also geodynamical numerical models of oblique collision suggest that when the transcurrent
component of basal velocity is smaller than 1.5 times the normal compressional component, strike-slip
deformation occurs simultaneously with thrusting. If the transcurrent component is greater than 1.5
times the compressional component strain is partitioned with the strike-slip component accommodated
on narrow, sub-vertical flower structures between the two major shear-up zones representing thrust
systems (Ellis and Beaumont, 1999). Regional narrow flower structures have not been described in
the EC, implying that if oblique compression occurred, the transcurrent component was not greater that

252
Chapter 6
1.5 times the normal component. However the possibility of important transcurent motions cannot be
ruled out, due to the lack of detailed structural studies. For example recently Branquet (1999) has
described flower structures in the eastern flank of the Cordillera. Additionally some structural
interpretations such as those by ESRI and Ecopetrol, (1994), Taboada et al. (2000, their Plate 6) and
Schelling (1994) show several pop-up blocks that may be interpreted as flower structures.
If in the EC subduction of mantle lithosphere would occur, the mechanism that initiated such a
subduction remains to be explained. As suggested by Roure (1998), stacking of the lower crust possibly
accounts for the over-thickening of the crust and consequent uplifting due to flexural isostasy.

Figure 6.23: Geodynamical model P2: weakly coupling between upper and lower crust. Other features as explained
in Fig 6.21 (from Beaumont et al., 2000).

253
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

Figure 6.24: Geodynamical model P3: strong coupling between upper and lower crust and between lower crust and
mantle. The assumption of strong coupling between upper and lower crust of this model is debatable (see text for
discussion). This model is characterised by an inherited mid-crustal weak zone. Other features as explained in Fig
6.21 (from Beaumont et al., 2000).

254
Chapter 6

Figure 6.25: Geodynamical model P4: weak coupling between upper and lower crust and strong coupling between
lower crust and mantle. This model also is characterised by an inherited mid-crustal weak zone. Other features as
explained in Fig 6.21 (from Beaumont et al, 2000).

255
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

Figure 6.26: Geodynamical model P5: strong coupling between upper and lower crust and between lower crust and
mantle. The assumption of strong coupling between upper and lower crust of this model is debatable (see text for
discussion). An inherited mid-crustal weak zone was applied as in previous model. This model is characterised by a
subduction load L. Other features as explained in Fig 6.21. The state of the model after 120 km of shortening has
some similarity with the structure of the EC if the pro-foreland basin represents the LLA Basin and the retro-
foreland Basin represents the MV Basin (from Beaumont et al., 2000).

256
Chapter 6

6.6. NEOGENE VOLCANISM AND DEEP-INTERMEDIATE SEISMICITY: TWO


ENIGMATIC FEATURES OF THE EASTERN CORDILLERA

Two particularities of the EC, the local presence of Neogene volcanic rocks and the deep-
intermediate seismicity containing a small, very active zone of seimicity known as the Bucaramanga
earthquake nest (Figs. 6.3, 6.4 and 6.5) may be the expression of a single process.

6.6.1. Neogene magmatism of Paipa Iza

In the northern part of the Tunja axial zone of the Cordillera and south-west of the Floresta
Massif, Neogene volcanic rocks occur between the Boyac and Soapaga faults (Figs 6.3). The volcanic
rocks at Iza intrude Cretaceous and Tertiary sedimentary rocks and form a small dome very close to the
Soapaga Fault (Romero and Rincn, 1990), that probably served as a channel for upward migration of
magma. According to Romero and Rincn (1990) volcanic rocks occur in a zone of Tertiary distension.
The western outcrops near Paipa cover the Gaduas, Socha, Picacho (Maastrichtian to Eocene,
Hbrard, 1985) and Pliocene Tilat formations (Renzoni et al., 1967). Based on stratigraphic relative
position, Fabre (1987) estimated the age of these volcanic rocks as Miocene to Pliocene.
At Iza petrography and chemical analyses indicate that lavas are rhyolites and alkaline rhyolites
(according to the Streckeisen, 1976 classification). According to the silica-alkali diagram (Hyndman,
1974 in Romero and Rincn, 1990) they are alkaline to subalkaline rocks (Romero and Rincn, 1990)
with anomalously high K2O values (7.8-12.0%), and low Na2O values (< 0.8%; Romero and Rincn,
1990). These authors assume an original K2O rich magma. Later intensive potassic metasomatism and
strong hydrothermal alteration affected the rocks. According to Martnez, (1989 in Taboada et al.,
1999) the rhyolites are K/Rb rich and their chemical composition suggests they are linked to partial
melting of the lower crust and not to oceanic subduction.
At Paipa volcanic rocks are rhyolites with a large amount of sanidine indicating potassium
enrichment. Hbrard (1985) documented a strong hydrothermal alteration. According to Romero and
Rincn (1990), Paipa rhyolites are calc-alkaline. Detailed geochemical or isotopic studies are not
available for these Neogene volcanic rocks.
The existence of several hot springs in the central part and eastern flank of the EC (Forero,
1958) may suggest a belt of magmatic activity at depth.

6.6.2. Deep-intermediate seismicity and the Bucaramanga earthquake nest

The Bucaramanga earthquake nest is a zone of intense seismicity beneath the western flank of
the EC, south of Bucaramanga. It is centred at 648 N latitude, 7310 W longitude, and 161-km depth
(Figs. 6.3, 6.4, 6.5, 6.7 and 6.9, in these figures the Bucaramanga nest is located within the cluster
around 150 km depth). The Bucaramanga nest is remarkable not only due to its unusually high-rate of
activity and concentration, but also because of the relative paucity of earthquakes in the surrounding
area (Schneider et al., 1987). However, new data from the National Seismological Network of
Colombia (NSNC) suggests that it is not an isolated phenomenon (Taboada et al., 2000). In a period of
10 years, 75% of earthquakes of magnitude mb 3.8 for the northern Colombia came from the nest
(Schneider et al., 1987). Most seismicity is concentrated in a small volume with a diameter not greater
than 20 km (Santo, 1969; Trygvason and Lawson, 1970; Dewey, 1972), and possibly 5 km (Pennington
et al., 1979, Schneider et al., 1987). The Bucaramanga earthquake nest has the following enigmatic
features in common with similar intermediate depth clusters from the Hindu Kush-Pamir region located
between Pakistan and Afghanistan (Khalturin et al., 1977): (1) Unusually high-rate of activity; (2)
Small volume; (3) Paucity of seismicity in the surrounding area (Schneider et al., 1987). However this
has been questioned by Taboada et al., (2000); (4) High values of Q. Q is a factor for transmission of
seismic energy. Attenuation of seismic energy is inversely proportional to Q (Khalturin et al., 1977;
Pennington et al, 1979; Pennington, 1981); (5) Considerable variability in focal mechanism fault-plane
solutions (Schneider et al., 1987). However, Rivera (1989, in Taboada et al., 1999) reported E-W

257
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

compression and N-S extension from a microseismic experiment around the Bucaramanga nest.
Taboada et al., (2000) confirmed E-W 1 compression, while 2 and 3 were not constrained, and
tectonic stress is in between compressional and strike-slip regimes; (6) Located at complex convergent
plate boundaries. Some of these features are also common to the Vrancea seismic cluster in Romania.
Schneider et al. (1987) showed that the NW-SE lineation of the nest is not an artefact of the location
processes (Fig. 6.3). The shape and size of the nest determined in their microearthquake study, are both
accurate to 10-20% in any dimension and exhibit a predominant NW-SE lineation with no discernible
dip.
Based on 161 intermediate-depth earthquakes Schneider et al. (1987) found that 142 are from
the Bucaramanga nest and only 12% of the total activity and a much smaller proportion of the total
energy released, comes from earthquakes that define an SEE trending alignment from 110 to 190 km
depth (Fig. 6.3). Pennington et al. (1979) found that the region close to the nest is characterised by
anomalously high values of Q. The average Q is the nest region is greater than 1000, which is almost
three times greater than the normal Q values obtained for the upper mantle (Coral, 1985). Focal
mechanisms of the Bucaramanga nest do not show correlation with their position in the nest (Schneider
et al., 1987). Based on the event-magnitude versus frequency relation observed from teleseismic events
and microseismicity Schneider et al. (1987) estimated a cumulative seismic moment rate for the nest of
approximately 5 x 1023 dyn-cm/yr. From the amplitude spectra of two seismic events Coral and
Sarmiento (1986) calculated the stress drop for these events applying the source model proposed by
Brune (1970). This ranges between 1 and 5 bars but is based on seismic moments that are low
compared to those calculated by Schneider et al. (1987).
Pennington (1981) suggested that the sparse intermediate-depth hypocentres outside the
Bucaramanga nest define a single Benioff zone (Figs. 6.4, 6.5 and 6.7). According to Pennington
(1981) the subducted slab is dipping approximately 20 to 25 toward N 109E at depths greater than
140 km. Pennington called this the Bucaramanga segment, running from 5.2 N to 12 N in latitude
and proposed that this subducted lithosphere is apparently continuous with the Caribbean Plate NW of
Colombia. Seismicity associated with this subduction changes orientation north of the Bucaramanga
nest.
New data from the NSNC confirm a NNE (N30E) trending subducting slab segment dipping
35 to 45 toward the SE, running 5.2 N to 7N in latitude along 150 km, from the north of Bogota to the
Bucaramanga nest, beneath the NW flank of the EC (Taboada et al., 2000, Fig. 6.5). This cluster was
not known before the installation of the NSNC. These authors interpret this seismicity cluster as a
subduction of cool and brittle infracontinental lithospheric mantle below the EC. Clearly the
Bucaramanga segment of Pennington (1981) running from 5.2 N to 12 N in latitude include in its
southern part the slab fragment running from 5.2 N to 7 N (trending N30 E and dipping 35 45 )
beneath the EC identified by Taboada et al., (1999, 2000) and a northern slab between 7 N and 12 in
latitude continuous with the Caribbean Plate identified by Prez et al. (1997) (trending NE and dipping
about 20 toward the SE, Fig. 6.7). If the Palaeo-Caribbean slab fragment below the EC is connected
to the Baud-Panam Block as suggested by Taboada et al. (2000) it was broken from the Caribbean
Plate which is subducting with a different velocity vector (Kellogg and Bonini, 1982; Prez et al.,
1997). Seismicity suggest that these two slabs have different subduction angles and geometry (Figs. 6.7
and 6.9). The Bucaramanga earthquake nest is located at the place where these two slabs are
converging with an oblique convergence angle (Fig. 6.9) which may explain the combination of
compressional and strike-slip fault-plane solutions from the Bucaramanga nest. The boundary between
these two slabs define an SEE alignment (Fig. 6.7) parallel to the SEE alignment of the Bucaramanga
nest (Fig. 6.3) identified by Schneider et al. (1987).

6.6.3. Similarities between the Bucaramanga earthquake nest and the Vrancea seismic cluster

Association of deep-intermediate seismic clusters and magmatic activity has also been observed
in the Eastern Carpathians. There, Pliocene-Quaternary alkaline basalts in the Perani Mountains and
Pliocene-Quaternary calc-alkaline high-K andesite and high-K dacite in the Harghita Mountains have

258
Chapter 6
been genetically correlated with the Vrancea seismic zone situated between 70 and 200 km depth
(Grbacea and Frisch, 1998). The rocks are characterised by their high K2O contents (Peccerillo and
Taylor, 1976 in Grbacea and Frisch, 1998), reflecting some similarity with the Neogene volcanic rocks
of the EC. It is interesting to note that in both the EC and in the Eastern Carpathians, both Neogene
volcanism and deep-intermediate seismic clusters are located in zones of maximum curvature of the
mountain belt in map view. Seismic clusters and volcanism also occur in the boundary zone of several
plates or microplates (tectonic blocks) i.e. the European, Moesian and Tisia-Dacia plates in the case of
the Carpathians and the South American Plate and the Andean and Maracaibo tectonic blocks
(microplates?) for the Colombian Andes.
Grbacea and Frisch (1998) have proposed a model involving slab break-off to explain the
association between Neogene volcanism and the seismic Vrancea cluster. Probably a similar
mechanism may explain the Bucaramanga earthquake nest and the Paipa-Iza Neogene volcanic rocks of
the EC (Fig 6.27).

6.6.4. The slab break-off model

This model has been developed to explain syn- to post-collisional magmatism (Davis and
Blanckenburg, 1995), and seismic complexities observed at the Moho depth in some mountain belts
(Sacks and Secor, 1990). In continental collision-type settings, the continental lithosphere is buoyant
and resists subduction, while the cold dense oceanic lithosphere generates a large downward force
(Davis and Blanckenburg, 1995). The net effect is extensional force acting on the transition region. The
strength of the subducting plate decreases in magnitude with increasing subduction. If the deformation
localises, the subducting slab began to brake in a narrow rifting mode, which continues until the
oceanic lithosphere detaches, thus producing slab break-off (Davis and Blanckenburg, 1995; Fig. 6.27).
As a result of rifting of the subducting slab, the asthenosphere upwells into the subsurface rift
generating a shallow thermal perturbation which may melt the metasomatised mechanical lithosphere
(given its lower solidus) and provide the source for syn-collisional magmatism (Davis and
Blanckenburg, 1995; Fig. 6.27). The upwelling asthenosphere could also partially melt the down-going
plate (Davis and Blanckenburg, 1995). The still cold and dense fragment of oceanic lithosphere may be
seismically active due to the pull of its negative buoyancy (Grbacea and Frisch, 1998). Davis and
Blanckenburg (1995) have assessed the plausibility of this process by quantitative models and found
that whether break-off will occur, and the depth at which it will occur, it is a function of temperature
and hence of the subduction velocity. The model has also been applied to explain syn-orogenic
magmatism in the Alps (Blanckenburg and Davies, 1995). For a subduction velocity of 1 cm/yr, break-
off could occur at depths comprised between 50 and 120 km, while at higher velocities it is still likely
to occur, but at greater depths (Davis and Blanckenburg, 1995). Numerical modelling studies (Schott
and Schmeling, in Willingshofer, 2000) have shown that the slab has to reach a depth of 100 170 km
to generate negative buoyancy forces to allow for delamination and slab detachment. With increasing
depth the thermal perturbation and magmatism will reduce (Davis and Blanckenburg, 1995).
In the Colombian case the slab brake-off model may explain (Fig. 6.27): (1) The Neogene
Paipa-Iza volcanism, the presence of several hot springs suggesting an anomalous heat input; (2) The
reduced lithosphere strength suggested by flexural models as compared to standard rheological models;
(3) The deep-intermediate seismic cluster roughly N30 E oriented and dipping towards ESE, where the
Bucaramanga earthquake nest is located; and (4) The rapid surface-uplift of the EC. The presence of a
subducting slab where the Bucaramanga nest is located would suggest that a relatively deep slab break-
off took place and hence could explain the reduced magmatic activity.

259
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

NW Paipa-Iz a Vol cani sm SE


75 W 74 W 72 W
MV EC LLA

100 km

200 km
H

300 km

400 km

0 100 km 200 km
Figure 6.27: Hypothetical deep cross-section of the EC illustrating how the slab break-off model may explain the
intermediate seismicity pattern, the presence of the K/Rb rich Paipa Iza volcanism. The deep subducting slab has
been inferred from mantle tomographic images by Taboada et al. (2000). See text for explanation (near surface
structural cross-section after Colletta et al., 1990; seismicity after Taboada et al., 1999, 2000).

260
Chapter 6
Probably slab break-off of an eastward subducting Caribbean-type mantle lithosphere under the
EC, such as the SEE dipping slab fragment identified by Pennington (1981) and Taboada et al. (1999,
2000) occurs. If this slab represents a fragment of the Palaeo-Caribbean Plate connected to the Baud-
Choc Blcok as suggested by Taboada et al. (2000), its crustal nature is similar to that of the relatively
buoyant Caribbean Plate. Subduction resistance of the relatively buoyant Caribbean type crust (Burke,
1988; Kerr et al., 1997) may have produced break-off of its denser mantle lithosphere under the
Colombian Andes or a fragment of a normal denser lithosphere of the old Farallon Plate (Fig 6.27).
This model would explain the east-dipping slab fragment suggested by seismological studies
(Pennington, 1981; Schneider et al., 1987; Taboada et al., 1999) and the Paipa-Iza volcanism. Although
the SE location of the Paipa-Iza volcanic rocks relative to the seismically active portion of subducted
slab under the EC seems to be inconsistent with the slab break-off model deep tomographic images
interpreted by Taboada et al. (2000) suggest that a deep slab subducts eastward of the EC up to 70 W
in longitude and up to 700 km depth. This can be a fragment of the Palaeo-Caribbean Plate as
suggested by Taboada et al. (2000), or alternatively can be a fragment of a normal denser lithosphere of
the old Farallon Plate. In these tomographic images the seismicity cluster where the Bucaramanga nest
is located occurs just at the place where the slab changes from a shallow dip angle to a high deep angle
(Fig. 6.27). The deep aseismic subducting slab east of the seismicity cluster has about 60 of dip,
suggesting its sinking due to its negative buoyancy. This part of the slab probably was detached and is
no longer connected to the surface plates (Fig. 6.27). The location of of the Paipa-Iza volcanic rocks
may in fact, be related with an upwelling hot asthenospheric flow at the place where the high angle
subduction slab fragment is being rifted from the shallower part (Fig. 6.27). Probably the hot
asthenospheric upwelling generated the thermal anomaly responsible for partial melting of the lower
crust under the EC and weakened its lithosphere. Partial melting of the lower crust may have generated
the Paipa-Iza magmas (Martnez, 1989, in Taboada et al., 1999). This may explain the chemical
composition of the volcanic rocks which is not related to oceanic subduction (Martnez, 1989, in
Taboada et al., 1999). Magmas could have migrated upward and south-eastward along north-west-
dipping upper crustal faults such as the Boyac and Soapaga faults. The location of the Iza volcanic
rocks, very close to the Soapaga Fault (Romero and Rincn, 1990), supports this interpretation. An
alternative hypothesis to explain the observed magmatism could be removal of the lithospheric root by
delamination of the lithospheric mantle (Bird, 1979) However, this would result in extensive
magmatism due to huge amounts of crustal melting, as in the Colorado Plateau.

6.6.5. Need of deep seismic refraction and reflection data

In order to understand the deep structure of the EC, there is a need for deep seismic refraction
and reflection data. All comparisons and inferences about the deep structure of the Cordillera made
here are necessarily merely hypotheses. Deep seismic data could support or contradict the present
hypotheses.

7. CONCLUSIONS

Two major controls on lithosphere dynamics of the extensional basin formation and inversion
history of the EC have been (1) lithosphere rheology, and (2) plate-related tectonic stresses. Standard
rheological models and flexural models indicate that during the Mesozoic and Palaeogene a weak
lithosphere resulted from lithosphere stretching in the area of Mesozoic extensional basins. However,
flexural models indicate a weak lithosphere under the EC in open disagreement with standard
rheological models calculated for the present time, that suggest a strong lithosphere in the EC. Two
important effects not considered in the standard rheological models may explain the difference: (1) The
presence of crustal discontinuities that weaken the lithosphere; and (2) The presence of an anomalous
heat input into the lithosphere, as indicated by local Neogene volcanic rocks and several hot springs in
the EC.
During the Triassic and Jurassic tensional/transtensional stresses, probably initially related to
the break-up of Pangea and later to backarc extension, produced lithosphere stretching and generated
narrow rifts (< 150 km wide), located in the places of the present day MV and the western flank of the

261
Rheological Evolution of the Lithosphere of the Eastern Cordillera and Hypotheses about its
Deep Structure

EC (Magdalena-Tablazo sub-basin). Repeated stretching events in the same area suggest strain
localisation affecting a weak lithosphere. Thermal heating associated with stretching and reactivation of
crustal discontinuities probably contributed to such a strain localisation.
During the Early Cretaceous tensional/transtensional stresses probably related to backarc
extension produced new episodes of lithosphere stretching and generated a wide (> 180 km wide)
system of asymmetric half-rift basins, one of which with a depocenter located along the palaeo-eastern
flank of the EC, with a major normal fault system in its eastern border (Cocuy sub-basin). A system of
horst blocks was located in the area of the Santander and Floresta massifs. A less developed second
order half-rift occurred in the place of the south-western flank of the Cordillera (Magdalena-Tablazo
sub-basin). While stretching of the Magdalena-Tablazo sub-basin affected lithosphere previously
weakened during earlier rifting episodes, development of the new Cocuy cub-basin only can be
explained by reactivation of older crustal discontinuities, probably the pre-Mesozoic Guaicramo
palaeo-fault
During the Palaeogene, compressional/transpresional stresses, probably initially related to
collision of the South American plate margin with a relatively buoyant oceanic plateau and later related
to acceleration of rates of plate convergence and collision to the Choc Block during Neogene time,
affected the weakened lithosphere (of the former extensional basins). This resulted in inversion
episodes of the former extensional basins: During the Palaeogene, incipient basin inversion affected the
weak lithosphere of the former Magdalena-Tablazo and Cocuy sub-basins, while during the Neogene,
the EC resulted from the complete inversion of the Mesozoic extensional basin area (e.g. Colletta et al.,
1990; Cooper et al., 1995; Casero et al., 1995, 1997).
Based on surface geological data, limited geophysical data (gravity, seismicity), comparison
with similar mountain belts and analogue and numerical models from the literature, it is possible to
narrow the uncertainity range on the deep structure of the EC. Probably it resembles that of the
Pyrenees: Lithospheric shortening has been accommodated in the upper brittle crust by development of
a double vergent asymmetric wedge, while the mantle lithosphere accommodated shortening by west-
dipping subduction of the cooler and denser LLA mantle lithosphere under the buoyant and hotter
Andean lithosphere. The lower ductile crust probably accommodated shortening by thickening.
Probably the EC has been strongly affected by transpression with important transcurrent components.
For example Branquet (1999) recently described flower structures in the eastern flank of the
Cordillera. Additionally some structural interpretations such as those by ESRI and Ecopetrol (1994) ,
Schelling (1997) and Taboada et al. (2000, their plate 6) show several pop-up blocks that may be
interpreted as flower structures.
Intermediate seismicity beneath the EC suggest that a subducted slab fragment is present below
its NW margin (Taboada et al., 1999, 2000). A small, but very active zone of intermediate-deep
seismicity, the Bucaramanga earthquake nest (Schneider et al., 1987), may result from deep oblique
convergence of Palaeo-Caribbean plate fragment connected to the Panam Block and the Caribbean
Plate from the north. This may explain the NW-SE lineation of the nest (Schneider et al., 1987). Two
particularities of the EC, i.e. the local presence of Neogene volcanic rocks at Paipa and Iza, and the
intermediate seismicity, may be the expression of slab break-off of the eastward-subducting Caribbean-
type mantle lithosphere fragment under the EC. Subduction resistance of the relatively buoyant
Caribbean type crust (Burke, 1988; Kerr et al., 1997) may have produced break-off of its denser mantle
lithosphere under the Colombian Andes or a fragment of a normal denser lithosphere of the old
Farallon Plate. The deep slab suggested by tomographic images (Taboada et al., 2000) may be a
subducted slab, which is no longer connected to the surface plates. This hypothesis would explain the
east-dipping slab fragment suggested by seismological studies (Pennington, 1981; Schneider et al.,
1987; Taboada et al., 1999, 2000) and the Paipa-Iza volcanic rocks. These volcanic rocks may be
related with upwelling hot asthenosphere where slab break-off occurs. Possibly partial melting of the
lower crust beneath the EC may have generated rising magmas, generated a thermal anomaly under the
EC and weakened its lithosphere.
Deep seismic studies could offer reliable data in order to support or contradict the suggested
hypotheses.

262
REFERENCES

Aalto, K. R., 1971. Petrografa de las areniscas de la seccin estratigrfica de Bogot, Colombia.
Geologa Colombiana, 3: 7-36.
Acosta, J. E., 1999. Structural evolution of the Eastern Cordillera Foothills between Nuncha and Paz
de Ariporo (Colombia). Actas X Congreso Latinoamericano de Geologa y VI Congreso
Nacional de Geologa Econmica, 2: 141-148.
Acosta, J., Coward, M. P. and Lonergan, L., 2000. Kinematics of the western foothills of the
Colombian Eastern Cordillera and regional implications. In press, London, 28 p.
Alfonso, C. A., 1985. Los episodios evaporticos de la Formacin Rosablanca. In: F. Etayo-Serna, F.
and Laverde-Montao (eds.), Proyecto Crtacico, contribuciones. Chapter XVI, Ingeominas
Publicacin Geolgica Especial 16, Bogot, 19 p.
Alfonso, C. A., 1989. Stratigraphy and regional structure of the western flank of the Cordillera
Oriental, Cimitarra area, Colombia. M.Sc. Thesis Univ. of South Carolina, Columbia, SC., 117
p.
Alfonso, C. A. and Ballesteros, I., 1987. Unpublished stratigraphic column, 1:5000 scale, Area 1. In: R.
Ressetar, and C.E. Macellari). ESRI Tech. Rept. 88-0006, 40 p.
Alfonso, C. A., Sacks, P. E. and Secor, Jr. D. T., 1989. Late Tertiary northwestard-vergent thrusting in
Valle del Cauca, Colombian Andes (abs.). A.A.P.G. Bull., 73: 327.
Alvarado, B. and Sarmiento, R., 1944. Informe general sobre los yacimientos de hierro, carbn, y caliza
de la regin de Paz de Ro, Departamento de Boyac. Servicio Geolgico Nacional Informe
468, Bogot, 132 p.
Alvarez, J., 1983. Geologa de la Cordillera Central y el Occidente Colombiano, y petroqumica de los
intrusivos granitoides Mesocenozicos. Ph.D. Thesis Univ. Chile. Boletin Geolgico
Ingeominas, Bogot, 26(2):1-175.
Alzate, J. C. and Roln, L. F., 1996. Modelo estratigrfico secuencial para el Cretcico Inferior en las
reas vecinas a los Macizos de Floresta y Santander (Cordillera Oriental), Memoria VII
Congreso Colombiano de Geologa, Santaf de Bogot, 14 p.
Alzate, J. and Bueno, M., 1994. Anlisis estratigrfico secuencial de las rocas cretcicas de la parte
oriental del Departamento de Boyac, Municipios de Sogamoso, Belencito y Aquitania. Tesis
Pregrado Geologa, Univ. Nal. de Colombia, Depto. de Geociencias, Bogot, 90 p.
Allemand, P. and Brun, J. P., 1991. Width of continental rifts and rheological layering of the
lithosphere. Tectonophysics 188: 63-69.
Allen, P. A. and Allen, J. R., 1990. Basin analysis, principles and applications. Blackwell Scientific
Publications, London, 451 p.
Allmendiger, R. W., Marret., R. A. and Cladouhos, T., 1989. Faultkin, a program for analyzing fault
slip data.
Amaya, S. F. and Santamara I., 1994. Cartografa geolgica y geometra estructural del sector
Peralonso Tetun al Sur de Ortega (Tolima) en el Valle Superior del Magdalena. In: F. Etayo
Serna (ed.), Estudios geolgicos del Valle Superior del Magdalena. Chapter VII, Univ.
Nacional de Colombia, Ecopetrol, Bogot, 20 p.
Amzquita K. F. and Montes, C., 1994. Seccin geolgica del Maco Buenavista: Estructura en el sector
occidental del Valle Superior del Magdalena. In: F. Etayo Serna (ed.), Estudios geolgicos del
Valle Superior del Magdalena. Chapter VI, Univ. Nacional de Colombia, Ecopetrol, Bogot, 36
p.
Anderson, T. A., 1970. Geology of the lower Tertiary Gualanday Group, Upper Magdalena Valley,
Colombia. Ph.D. Thesis, Princeton Univ., Princeton, 86 p.
Anderson, T. A., 1972. Paleogene nonmarine Gualanday Group, Neiva basin, Colombia, and regional
development of the Colombian Andes. Geol. Soc. Am. Bull. 83: 2423-2438.
Andriessen, P. A. M., 1995. Fission-track analysis: principles, methodology and implications for
tectono-thermal histories of sedimentary basins, orogenic belts, and continental margins.
Geologie en Mijnbouw. 74, p. 1-12.
References
Andriessen, P. A. M., Helmens, K. F., Hooghiemstra, H., Riezebos, P. A. and Van der Hammen, T.,
1993. Absolute chronology of the Pliocene-Quaternary sediment sequence of the Bogot Area,
Colombia. Quaternary Sci. Rev. 12: 483-503.
Arango, F. and Venegas, D., 1994. Anlisis Estratigrfico de una Regin al Occidente de la Poblacin
de Paz de Ariporo (Casanare). Tesis Pregrado Geologa, Universidad Nacional de Colombia.
Santaf de Bogot.
Araujo, M., 1997. Experiencias en la exploracin de frente de montaa de la Sierra de Perij, Cuenca
de Maracaibo, Venezuela. VI Simpolsio Bolivariano Exploracion Petrolera en las Cuencas
Subandinas. Mem. Tomo I: 448-456.
Arias, A. and Vargas, R., 1978. Geologa de las planchas 86, Abrego y 97, Cchira, Departamento de
Norte de Santander, Boletn Geolgico Ingeominas Bogot, 23(2): 3-38.
Aspden, J. A. and McCourt, W. J., 1986. Mesozoic Oceanic terrane in the Central Andes of Colombia.
Geology, 14: 415-418.
Aspden, J. A., McCourt, W. and Brook, M., 1987. Geometrical control of subduction-related
magmatism: the Mesozoic and Cenozoic plutonic history of Western Colombia. Jour. Geol.
Soc. London, 144: 893-905.
Audemard, F., 1991. Tectonics of Western Venezuela. Ph.D. Thesis, Rice University, Houston, Texas,
243 p.
Audibert, M., 1991. Dformation discontinue et rotations de blocs. Mthodes numriques de
restauration. Mmoire. Documental. Centre. Armoricain Etudies. Structurales. Socles, Rennes,
France. 40, 239 p.
Av Lallemant, H. G., 1997. Transpression, displacement partitioning, and exhumation in the eastern
Caribbean/South American plate boundary zone. Tectonics, 16(2): 272-289.
Avraham, Z., 1992. Development of asymmetric basins along continental transform faults. In: P.
Ziegler (ed.), Geodynamics of rifting, Volume III, Thematic discussions. Tectonophysics,
215(1-2): 209-220.
Avraham, B. Z. and Nur, A., 1987. Effects of collisions at trenches on oceanic ridges and passive
margins. In: J.W.H. Monger and J. Francheteau (eds.), Circum-Pacific orogenic belts and
evolution of the Pacific Ocean basin. American Geophysical Union, Geodynamics Series, v.
18, p. 9-18.
Baby, P., Rivadeneira, M., Bernal, C., Christophoul, F., Dvila, F. and Galrraga, 1998. Strcutural style
and timing of hydrocarbon entrapments in the Ecuadorian Oriente Basin. A.A.P.G.
International conference and exhibition, abstracts. A.A.P.G. Bull 82(19): 1889.
Bachu, S., Ramn, J.C., Undershultz, J.R. and Villegas, M., 1993. Evaluation of fluid and heat flow
regimes in the Llanos Basin, Colombia. Ecopetrol ICP, Petrocanada Alberta Reseach Council.,
Edmonton, 91 p., 7 appendixes.
Bachu, S., Ramn, J. C., Villegas, M. E. and Underschultz, J. R., 1995. Geothermal regime and thermal
history of the Llanos Basin, Colombia. A.A.P.G. Bull., 79(1): 116-129.
Ballard, J. F., Brun, J. P., Van Den Driesche, J. and Allemand, P., 1987. Propagation des
chevauchements au dessus des zones de dcollement: modles exprimentaux. C.R. Acad. Sci.
Paris, 305, II: 1249-1253.
Ballesteros I. and Nivia A., 1985. La Formacin Ritoque: Registro sedimentario de una albufera de
comienzos del Cretcico. In: F. Etayo-Serna and F. Laverde-Montao (eds.), Proyecto
Crtacico, contribuciones. Chapter XIV, Ingeominas Publicacin Geolgica Especial 16,
Bogot, 17 p.
Bally, A. W., 1984. Tectognse et sismique rflexion. Bull. Soc. Gol. Fr., 24: 279-285.
Barrero, D., 1979. Geology of the Western Cordillera, west of Buga and Roldanillo, Colombia.
Ingeominas Publicacin Geolgica Especial 4, Bogot, 75 p.
Barrientos, A. and Torres, R., 1994. Anlisis paleoambiental de la secuencia del Terciario Inferior
Agua Clara Casanare (Colombia). Tesis Pregrado Geologa Univ. Nacional de Colombia,
Bogot, 80 p.
Barrio, C. A. and Coffield, D. Q., 1992. Late Cretaceous stratigraphy of the Upper Magdalena Basin in
the Payand-Chaparral segment (western Girardot Sub-Basin), Colombia. Jour. South Am.
Earth Sci., 5(2): 132-139.

264
References
Bartels, H., 1986. El "Grupo Palmichal" y su relacin estratigrfica con el Grupo Guadalupe. Tesis
Pregrado Geologa. Univ. Nacional de Colombia, Depto de Geociencias, Bogot, 148 p.
Bayona, G. A., Garca D. F. and Mora G., 1994. La Formacin Saldaa: producto de la actividad de
estratovolcanes continentales en un dominio de retroarco. In: F. Etayo Serna (ed.), Estudios
geolgicos del Valle Superior del Magdalena. Chapter I, Univ. Nacional de Colombia,
Ecopetrol, Bogot, 21 p.
Beaumont, C., 1981. Foreland basins. Geophys. J. R. Astr. Soc., 55: 471-497.
Beaumont, C. and Quinlan, G., 1994. A geodynamic framework for interpreting crustal-scale seismic-
reflectivity patterns in compressional orogens. Geoph. Jour. Int., 116, 754-783.
Beaumont, C., Fullsack, P. and Hamilton, J., 1994. Styles of crustal deformation in compressional
orogens caused by subduction of the underlying lithosphere. Tectonophysics, 232, 119-132.
Beaumont, C., Muoz J. A., Hamilton, J. and Fullsack, P., 2000. Factors controlling the Alpine
evolution of the central Pyrenees inferred from a comparison of observations and geodynamical
models. J. G. R., 105 (B4): 8121-8145.
Beekman, F., 1994. Tectonic modelling of thick-skinned compressional intraplate deformation. Ph.D.
Thesis Vrije Universiteit, Amsterdam, 152p.
Beltran, N. and Gallo, J., 1968. The geology of the Neiva subbasin, Upper Magdalena basin (Southern
portion): 9th Annual Field Conference Guidebook, Colombian Association Petroleum
Geologists and Geophysicists., Bogot, p. 253-274.
Bessis, F., 1986. Some remarks on the study of subsidence of sedimentary basins; application to the
Gulf of Lions margin (western Mediterranean). Marine and Petroleum Geol., 3(1): 37-63.
Bird, P., 1979. Continental delamination and the Colorado Plateau. J. G. R., 84: 7561-7571.
Blanckenburg, F. and Davies J. H., 1995. Slab breakoff: A model for syncollisional magmatism and
tectonics in the Alps. Tectonics, 14(1): 120-131.
Bodine, J. H., 1981. The thermomechanical properties of the oceanic lithosphere, Ph.D. Thesis,
Columbia University, 332 p.
Bogot, J. and Aluja, J., 1981. Geologa de la Serrana de San Lucas. Geologa Norandina, Bogot, 4:
49-55.
Boinet, T., 1985. La frontiere meridionale de la plaque Caraibe aux confins Colombo-Venezuelensis
(Norte de Santander, Colombie: donnes geologiques. These Diplome Docteur 3eme cycle.
L'universite Pierre et Marie Curie, Paris 6, Paris, 200 p.
Boinet, T. H., Bourgois, J. and Mendoza, H., 1982. Tectnica de sobrecorrimiento y sus implicaciones
estructurales en el rea Pamplona, Cordillera Oriental de Colombia. Boletn Geolgico Univ.
Industrial de Santander, Bucaramanga, 15(29): 81.97.
Boinet, T., Bourgois, J., Bellon, H. and Toussaint, J. F., 1985a. Age et rpartition du magmatisme
Prmsozoique des Andes de Colombie. C. R. Acad. Sci. Paris. 300, II(10): 445-450.
Boinet, T, Bourgois, J., Mendoza, H. and Vargas, R., 1985b. Le poinon de Pamplona (Colombie): un
jalon de la frontire mridionale de la plaque Carabe. Bull. Soc. Gol. France, 8(1-3): 403-413.
Bond, G. and Kominz, M. A., 1984. Construction of tectonic subsidence curves for the early Paleozoic
miogeocline, southern Canadian Rocky Mountains: implications for subsidence mechanisms,
age of break-up, and crustal thinning. Geol. Soc. Am. Bull. 95: 155-173.
Bonini, W. E., Hargraves, R. B. and Shagam, R (eds.)., 1984. The Caribbean South American plate
boundary and regional tectonics. Geol. Soc. Am. Memoir 162, 421 p.
Botero-Arango, G., 1963. Contribucin al conocimiento de la geologa de la zona central de Antioquia.
Anales Facultad de Minas., Medelln, 57: 1-101.
Bott, M. H. P., 1992. Passive margins and their subsidence. Jour. Geol. Soc., London, 149: 805-812.
Bourgois J., Azma J., Tournon, J., Bellon, Calle B., Parra, E., Toussaint J. F., Glaon G., Feinberg H.,
De Wever P. and Origlia I., 1982a. Ages et structures des complexes basiques et ultrabasiques
de la facade Pacifique entre 3 N et 12 N (Colombie, Panam et Costa-Rica).Bull. Soc. Gol.
France, 7 t XXIV (3): 545-554.
Bourgois, J., Calle, B., Tournon, J. and Toussaint, J. F., 1982b. The Andean ophiolitic megastructures
on the Buga-Buenaventura transverse (Western Cordillera Valle Colombia). Tectonophysics,
82: 207-229.

265
References
Bourgois J., Toussaint J. F., Gonzales H., Azma J., Calle B., Desmet A., Murcia L. A., Acevedo, A.
P., Parra E. and Tournon, J., 1987. Geological history of the Cretaceous ophiolitic complexes of
Northwestern South America (Colombian Andes). Tectonophysiscs 143: 307-327.
Bourgeois, O., Cobbold, P. R., Rouby, D., Thomas, J. C. and Shein, V., 1997. Least squares restoration
of Tertiary thrust sheets in map view, Tajik depression, central Asia. J. Geophys. Res.
102(B12): 27.553-27.573. Bowin, C., 1976, Caribbean gravity field and plate tectonics, Spec.
Pap. Geol. Soc. Am., Boulder Colorado, 169, 79 p.
Branquet Y., 1999. Etude structurale et mtallognique des gisements dmeraude de Colombie:
Contribution lhistoire tectono-sdimentaire de la Cordillre Orientale de Colombie. Thse de
Doctorat de lInstitut National Polytechnique de Lorraine, Specialit Geosciences, Matires
premires et environement, Institut National Polytechnique de Lorraine, Centre de recherches
ptrographiques et gochimiques (CRPG/CNRS), 265 p
Branquet, Y., Laumonier, B., Lopes, B., Cheilletz, A., Giuliani, G. and Rueda, F., 1996. Evidences of
compressive structures in the Muzo and Coscez emerald deposits, Eastern Cordillera of
Colombia. Third ISAG, St Malo, Abs., p. 675-677.
Branquet, Y., Cheilletz, A., Cobbold, P. R., Baby, P., Laumonier, B. and Guiliani, G., 1999a. Andean
transpressive tectonics at the eastern edge of the Cordillera Oriental, Colombia (Chivor-Guavio
area). Fourth ISAG, 4-6 October 1999, Extended Abstracts, p. 103-105.
Branquet, Y., Chellietz, A., Giuliani, G., Laumonier, B. and Blanco, O., 1999b. Fluidized hydrothermal
breccia in dilatant faults during thrusting, the Colombian emerald deposits. In: K.J.W.,
McCaffrey, L., Lonergan and J., Wilkinson (eds.), Geol. Soc. Spec. Publ. Geol. Soc., London,
155: 183-195.
Braun, J. and Beaumont, C., 1987. Styles of continental rifting: Results from dynamic models of
lihtospheric extension, In: C. Beaumont and A.J. Tankard (eds.), Sedimentary Basins and
Basin-Forming Mechanisms, Can. Soc. Petr. Geol. Memoir, 12: 241-258.
Braun, J. and Beaumont, C., 1989. Contrasting styles of lithospheric extension: implications for
differences between Basin and Range province and rifted continental margins. In: A.J. Tankard
and H.R. Balkwill (eds.), Extensional tectonics and stratigraphy of the North Atlantic margins.
A.A.P.G. Mem. 46: 53-79.
Breen, N. A., 1989. Structural effect of Magdalena fan deposition in the northern Colombia convergent
margin. Geology, 17:34-37.
British Petroleum., 1989. Informe tcnico contrato de evaluacin tcnica Boyac, aos 1987, 1988 y
1989, 16 Vols.; Ecopetrol Bogot, ISN 11297, Informe 2253.
Brune, J. N., 1970. Tectonic stress and the spectra of seismic shear waves from earthquakes. J. G. R.,
75: 4.997-5.009.
Buck, W. R., 1991. Modes of continental lithosphere extension. J. Geophys. Res. , 96: 20.161-20178.
Buitrago, J., 1994. Petroleum systems of the Neiva area, Upper Magdalena Valley, Colombia. In: L. B.
Magoon and W. G. Dow (eds). The petroleum system; from source to trap. A.A.P.G. Mem. 60:
483-497.
Brgl, H., 1958. El Jursico e Infracretceo del ro Bat, Boyac. Boletn Geolgico, Ingeominas,
Bogot, 6(1-3):169-211.
Brgl, H., 1959. Apuntes sobre la estratigrafa de los alrededores de Neiva, Huila, Servicio Geolgico
Nacional, Informe 1318, Bogot, 14 p.
Brgl, H., 1960. El Jursico e Infracretceo del Ro Bat, Boyac: Servicio Geolgico Nacional,
Informe 1319, Boletn Geolgico, Bogot, 1-3: 169-211.
Brgl, H., 1961a. Geologa de los alrededores de Ortega, Tolima. Boletn Geolgico Univ. Industrial de
Santander, Bucaramanga, 8: 21-38.
Brgl, H., 1961b. Historia geolgica de Colombia. Revista Academia Colombiana Ciencias Exactas
Fisicas y Naturales, Bogot, 9(43): 137-191.
Brgl, H., 1961c. Sedimentacin cclica en el geosinclinal Cretceo de la Cordillera Oriental de
Colombia. Boletn Geolgico Servicio Geolgico Nacional, Bogot, 7(1-3): 85-118.
Brgl, H., 1964. El Jura-Trisico de Colombia. Boletn Geolgico Servicio Geolgico Nacional,
Bogot, 12(1-3):5-31.
Brgl, H., 1967. The orogenesis of the Andean system of Colombia. Tectonophysics, 4(4-6): 429-443.

266
References
Burgl, H. and Dumit-Tobon, Y., 1954. El Cretceo Superior en la regin de Girardot. Boletn
Geolgico Instituto Geolgico Nacional, Bogot, 2(1): 23-48.
Burgl, H. and Radelli, L., 1962. Nuevas localidades fosilferas en la Cordillera Central de Colombia.
Geologa Colombiana, Bogot, 3: 133-138.
Burke, K., 1988. The tectonic evolution of the Caribbean. Ann. Rev. Earth Planet. Sci. 16: 210-230.
Burke, K., Cooper, K., Dewey, J., Mann, P. and Pindell, J., 1984. Caribbean tectonics and relative plate
motions. In: W. Bonini, R. Hargraves, R. and Shagam, (eds.), The Caribbean South American
plate boundary and regional tectonics. Geol. Soc. Am. Mem., 162: 31-63.
Burov, E. B. and Diament, M., 1995. The effective elastic thickness (Te) of continental lithosphere:
What does it really mean?. J. Geophys. Res. , 100(B3): 3905-3927.
Burov, E. B., Lobkovsky, L.I., Cloetingh, S. and Nikishin, A.M., 1993. Continental lithosphere folding
in Central Asia (Part II): Constraints from gravity and topography. Tectonophysics, 226: 73-87.
Burtman, V.S. and Molnar, P., 1993. Geological and geophysical evidence for deep subduction of
continental crust beneath the Pamir, Geol. Soc Am. Boulder Colorado, Spec Paper 281, 51 p.
Busbey, A. B., 1986. New material of Sebecus cf. Huilensis (Crocodilia) from the Miocene La Venta
Formation, Colombia. Jour. Vertebrate Paleont., 6(1): 20-27.
Butler, K., 1983. Andean-type foreland deformation: structural development of the Neiva Basin, Upper
Magdalena Valley, Colombia. Ecopetrol, Bogot, Informe Geolgico 1509.
Butler, K. and Schamel, S., 1988. Structure along the eastern margin of the Central Cordillera, Upper
Magdalena Valley, Colombia. Jour. South Am. Earth Sci., 1(1): 109-120.
Butler, W.H., Harris, B.W. and Whittington, G., 1997. Interactions between deformation, magmatism
and hydrothermal activity during active crustal thickening, a field example from Nanga Parbat,
Pakistan Himalayas. Mineralogical Magazine, London, 61(1): 37-52.
Byerlee, J. D., 1978. Friction in rocks. Pure Appl. Geophys. 116: 615-626.
Cceres, C. and Etayo-Serna, F., 1969. Bosquejo geolgico de la regin del Tequendama, Opsculo
gua de la Excursin Pre-Congreso. 1er Congreso Colombiano de Geologa, Opsculo, Bogot,
23 p.
Caicedo J. and Roncancio, J. H., 1994. El Grupo Gualanday como ejemplo de acumulacin sintectnica
en al Valle Superior del Magdalena durante el Paleoceno. In: F. Etayo Serna (ed.), Estudios
geolgicos del Valle Superior del Magdalena. Chapter X, Univ. Nacional de Colombia,
Ecopetrol, Bogot, 20 p.
Calvache, J. A. and Muoz., F. A., 1984. Modelo de la corteza terrestre en un rea aledaa al municipio
de Cumaral. Departamento del Meta. Tesis Pregrado Geologa, Univ. Nacional de Colombia.
Depart. Geoc., Bogot. 116 p.
Campbell, C. J., 1968. The Santa Marta wrench fault of Colombia and its regional setting, Trans. IV
Caribb. Geol. Conf., Trinidad, p. 247-262.
Campbell, C. J., 1974. Colombian Andes. In: A. M. Spencer (ed.), Mesozoic-Cenozoic orogenic belts:
data for orogenic studies: Edinburgh, Scottish Academic Press, p. 705-724.
Campbell, J. C. and Brgl, H., 1965. Section through the Eastern Cordillera of Colombia, South
America. Geol. Soc. Am. Bull., 76: 567-590.
Cardona J. and Guterrez Z., 1995. Estratigrafa y ambientes de depsito de la Formacin Carbonera en
un rea al noroeste de Yopal-Casanare (Colombia). Tesis Pregrado Geologa, Univ. Nacional
de Colombia, Bogot, 80 p.
Cardozo E. and Ramrez C., 1985. Ambientes de depsito de la Formacin Rosablanca: rea de Villa
de Leiva. In: F. Etayo-Serna and F. Laverde-Montao (eds.), Proyecto Crtacico,
contribuciones. Chapter XIII, Ingeominas Publicacin Geolgica Especial 16, Bogot, 13 p.
Cardozo, E. and Sarmiento, L. F., 1987. Unpublished stratigraphic column, 1:5000 scale, Area 3. In: R.
Ressetar, and C. E. Macellari. ESRI Tech. Rept. 88-0006, 40 p.
Cardozo, N. F. and Ziga, J., 1995. Anlisis estructural de las zonas del bloque colgante (hanging-
wall) y bloque yacente (foot-wall) de la Falla de Soapaga entre Corrales y Paz de Ro (Boyac),
Tesis Pregrado Univ. Nacional de Colombia, Empresa Colombiana de Petrleos Ecopetrol,
Bogot, 90 p.
Cardozo-Puentes, E., 1989. Structural style of the central western flank of the Cordillera Oriental west
of Bogot, Colombia. M.Sc. Thesis, University of South Carolina, Columbia S.C., 146 p.

267
References
Carillo, V., 1982. La Quebrada Confines (Municipio de Cucutilla, Norte de Santander), una nueva
localidad fosilfera del Paleozico Superior en la Cordillera Oriental de Colombia. Geologa
Norandina, Bogot, 5: 32-38.
Carter, N. L. and Tsenn, M. C., 1987. Flow properties of continental lithosphere. Tectonophysics, 136:
27-63.
Casero, P., Salel, J. F. and Rossato, A., 1995. Structural evolution of the margin and foothills belt of the
Cordillera Oriental of Colombia (southwest of Cusiana). VI Congreso Colombiano del
Petrleo. Abs., Bogot, 1: 33-41
Casero, P., Salel, J. F. and Rossato, A., 1997. Multidisciplinary correlative evidences for polyphase
geological evolution of the foothills of the Cordillera Oriental (Colombia). VI Simposio
Bolivariano Exploracin Petrolera en las Cuencas Subandinas, Cartagena, Abs., 7.Tomo II:
100-118.
Castillo, J. E. and Mojica, J., 1990. Determinacin de la orientacin de los esfuerzos actuales a partir de
elongaciones tectnicas (breakouts) en algunos pozos petroleros de los Llanos Orientales y
del Valle Medio del Magdalena, Colombia. Geologa Colombiana, Bogot, 17: 123-132.
Cazier, E. C., Hayward, A. B., Espinosa, G., Velandia, J., Mugniot, J. F. and Leel Jr W. G., 1995.
Petroleum geology of the Cusiana field, Llanos Basin Foothills. Colombia, A.A.P.G. Bull.,
79(10): 1444-1463.
Cazier, E. C., Cooper, M. A., Eaton, S. G. and Pulham, A. J., 1997. Basin development and tectonic
history of the Llanos Basin, eastern Cordillera, and Middle Magdalena Valley, Colombia:
reply. A.A.P.G. Bull., 81: 1332-1335.
Cediel, F., 1968. El Grupo Girn, una molasa Mesozica de la Cordillera Oriental. Servicio Geolgico
Nacional Boletn Geolgico, Bogot, 16(1-3): 5-96.
Cediel, F., 1982. Modelo estructural para los Llanos Orientales. Simposio Exploracin Petrolera en las
Cuencas Subandinas de Venezuela, Colombia, Ecuador y Per. Colombian Society Petroleum
Geologists and Geophysicists, Bogot, 8 p.
Cediel, F., 1983. El borde continental Colombiano durante el Trasico-Jursico y el origen de la Placa
Caribe. Abstract. Mem. 10th Caribbean Geological Conference, Ingeominas, Bogot, p. 170.
Cediel, F., Mojica, J. and Macia, C., 1981. Las formaciones Luisa, Payand, Saldaa sus columnas
estratigrficas caractersticas. Geologa Norandina, Bogot, 3: 11-19.
Cediel, F., Etayo, F. and Cceres, C., 1997. Distribucin de facies sedimentarias y su marco tectnico
durante el Fanerozico en Colombia. IV Simposio Bolivariano Exploracin Petrolera en las
Cuencas Subandinas, Cartagena, Septemb. 14-17, Tomo II: 34-37.
Cspedes S. and Pea L., 1995. Relaciones estratigrficas y ambientes de depsito de las formaciones
del Terciario Inferior aflorante entre Tunja y Paz de Rio (Boyac), Tesis Pregrado Geologa,
Univ. Nacional de Colombia Depart. Geoc., Santaf de Bogot, 50 p.
Chandra, U., 1981. Focal mechanism solutions and their tectonic implications for the Eastern Alpine-
Himalayan Region. In H. K. Gupta and F. M. Delany (Eds.), Zagros. Indu Kush. Himalaya.
Geodynamic evolution. Geodynamic Series volume 3. Am. Geoph. Union, Washington, D. C.
Geol. Soc. Am., Boulder Colorado: 243-271.
Chapman, D. S., 1986. Thermal gradients in the continental crust, In: J.B. Dawson, D.A., Carswell, J.
Hall, and K.H., Wedepolhl (eds.), The Nature of the Lower Continental Crust. Geological
Society Spec. Publ., 24: 63-70.
Cheilletz, A., Giuliani, G. and Arhan, T., 1993. Late Eocene-Oligocene episode in the Eastern
Cordillera of Colombia viewed by emerald dating. II Symp. Internat. Godynamique Andine,
ISAG 93, Oxford, ORSTOM. Paris, p 473-476.
Cheilletz, A., Giuliani, G., Branquet, Y., Laumonier, B., Snchez, M.A.J., Rueda, F., Fraud, G. and
Arhan, T., 1997. Datation K-Ar et 40Ar/39Ar 65 M.a. des gisements dmeraude du district
Chivor-Macanal, argument en faveur dune dformation prcoce dans la Cordillre Orientale de
Colombie. C.R. Acad. Sci., Sries II, Paris, 324(5): 369-377.
Chigne, N. and Rojas, E., 1997. El piedemonte de la Cordillera Oriental de Colombia y de los Andes de
Mrida: Estilos estructurales y consideraciones sobre la gnesis y migracin de hidrrocarburos.
VI Simposio Bolivariano Exploracin Petrolera en las Cuencas Subandinas. Mem. Tomo I:
457-477.

268
References
Choukroune, P. and ECORS Team., 1989. The ECORS Pyrenean deep seismic profile reflection data
and the overall structure of an orognic belt. Tectonics 8(1): 23-39.
Choukroune, P., Pinet, B., Roure, F. and Cazes, M., 1989. Major Hercynian structure along the ECORS
Pyrenees and Biscaye lines. Bull. Soc. Gol. France.
Cisternas, A. and Gaulon, R., 1984. Sntesis sismotectnica de nordeste de Venezuela, Revista de
Geofsica, Madrid, 40(1): 3-10.
Clavijo, J., 1985. La secuencia de la Formacin Los Santos en la Quebrada Piedra Azul: registro de una
hoya fluvial evanescente. In: F. Etayo-Serna and F. Laverde-Montao (eds.), Proyecto
Crtacico, contribuciones. Chapter IV, Ingeominas Publicacin Geolgica Especial 16, Bogot,
18 p.
Clavijo, J., 1995. Mapa geolgico de Colombia, plancha 75 Aguachica Escala 1:100.000, Memoria
explicativa. Ingeominas, Bucaramanga, 48 p.
Clavijo, J., Barbosa, G., Bernal, L.E., and others, 1992. Mapa geolgico plancha 75 Aguachica, Scale
1:100.000. Bucaramanga.
Cloetingh, S., 1988. Intra-plate stress: a new element in basin analysis. In: K.L. Kleinspehn and C.
Paola (eds.), Frontiers in sedimentary geology New perspectives in basin analysis. Springer-
Verlag, New York, NY, p. 205-230.
Cloetingh, S. and Banda, E., 1992. Europes lithosphere physical properties: mechanical structure. In:
Blundell, D., Freeman, R., and S. Mueller (eds.), A continent revealed the European
geotraverse. Cambridge University Press, European Science Foundation, Cambridge, p. 80-91.
Cloetingh, S. and Kooi, H., 1992. Intraplate stress and dynamical aspects of rift basins. In: P. A. Ziegler
(ed.), Geodynamics of rifting, v. II. Thematic discussions, Tectonophysics, 215: 167-185.
Cloetingh, S. and Burov, E., 1996. Thermomechanical structure of European continental lithosphere;
constraints from rheological profiles and EET estimates. Geophys. Jour. Int., 124: 695-723.
Cloetingh, S., McQueen, H. and Lambeck, K., 1985. On a tectonic mechanism for regional sea level
variations. Earth Plan. Sci. Lett., 75: 157-166.
Cloetingh, S., Sassi, W. and Horvath, F., (eds.), 1993. The origin of sedimentary basins: Inference from
quantitative modeling and sedimentary basin analysis. Tectonophysics, 226, 518 p.
Cloetingh, S., Eldholm, O., Larsen, B. T., Gabrielsen, R. H. and Sassi, W., 1994. Dynamics of
extensional basin formation and inversion: introduction. Tectonophysics, 240: 1-9.
Cloetingh, S., Burov, E. and Poliakov, A., 1999. Lithosphere folding: Primary response to
compression? (from central Asia to Paris Basin). Tectonics, 18(6): 1064-1083.
Cloos, M., 1993. Lithospheric buoyancy and collisional orogenesis: subduction of oceanic plateaus,
continental margins, island arcs, spreading ridges and seamounts. Geol. Soc. Am. Bull., 105:
715-735.
Cobbold, P. R. and Percevault, M. N., 1983. Spatial integration of strains using finite elements. Jour.
Struct. Geol., 5: 299-305.
Cobbold, P. R., Renzoni, G. and Ulloa, C., 1988. Cenozoic tectonics of the Cordillera Oriental: A
working hypothesis. Ingeominas Internal Report. Bogot, 8 p.
Colletta, B., Hbrard F., Letouzey, J., Werner, P. and Rudkiewikz, J. L., 1990. Tectonic style and
crustal structure of the Eastern Cordillera (Colombia), from a balanced cross section. In: J.
Letouzey (ed.), Petroleum and tectonics in mobile belts. Editions Technip, Paris 11990, p. 81-
100.
Colletta B., Letouzey, J., Faure, J. L. and Casero, P., 1995. Kinematic and physical modeling of thrusts.
Comparisons with the Colombian foothills structures. VI Congreso Colombiano del Petrleo.
Bogot, Abs. 1 p. 151-154.
Colletta, B., Roure, F., De Toni, B., Loureiro, D. and Passalacqua, H., 1997. Tectonic inheritance,
crustal architecture, and contrasting structural styles in the Venezuelan Andes. Tectonics,
16(5): 777-794.
Cooper, M. A., Addison, F. T., Alvarez, R., Coral, M., Graham, R. H., Hayward, A. B., Howe, S.,
Martinez, J., Naar, J., Peas, R., Pulham, A. J. and Taborda, A., 1995. Basin development and
tectonic history of the Llanos Basin, Eastern Cordillera, and Middle Magdalena Valley,
Colombia, A.A.P.G. Bull., 79 (10): 1421-1443.

269
References
Coral, C., 1985. Contribucin al estudio de la actividad ssmica en Santander (Colombia), Memorias VI
Congreso Latinoamericano de Geologa, Tomo II: 271-292.
Coral, C. and Sarmiento, L., 1986. Resultados preliminares del estudio de parmetros focales en la
regin de Santander (Colombia), Revista CIAF, Bogot, 11(1-3): 435-455.
Corredor, F., 1997. Evidence for a passive roof duplex along the Northeastern thrust front of Eastern
Cordillera, Colombia, and implications for oil exploration. VI Simposio Exploracin Petrolera
en las Cuencas Subandinas, Cartagena, Tomo II: 42-47.
Corrigan, H.T., 1989. Introduction to the geology of the Sabana and southern protions. 26th Field
Conference. Texaco. In: Colom. Soc. Petrol. Geol. Geoph. (Ed.), Geological Field Trips,
Colombia 1980-1989. Colombian Society Petroleum Geologists and Geophysicists, Bogot,
312-332, 1 appendix.
Cortes, R. and De la Espriella, R., 1984. Contibucin al conocimiento del Paleozico Superior en la
seccin Quetame-Villavicencio. Boletn Geolgico Univ. Industrial de Santander,
Bucaramanga, 16(30): 83-101.
Covey, M. C. and Dengo, C. A., 1987. Structural evluation of exploration leads in the Paz de Ro and
Tunja blocks, Eastern Cordillera, Colombia. Ecopetrol, Bogot, ISN 12535, Informe 2310.
Cross, T. A. and Pilger, R. H., 1982. Controls of subduction geometry, location of magmatic arcs, and
tectonics of arc and back-arc regions. Geol. Soc. Am. Bull. 93: 545-562.
Cross T. A., Anderson, D. S., McDonough K. J., Ramn, J. C., Sanchez, A. and Olaya, I., 1996a.
Stratigraphic framework of the Middle Magdalena Basin, Report prepared for Ecopetrol and
Instituto Colombiano del Petrleo by Strategic Stratigraphy, Inc Evergreen, Colorado, 39 p.
Cross T. A., Anderson, D. S., McDonough K. J., Ramn, J. C., Sanchez, A. and Olaya, I., 1996b.
Middle Magdalena Basin stratigraphic play assessment, Phase 2, Report prepared for Ecopetrol
and Instituto Colombiano del Petrleo by Strategic Stratigraphy, Inc Evergreen, Colorado, 8 p.
Cuervo, E. A. and Ramrez, A., 1985. Estratigrafa y ambiente de sedimentacin de la Formacin
Cacho en los alrededores de Bogot. Tesis Pregrado Geologa. Univ. Nacional de Colombia,
Depto. de Geociencias, Bogot, 100 p.
Czaplewski, N. J., 1989. Miocene bats from La Venta fauna, Colombia. In: W. D., Krause and R. D.,
Kenneth (eds.), Abstracts for papers: Forty-ninth annual meeting, Assoc. of Vertebrate Paleont.
Jour. Vertebrate Paleont. 9(3) Supple. p. 18A.
Daconte, R. and Salinas, R., 1980. Geologa de las planchas 66, Miraflores y 76, Ocaa, Colombia.
Ingeominas, Bogot, Informe 1844: 1-105.
Dahlstrom, C. D. A., 1969. Balanced cross-sections. Can. Jour. Earth Sci., 6: 743-757.
Daly, M. C., 1989. Correlations between Nazca/Farallon Plate kinematics and forearc basin evolution
in Ecuador. Tectonics, 8: 769-790.
Daniels, A. T., 1991. Stress and strain in the Northern Andes from earthquake data. M.Sc. Thesis,
Univ. of South Carolina, Columbia, S.C., 100 p.
Dashwood, M. F. and Abbotts, I. L., 1990. Aspects of the petroleum geology of the Oriente Basin,
Ecuador, In: J. Brooks (Ed.), Classic petroleum provinces. Geol. Soc. Spec. Publ. London, 50:
89-117.
Davis, J. H. and Blanckenburg, F., 1995. Slab breakoff: A model of lithosphere detachment and its test
in the magmatism and deformation of collisional orogens. Earth and Planet. Scien. Lett., 129:
85-102.
Davy, Ph. and Gillet, Ph., 1986. The stacking of thrust slices in collision zones and its thermal
consequences. Tectonics, 5: 913-929.
De Freitas, M. G., Francolin, J. B. L. and Cobbold, P. R., 1997. The structure of the axial zone of the
Cordillera Oriental, Colombia. IV Simposio Bolivariano Exploracin Petrolera en las Cuencas
Subandinas.Asociacion Colombiana Geologos y Geofsicos del Petrleo, Bogot, p. 38-41.
De Jong, K., 1991. Tectono-metamorphic studies and radiometric dating in the Betic Cordilleras (SE
Spain) with implications for the dynamics of extension and compression in the Western
Mediterranean Area. Ph.D. Thesis, Vrije Universiteit, Amsterdam, 204 p.
De la Espriella, R. and Cortes, R., 1982. Contribucin a la estratigrafa del Grupo Quetame. IV Congr.
Colombiano de Geologa. Cali, (In press) 15 p.

270
References
Dengo, C. A. and Covey, M. C., 1993. Structure of the Eastern Cordillera of Colombia: Implications
for trap styles and regional tectonics, A.A.P.G. Bull., 77(8): 1315-1337.
Desgaulx, P., Roure, F. and Villien, A., 1990. Structural evolution of the Pyrenees: Tectonic heritage
and flexural behaviour of the continental crust. In J Letouzey (ed.), Petroleum and tectonics in
mobile belts. IFP Exploration and production research conferences, Editions Technip, Paris, p.
31-48.
Desgaulx, P., Kooi, H. and Cloetingh, S., 1991. Consequences of foreland basin development on
thinned continental lithosphere: Applications to the Aquitane Basin (SW France). Earth Planet.
Sci. Lett., 106: 116-132.
Dewey, J. F., 1980. Episodicity, sequence, and style at convergent plate boundaries. In: D.W.
Strangeway (ed.), The continental crust and its mineral deposits, Spec. Paper Geol. Assoc.
Canada, 20: 553-573.
Dewey, J. F. and Pindell, J. L., 1985. Neogene block tectonics of eastern Turkey and northern South
America; continental applications of the finite difference method. Tectonics, 4(1): 71-83.
Dewey, J. W., 1972. Seismicity and tectonics of western Venezuela, Bull. Seism. Soc. Am., 62: 1.711-
1.752.
Diaz L., 1994a. Distribucin de las facies siliciclsticas correspondientes a la Formacin Arenisca
Tierna y equivalents en el Valle Superior del Magdalena. In: F. Etayo Serna (ed.), Estudios
geolgicos del Valle Superior del Magdalena. Chapter IV, Univ. Nacional de Colombia,
Ecopetrol, Bogot, 15 p.
Diaz L., 1994b. Reconstruccin de la Cuenca del Valle Superior del Magdalena, a finales del Cretcico.
In: F. Etayo Serna (ed.), Estudios geolgicos del Valle Superior del Magdalena. Chapter XI,
Univ. Nacional de Colombia, Ecopetrol, Bogot, 13 p.
Daz, J. M. and Sotelo, C. I., 1995. Anlisis estructural de la Falla de Boyac en un rea al oeste de los
Municipios de Paipa-Duitama. Tesis Pregrado Geologa, Univ. Nacional de Colombia,
Depart. Geoc., Bogot, 75 p.
Dickey, P. A., 1941. Pre-Cretaceous sediments in Cordillera Oriental of Colombia. Bull. A.A.P.G.,
25(9): 1789-1795.
Dickinson, W. R., 1985. Interpreting provenance relations from detrital modes of sandstones. In: G. P.,
Zuffa (ed.), Provenance of arenites. NATO ASI Series C. Mathematical and Physical Sciences
148: 333-361.
Divies, R., 1997. FOLDIS, un modle cinmatique de bassins sdimentaires par lments discrets
associants plis, failles, rosion/sdimentation et compaction. Ph.D. Thesis. Univ. of Grenoble I,
Grenoble, 220 p.
Dorado, J., 1984. Contribucin al conocimiento de la estratigrafa de la Formacin Brechas de
Buenavista (lmite Jursico Cretcico), regin noroeste de Villavicencio (Meta). Geologa
Colombiana, Bogot, 17: 7-40.
Dueas, H. and Castro, G., 1981. Asociacin palinolgica de la Formacin Mesa en la regin de Faln,
Tolima, Colombia. Geologa Norandina., Bogot, 3: 27-36.
Duncan R. A. and Hargraves R. B., 1984. Plate tectonic evolution of the Caribbean region in the mantle
reference frame. In: W.E. Bonini, R.B. Hargraves and R. Shagam (eds). The Caribbean-South
American plate boundary and regional tectonics, Geol. Soc. Am., Mem., 162: 81-93.
Duque-Caro, H., 1979. Major estructural elements and evolution of northwestern Colombia. In: J.S.
Watkins, L. Montadert, and P.W. Dickerson. (Eds.), Geological and geophysical investigations
of continental margins. A.A.P.G. Mem. 29: 329-351.
Duque-Caro, H., 1984. Structural style, diapirism, and accretionary episodes of the Sin-San Jacinto
terrane, southwestern Caribbean borderland. Geol. Soc. Am. Mem. 162: 303-316.
Duque-Caro, H., 1990. The Choc block in the northwestern corner of South America: Structural,
tectonostratigraphic, and paleogeographic implications. Jour. of South Am. Earth Sci., 3(1): 71-
84.
Duval, B. C. and Valds, G. E., 1995. Giant oil fields of the 80s associated with an 'A' subduction in S.
America. Oil and Gas Jour., July 1995: 67-70.
Ebinger, C. J., 1989. Geometric and kinematic development of border faults and accommodation zones,
Kivu-Rusizi rift, Africa. Tectonics, 8: 117-133.

271
References
Ecopetrol., 1970. Mapa Geolgico de los Llanos Orientales (rea Pisba - Tmara), Esc.1:50.000.
Ecopetrol-ICP., 1995. rea volcanera: evaluacin tcnica de subsuelo. Ecopetrol-ICP, Division de
Exploracin y Explortacin, EXE, Piedecuesta, 42 p.
Ecopetrol-ICP., 1996. Informe Proyecto Estratigrafa del Terciario del Valle Medio del Magdalena.
Compact disc., Piedecuesta.
Ecopetrol, Vicepresidencia de Exploracin., 1997. Topes de formaciones de pozos del Valle del
Magdalena. Informe interno, Santaf de Bogot, 100 p,
Ecopetrol and Beicip., 1995. Cuenca de los Llanos Orientales, Estudio geolgico regional, Santaf de
Bogot, Volume 1, Text, 198 p., Volume 2, figures.
Ecopetrol, Esso and Exxon Exploration Company., 1994. Integrated technical evaluation Santander
sector Colombia., 1991-1994. Report, text and figures. Final Report, Houston, 38 p., 49 figs.
Ellis, S. and Beaumont, C., 1999. Models of convergent boundary tectonics, implications for the
interpretation of Lithoprobe data, Canadian Jour. of Earth Sci., 36(10): 1.711-1.741.
England, P. and Molnar, P., 1990. Surface uplift, uplift of rocks, and exhumation of rocks. Geology,
18: 1173-1177.
England, P. C. and Thompson, A. B., 1986. Pressure-temperature-time paths of regional metamorphism
I. Heat transfer during the evolution of regions of thickened continental crust. Jour. Petrol., 25:
894-928.
ESRI - HOCOL., 1989. Proyecto de Evaluacin Tcnica del Flanco Oriental de los Valles Medio y
Alto del Magdalena. Informe Geolgico de Ecopetrol Nro. 2424.
ESRI and Ecopetrol., 1994. Geology and hydrocarbon potential of the Cordillera Oriental Colombia, 4
volumes. Earth Sciences and Resources Institute, University of South Carolina. ESRI Technical
Report 94-07-418, Columbia, SC., 500 p.
Estrada, A., 1972. Geology and plate tectonics history of the Colombian Andes. M.Sc. Thesis Stanford
Univ., California, 115 p.
Etayo-Serna, F., 1968. El sistema Cretceo en la regin de Villa de Leiva y zonas prximas. Geologa
Colombiana, Bogot, 5: 5-74.
Etayo-Serna, F., 1979. Zonation of the Cretaceous of Central Colombia by ammonites. Publicacin
Geolgica Especial Ingeominas, Bogot, 2: 1-186.
Etayo-Serna, F., 1985a. Documentacin paleontolgica del Infracretcico de San Flix y Valle Alto,
Cordillera Central. In: F. Etayo-Serna and F. Laverde-Montao (eds.), Proyecto Crtacico,
contribuciones. Chapter XXV, Ingeominas Publicacin Geolgica Especial 16, Bogot, 7 p.
Etayo-Serna, F., 1985b. El lmite Jursico Cretcico en Colombia. In: F. Etayo-Serna and F. Laverde-
Montao (eds.), Proyecto Crtacico, contribuciones. Chapter XXII, Ingeominas Publicacin
Geolgica Especial 16, Bogot, 4 p.
Etayo-Serna, F., 1985c. Paleontologa estratigrfica del Sistema Cretcico en la Sierra Nevada del
Cocuy. In: F. Etayo-Serna and F. Laverde-Montao (eds.), Proyecto Crtacico, contribuciones.
Chapter XXIII, Ingeominas Publicacin Geolgica Especial 16, Bogot, 47 p.
Etayo-Serna, F., 1994. Eplogo: A modo de historia geolgica del Cretcico del Valle Superior del
Magdalena. In: F. Etayo Serna (ed.), Estudios geolgicos del Valle Superior del Magdalena.
Chapter XX, Univ. Nacional de Colombia, Ecopetrol, Bogot, 6 p.
Etayo-Serna, F. and Laverde-Montao, F., (eds.), 1985. Proyecto Crtacico, contribuciones.
Ingeominas, Publ. Geol. Esp. 16, Bogot, 200 p.
Etayo-Serna, F. and Rodrguez G., 1985. Edad de la Formacin Los Santos. In: F. Etayo-Serna and F.
Laverde-Montao (eds.), Proyecto Crtacico, contribuciones. Chapter XXVI, Ingeominas
Publicacin Geolgica Especial 16, Bogot, 13 p.
Etayo-Serna, F. and Florez, J. M., 1994. Estratigrafa y estructura de la Quebrada Calambe y el cerro El
Azucar, Olaya Herrera Tolima. In: F. Etayo Serna (ed.), Estudios geolgicos del Valle Superior
del Magdalena. Chapter XII, Univ. Nacional de Colombia, Ecopetrol, Bogot, 23 p.
Etayo, F., Renzoni, G. and Barrero, D., 1976. Contornos sucesivos del mar Cretcico en Colombia.
Primer Congreso Colombiano de Geologa, Mem., Bogot, p. 217-252.
Etayo-Serna, F., Barrero, D., Lozano, H., Espinosa, A., Gonzlez, H., Orego, A., Zambrano, F., Duque,
H., Vargas, R., Nez, A., lvarez, J., Ropan, C., Ballesteros, I., Cardozo, E., Forero, H.,

272
References
Galvis, N., Ramrez, C. and Sarmiento, L., 1983. Mapa de terrenos geolgicos de Colombia,
Ingeominas Publicacin Geolgica Especial 14, Bogot, 235 p.
Eurocan., 1988. Informe operacional y tcnico para Ecopetrol 1988. Contrato de Asociacin Tunja.
Ecopetrol Informe No 2133, Bogot, 20p.
Fabre, A., 1981. Estratigrafa de la Sierra Nevada del Cocuy, Boyac y Arauca, Cordillera Oriental
(Colombia). Geologa Norandina, Bogot, 4: 3-12.
Fabre, A., 1983a. La subsidencia de la Cuenca del Cocuy (Cordillera Oriental de Colombia) durante el
Cretceo y el Terciario Inferior. Primera parte: Estudio cuantitativo de la subsidencia. Geologa
Norandina, Bogot, 8: 49-61.
Fabre, A., 1983b. La subsidencia de la Cuenca del Cocuy (Cordillera Oriental de Colombia) durante el
Cretceo y el Terciario Inferior. Segunda parte: Esquema de evolucin tectnica. Geologa
Norandina, Bogot, 8: 21-27.
Fabre A., 1985a. Dinmica de la sedimentacin Cretcica en la regin de la Sierra Nevada del Cocuy
(Cordillera Oriental de Colombia). In: F. Etayo-Serna and F. Laverde-Montao (eds.), Proyecto
Crtacico, contribuciones. Chapter XIX, Ingeominas Publicacin Geolgica Especial 16,
Bogot, 20 p.
Fabre, A., 1985b. Subsidencia y maduracin de la materia orgnica; un modelo de la evolucin de la
Cordillera Oriental y los Llanos durante el Cretcico y el Terciario (abstract). Mem. IV Congr.
Latinoamericano de Geologa., Bogot, 2: 421-422.
Fabre, A., 1986. Gologie de la Sierra Nevada del Cocuy (Cordillre Orientale de Colombie). Thse
Universit de Genve de Docteur es Sciences de la Terre, No 2217, Genve, 394 p.
Fabre, A., 1987. Tectonique et gneration dhydrocarbures: Un modle de lvolution de la Cordillre
Orientale de Colombie et du Bassin des Llanos pendant le Crtac et le Tertiaire. Arch. Sc.
Genve, 40 (Fasc. 2): 145-190.
Fabre, A. and Delaloye, M., 1983. Intrusiones bsicas Cretcicas de la Cordillera Oriental. Geologa.
Norandina, Bogot, 6:19-28.
Fajardo, A., 1995. 4-D stratigraphic architecture and 3-D reservoir fluid-flow model of the Mirador
Formation, Cusiana Field, foothills area of the Cordillera Oriental, Colombia. MSc. Thesis
Colorado School of Mines, Golden, 171 p.
Fajardo A., Rubiano J. L. and Reyes A., 1993. Estratigrafa de secuencias de las rocas del Cretceo
Tardo al Eoceno Tardo en el sector central de la cuenca de los Llanos Orientales,
Departamento del Casanare, Report ECP-ICP-001-93, Piedecuesta, Santander, 69 p.
Fajardo, G. and Chamorro, M. E., 1994. Interpretacin geolgica y geofsica Proyecto Tunja-94.
Ecopetrol, Informe Geol. 3046. Santaf de Bogot.
Fajardo-Pea, G., 1998. Structural analysis and basin inversion evolutionary model of the Arcabuco,
Tunja and Sogamoso regions, Eastern Cordillera, Colombia. M.Sc. Thesis, University of
Colorado, Boulder, 114 p.
Faure, G., 1986. Principles of isotope geology. John Wiley, New York, 589 p.
Feininger, T., 1970. The Palestina Fault, Colombia. Geol. Soc. Am. Bull., 81(4): 1201-1216.
Feininger, T., 1980. Cretaceous and Paleogene geologic history of coastal Ecuador. Geologishe
Rundschau, Bd. 69 (Heft. 3): 849-874.
Feininger, T., 1985. Allochthonous terranes in the Andes of Ecuador and northwestern Per.
Geological Survey of Canada, I Observatory Crescent, Ottawa, Ontario. Canada KIA OY3.
Feininger, T., 1986. Allochthonous terranes in the Andes of Ecuador and northwestern Per. Can. Jour.
of Earth Sci., 24: 266-278.
Findlay, A. L., 1988. Cortes geolgicos regionales en los Andes Venezolanos y sus implicaciones
petroleras y estructurales. Presentacin por afiches. III Simposio Bolivariano Exploracin
Petrolera en las Cuuencas Subandinas, Caracas, Memorias, Tomo I: 401-429.
Flint, S., Turner, P., Jolley, E. J. and Hartley, A. J., 1993. Extensional tectonics in convergent margin
basins: an example from the Salar de Atacama, Chilean Andes. Geol. Soc. Am. Bull. 105: 603-
617.
Flrez J, M. and Carrillo G. A., 1994. Estratigrafa de la sucesin litolgica basal del Cretcico del
Valle Superior del Magdalena. In: F. Etayo Serna (ed.), Estudios geolgicos del Valle Superior
del Magdalena. Chapter II, Univ. Nacional de Colombia, Ecopetrol, Bogot, 26 p.

273
References
Fleh, E. R., 1973. Additional gravity interpretation along the active plate boundary in southweast,
Colombia. In: J. E., Ramirez and L. T.Aldrich (eds.), La transicin oceano-continente en al
suroeste de Colombia, Proyecto Cooperativo Internacional Nario, Univ. Javeriana, Bogot, p.
199-204.
Fleh, E. R., Milkereit, B., Meissner, R., Meyer, R. P., Ramrez, J. E., Quintero, J. and Udias, A., 1981.
Observaciones de refraccin ssmica en el Noroeste Colombiano en la latitud 5.5 N. In:
Goberna, J.R. (ed.), Investigaciones geofsicas sobre las estructuras oceano-continentales del
Occidente Colombiano, Proyectos Cooperativos Internacionales Nario II 1976, Nario III
1978. Instituto Geofsico Univ. Javeriana, Bogot, p. 83-96.
Fllmi K. B., Garrison R. E., Ramirez P. C., Zabrano-Ortz, Kennedy W. J. and Lehner B. L., 1992.
Cyclic phosphate-rich successions in the Upper Cretaceous of Colombia. Palaeogeography,
Palaeoclimatology, Palaeoecology, 93: 151-182.
Forero, H., 1958. Fuentes termales de Colombia. Informe 1295, Servicio Geolgico Nacional, Bogot,
25 p.
Forero, H. and Sarmiento L., 1985. La facies evaportica de la Formacin Paja en la regin de Villa de
Leiva. In: F. Etayo-Serna and F. Laverde-Montao (eds.), Proyecto Crtacico, contribuciones.
Chapter XVII, Ingeominas Publicacin Geolgica Especial 16, Bogot, 16 p.
Freymueller, J. T., Kellogg, J. N. and Vega, V., 1993. Plate motions in the North Andean region. J.
Geophys. Res. , 98 (B12): 21.853-21.863.
Galindo, V., Garzn, H., Rueda, M. C. and Brown, P., 1994. Evaluacin del potencial de
entrampamiento estructural en el Cretceo (Cuenca del Valle Supeior del Magdalena),
Ecopetrol, Vicepresidencia de exploracin y produccin. Gerencia de Exploracin , 22 p.
Anexo Regional overview of the Upper Magdalena Valley: Tectonic evolution by P. Brown. 26
p. Ecopetrol Informe Geolgico, Santaf de Bogot, 15 p.
Galvis, N. and Rubiano J., 1985. Redefinicin estratigrfica de la Formacin Arcabuco, con base en
anlisis facial. In: F. Etayo-Serna and F. Laverde-Montao (eds.), Proyecto Crtacico,
contribuciones. Chapter VII, Ingeominas Publicacin Geolgica Especial 16, Bogot, 16 p.
Garca-Castellanos, D., 1998. Desarrollo de modelos numricos de flexin litosfrica. Ph.D. Thesis.
Institut de Cincies de la Terre Jaume Almera. Consejo Superior de Investigaciones Cientficas
(CSIC), Barcelona, 166 p.
Garca-Gonzlez. M., Velandia-Guevara, J. and Galindo-Torres, G., 1997. El potencial de generacin
de gas y petrleo de la Formacin Guaduas de la Cordillera Oriental, y de la Formacin
Guachinte-Ferreira de el Valle del Cauca. VI Simposio Bolivariano Exploracin Petrolera en
las Cuencas Subandinas, Cartagena, Abs. 14.
George. R. P., Pindell, J. L. and Cristancho, J. H., 1997. Eocene paleostructure of Colombia and
implications for history of generation and migration of hydrocarbons. IV Simposio Bolivariano
Exploracin Petrolera en las Cuencas Subandinas: Asociacin Colombiana Geologos y
Geofsicos del Petrleo, Bogot, Tomo I: 133-139.
Geotec., 1976. Mapa geolgico de Colombia. Scale 1: 1.000.000. Ed. Geotec, Bogot, Colombia.
Geotec., 1992. Facies distribution and tectonic setting through the Phanerozoic of Colombia. A regional
synthesis combining outcrop and subsurface data presented in 17 consecutive rock-time slices.
Bogot.100 p.
Geotec., 1994. Stratigraphy and tectonic study of the Upper and Middle Magdalena Basins Colombia,
Report for Maxus, Bogot, 55 p.
Germeraad, J. H., Hoping, C. A. and Muller, J., 1968. Palynology of Tertiary sediments from tropical
area. Rev. Paleobotan. Palynology, 6: 139-348.
Geyer, O., 1976. La fauna de amonitas del perfil tpico de Morrocoyal (Lisico inferior). I Congr.
Colombiano de Geologa, Bogot, Mem.: 111-134.
Geyer, O., 1982. Comparaciones estratigrficas y faciales en el Trisico Norandino. Geologa.
Norandina, Bogot, 5: 27-31.
Gibbs, A. D., 1987. Linked tectonics of the Northern North Sea basin. In: C. Beaumont and A. J.
Tankard (eds.), Sedimentary basins and basin forming mechanisms. Can. Soc. Petrol. Geol.,
Mem. 12: 163-171.

274
References
Grbacea, R. and Frisch, W., 1998. Slab in the wrong place: Lower lithospheric mantle delamination in
the last stage of the eastern Carpathian subduction retreat. Geology, 26(7): 611-614.
Goetze, C. and Evans, B., 1979. Stress and temperature in the bending lithposhere as constrained by
experimental rock mechanics. Geophys. J. Roy. Astr. Soc., 59: 463-478.
Goldsmith, R., Marvin, R. F. and Mehnert, H. H., 1971. Radiometric ages in the Santander Massif,
Eastern Cordillera, Colombian Andes, U.S. Geol. Surv. Prof. Paper, Reston, V.A. P0750-D:
D44-D49.
Gomez, E., 1997. Quantified Cenozoic evolution of Cenozoic deformation of the Middle Magdalena
(Colombia). A.A.P.G. Bull., 81(10): 1774.
Gmez, E., 1999. Geodinmica, tectonoestratigrafa, cronologa y evolucin termal de cuencas
sedimentarias complejas y de sus reas fuente. Mtodos modernos de anlisis cuantitativo y
aplicaciones, implicaciones para exploracin y produccin de hidrocarburos. Curso-Taller con
salida de campo. Santaf de Bogot, Nov. 26 30., 55 p.
Gmez E. and Pedraza P., 1994. El Maastrichtiano de la regin Honda Guaduas, lmite N del Valle
Superior del Magdalena: Registro sedimentario de un delta dominado por ros trenzados. In: F.
Etayo Serna (ed.), Estudios geolgicos del Valle Superior del Magdalena. Chapter III, Univ.
Nacional de Colombia, Ecopetrol, Bogot, 20 p.
Gmez, E. and Jordan, T., 1998. Relative timing of Cenozoic deformation of the Middle Magdalena
Valley Basin and bounding mountain ranges, Colombia A.A.P.G./SEPM Annual Meeting,
Abstr. with Progr., Salt Lake City, Utah.
Gomez, E., Jordan, T., Hegarty, K. and Kelley, S., 1999. Diachronous deformation of the Central and
Eastern Andean cordilleras of Colombia and syntectonic sedimentation in the Middle
Magdalena Valley Basin. Fourth ISAG, Gottingen, Germany, 4-6 October, p. 287-290.
Gradstein, F. M. and Ogg, J. G., 1996, A Phanerozoic time scale. Episodes, 9(1-2): 3-5.
Gratier, J. P., Guillier, B. and Delorme, A., 1991. Restoration and balance of a folded and faulted
surface by best-fitting of finite elements: Principle and applications. Jour. Struct. Geol., 13: 11-
115.
Grupo de Geoqumica-ICP., 1994. Evaluacin geoqumica Cuenca Valle Superior del Magdalena.
Informe final presentado a la Gerencia de Exploracin Ecopetrol. Proyecto Valle Superior del
Magdalena, Piedecuesta, 195p.
Guatame, R. A. and Lara, H. Y., 1995. Anlisis geoqumico de las formaciones pre-Albianas Fmeque
y Tibasosa en un rea al norte de Tunja entre los Municipios de Pesca, Nobsa, Santa Rosa de
Viterbo, Beln y Betitiva. Tesis Pregrado Geologa, Univ. Nacional de Colombia, Depto.
Geoc., Bogot, 100 p.
Guerrero, J., 1993. Magnetostratigraphy of the upper part of the Honda Group and Neiva Formation.
Miocene uplift of the Colombian Andes. Ph.D. Thesis, Duke University, 108 p. Appendix
Orthogonal Vector plots, 108 p.
Guerrero, J. and Sarmiento, G., 1996. Estratigrafa fsica, palinologca, sedimentolgica y secuencial
del Cretcico Superior y Paleoceno del Piedemonte Llanero, Implicaciones en Exploracin
Petrolera. Geologa Colombiana, 20: 3-66.
Guillande, M. R., 1988. Evolution Mso-Cnozoque dune valle intercordillraine andine: La Haute
Valle du Rio Magdalena (Colombie), Ph.D. Thesis Univer. Paris 6, Paris, 352 p.
Guzmn, G., 1985. Los grifidos infraCretcicos Aetostreon couloni y Ceratostreon boussingaulti de la
Formacin Rosablanca, como indicadores de occilaciones marinas, In: F. Etayo-Serna and F.
Laverde-Montao (eds.), Proyecto Crtacico, contribuciones. Chapter XII, Ingeominas
Publicacin Geolgica Especial 16, Bogot, 16 p.
Hall, R., Alvarez, J. and Rico, H., 1972. Geologa de los departamentos de Antioquia y Caldas
(subzona II-A). Boletn Geolgico Ingeominas, Bogot, 20(1): 1-85.
Haq, B. U., Hardenbol, J. and Vail, P. R., 1987. Chronology of fluctuating sea levels since the Triassic
(250 Myr ago to Present). Science, 235: 1156-1167.
Harry, D. L. and Sawyer, D. S., 1992. Basaltic volcanism, mantle plumes, and the mechanics of rifting:
The Paran flood basalt province of South America. Geology 20: 207-210.

275
References
Hathon, L. and Espejo, I., 1997. Detrital modes of Tertiary units, Central Middle Magdalena Basin,
Colombia. IV Simposio Bolivariano, Exploracin Petrolera en las Cuencas Subandinas.
Cartagena. Abs., 16.
Hayashida, A., 1984. Paleomagnetic study of the Miocene continental deposits in La Venta badlands,
Colombia.
Hbrard F., 1985. Les foothills de la Cordillre Orientale de Colombie entre les rios Casanare et
Cusiana. Evolution godynamique depuis dEo Crtac. Ph.D. Thesis, Univ. Pierre et Marie
Curie, Paris, 162 p.
Hbrard, F., Leikine, M., Bourgois, J., Ricateau, R. and Charitat, P., 1987a. Metamorphisme et
diagense des sries Crtaces et Tertiaires du versant est de la Cordillre Orientale de
Colombie, un exemple dvolution thermique lie louverture dun bassin au Crtac. Eclog.
Geol. Helv.
Hbrard, F., Leikine, M., Bourgois, J., Ricateau, R. and Charitat, P., 1987b. Mtamorphisme et
diagense des series Crtaces et Tertiaires du versant est de la Cordillre Orientale de
Colombie, un exemple dvolution thermique lie louverture dun bassin au Crtac. II
Simposio Bolivariano Exploracin Petrolera en las Cuencas Subandinas, Bogot, Abs.
Helmens, K. F., 1988. Late Pleistocene glacial sequence in the area of the high plain of Bogot (Eastern
Cordillera, Colombia). Palaeogeography, Palaeoclimatology, Palaeoecology, 67: 263-283.
Helmens, K. F., 1990. Neogene-Quaternary geology of the high plain of Bogot, Eastern Cordillera,
Colombia (stratigraphy, paleoenvironments and landscape evolution). Diss. Botanicae 163, J.
Cramer, Berlin, 202 p.
Helmens, K., Barendregt, R., Randolph, E., Baker, J. and Andriessen, P. A. M., 1997. Magnetic
polarity and fission-track chronology of a late Pliocene-Pleistocene paleoclimatic proxy record
in the tropical Andes. Quaternary Research, 48: 15-28.
Henneberg, H. E., 1983. Geodetic control of neotectonics in Venezuela. Tectonophysics, 97: 1-15.
Hill, R. I., 1993. Mantle plumes and continental tectonics. Lithos 30: 193-206.
Hirsfeld S. E. and Marshall, L. G., 1976. Revised fauna list of the La Venta Fauna (Friasian, Miocene)
of Colombia, South America. Jour. Paleontology, 50: 433-436.
Hoogiemstra, H., 1984. Vegetational and climatic history of the high plain of Bogot, Colombia: a
continuous record of the last 3.5 million years. Diss. Botanicae, 79, 368 p.
Hooghiemstra, H., 1989. Quaternary and Upper-Pliocene glaciations and forest development in the
tropical Andes: evidence from long high-resolution pollen record from the sedimentary basin of
Bogot, Colombia. Palaeogeography, Palaeoclimatology, Palaeoecology, 72: 11-26.
Hoorn, C., 1988. Quebarada Mochuelo, type locality of the Bogot Formation: A Sedimentological,
petrographical and palynological study. Hugo de Vries Lab., Univ. of Amsterdam, Amsterdam,
21 p.
Hossack, J., Martnez, J., Estrada, C. and Herbert, R., 1999. Structural evolution of the Llanos fold and
thrust belt, Colombia. In: K. McClay (ed.), Thrust Tectonics 99 Meeting, Royal Halloway
Univ. of London, April 26-29, June 1999, Program, London, p. 110.
Howe, M. W., 1969. Geologic studies of the Mesa Group (Pliocene?), Upper Magdalena Valley,
Colombia. Ph.D. Thesis, Princeton Univ., Princeton, 67 p.
Howe, M. W., 1974. Nonmarine Neiva Formation (Pliocene?), Upper Magdlena Valley, Colombia:
regional tectonism. Geol. Soc. Am., Bull., 85: 228-246.
Hubach, E., 1957. Estratigrafa de la Sabana de Bogot y alrededores. Boletn Geolgico, Servicio
Geolgico Nacional, Bogot, 5(2): 93-112.
ICP-GEX-DYA, Grupo Neiva, 1995. Estudio hidrodinmico de la Formacin Caballos en la Subcuenca
de Neiva. Cuenca Valle Superior del Magdalena. Informe final, Piedecuesta, 39 p.
Irving, E. M., 1971. La evolucin estructural de los Andes mas septentrionales de Colombia. Boletn
Geolgico Ingeominas, Bogot, 19 (2): 1-190.
Isacks, B. and Molnar, P., 1971. Distribution of stress in the descending lithosphere from a global
survey of focal mechanism solutions of mantle earthquakes, Rev. Geoph., 9: 103-174.
Jaillard, E. P., Solar, P., Carlier, G. and Mourier, T., 1990. Geodynamic evolution of the northern and
central Andes during early to midle Mesozoic times: a Tethyan model. Jour. Geol. Soc.,
London, 147: 1009-1022.

276
References
Janssen, M., Stephenson, R. A. and Cloetingh, S., 1995. Temporal and spatial correlations between
changes in plate motions and the evolution of rifted basins in Africa. Geol. Soc. Am. Bull., 107:
1317-1332.
Jaramillo, J. M., 1978. Determinacin de las edades de algunas rocas de la Cordillera Central de
Colombia por el mtodo de huellas de fisin. 2do. Congreso Colomb. Geol. Resmenes,
Bogot, P. 19-20.
Jaramillo, J. M., 1981. Determinacin de las edades de algunas rocas de la Cordillera Central de
Colombia por el mtodo de huellas de fisin. Boletn Ciencias de la Tierra, Medelln, (5-6):
145-146.
Jaramillo L, Roa. E. and Torres M., 1993. Relaciones estratigrficas entre las unidades arenosas del
Paleogeno (Paleoceno) del piedemonte llanero y la parte media de la Cordillera Oriental. Tesis
Pregrado Geologa Univ.Nacional de Colombia, Bogot, 80 p.
Jaramillo C. and Ypez O., 1994. Palinostratigrafa del Grupo Olini (Coniaciano-Campaniano), Valle
Superior del Magdalena, Colombia. In: F. Etayo Serna (ed.), Estudios geolgicos del Valle
Superior del Magdalena. Chapter XIII, Univ. Nacional de Colombia, Ecopetrol, Bogot, 20 p.
Jarvis, G. T. and McKenzie, D., 1978. Sedimentary basin formation with finite extension rates. Earth
Planet. Sci. Lett., 48: 42-52.
Jones, P. B., 1995. Geodynamic evolution of the Eastern Andes, Colombia An alternative hypothesis.
In: A.J. Tankard, R. Surez S., and H. J., Welsnik (eds.), Petroleum basins of South America:
A.A.P.G. Mem., 62: 647-658.
Jordan, T. and Alonso, R., 1987. Cenozoic stratigraphy and basin tectonics of the Andes Mountains,
20 - 28 south latitude. A.A.P.G. Bull., 71: 49-64.
Jordan, T. and Gomez, E., 1996. Proposal of Middle Magdalena Valley stratigraphic study of Tertiary
sediments to Ecopetrol. Introduction, Ecopetrol, Bogot, 12 p.
Julivert, M., 1958a. La morfoestructura de la zona de mesas al SW de Bucaramanga (Colombia S.A.).
Boletn Geolgico Univ. Industrial de Santander, Bucaramanga, 1: 7-43.
Julivert, M., 1958b. Geologa de la zona tabular entre San Gil y Chiquinquir, Cordillera Oriental,
Colombia. Boletn Geolgico Univ. Industrial de Santander, Bucaramanga, 2: 33-45.
Julivert, M., 1959. Geologa de la vertiente W del macizo de Santander en el sector de Bucaramanga.
Boletn Geolgico Univ. Industrial de Santander, Bucaramanga, 3: 15-33.
Julivert, M., 1960. Geologa de la regin occidental de Garca Rovira. Boletn Geolgico Univ.
Industrial de Santander, Bucaramanga, 5: 5-32.
Julivert, M., 1961. Las estructuras del Valle Medio del Magdalena y su significacin. Univ. Industrial
de Santander, Bucaramanga, Boletn Geolgico, 6: 33-52.
Julivert, M., 1962. La estratigrafa de la Formacin Guadalupe y las estructuras por gravedad en la
serrana de Cha, Sabana de Bogot. Boletn Geolgico Univ. Industrial de Santander,
Bucaramanga, 11: 5-21.
Julivert, M., 1968. Colombie. Lexique stratigraphique International. Precambrian, Paleozoique et
Mesozoique. Paris, vol. V, Fasc. 4a, premiere partie: 651 p.
Julivert, M., 1970. Cover and basement tectonics in the Cordillera Oriental of Colombia, and a
comparison with other folded chains. Geol. Soc. Am. Bull., 81: 3623-3646.
Julivert, M., Barrero, D. and Navas, J., 1964. Geologa de la Messa de Los Santos. Boletn Geolgico
Uiv. Industrial de Santander, Bucaramanga, 18: 5-11.
Kaila, K. L., 1981. Structure and seismotectonics of the Himalaya-Pamir Hindukush region and the
Indian Plate boundary. In H. K. Gupta and F. M. Delany (Eds.), Zagros. Indu Kush. Himalaya.
Geodynamic evolution. Geodynamic Series volume 3. Am. Geoph. Union, Washington, D. C.
Geol. Soc. Am., Boulder Colorado: 272-293.
Kammer, A., 1993a. Las Fallas de Romeral y su relacin con la tectnica de la Cordillera Central.
Geologa Colombiana, Bogot, 18: 27-46.
Kammer, A., 1993b. Steeply dipping basement faults and associated structures of the Santander Massif,
Eastern Cordillera, Colombian Andes. Geologa Colombiana, Bogot, 18: 47-64.
Kammer, A., 1996. Estructuras y deformaciones del borde oriental del Macizo de Floresta. Geologa
Colombiana, Bogot, 21: 65-80.

277
References
Kammer, A., 1999a. Structural styles of the folded Bogot segment, eastern Cordillera Oriental of
Colombia. Fourth ISAG, Gottingen, Germany, Oct. 4-6, 1999, Extended Abstracts, p. 381-384.
Kammer, A., 1999b. Observaciones acerca de un origen traanspresivo de la Cordillera Oriental.
Geologa Colombiana, Bogot, 24: 29-53.
Kammer, A., 2001. Tectnica extensiva Jursica al margen continental Norandina y el origen de la
Falla de Bucaramanga. Resumen Coloquio Trisico-Jursico en Colombia. Ingeominas, Univ.
Indust. Santander, Sociedad Colombiana de Geologa. Bucaramanga, 1 p.
Kammer, A. and Mojica, J., 1995. Los pliegues de la Barrera de GuataquiGirardot: Producto de un
despegue de la Cobertera Cretcica de la Cordillera Oriental?. Geologa Colombiana, Bogot,
19: 33-48.
Kammer, A. and Mojica, J., 1996. Una comparacin de la tectnica de basamento de las cordilleras
Central y Oriental. Geologa Colombiana, Bogot, 20: 93-106.
Karner, G. D. and Watts, A. B., 1983. Gravity anomalies and flexure of the lithosphere at mountain
ranges. J. Geophys. Res. , 88: 10449-10477.
Kay R. F., Madden, R. H. and Guerrero, J., 1989. Nuevos hallazgos de monos in el Mioceno de
Colombia. Ameghiana, 25(3): 203-212.
Keal, J. E., 1985. Source rock studies of the Eastern Cordillera adjacent to the Llanos Basin, Colombia.
Mem. II Simposio. Bolivariano Exploracin Petrolera en las Cuencas Subandinas, Bogot, 2: 1-
19.
Keeley, M. L. and Sarez, A. E., 1997. Exploring the sub-thrust play along the eastern margin of the
Middle Magdalena Vallley: The example of Chawina (Apulo), Cundinamarca, Central
Colombia. VI Simposio Bolivariano Exporacin en las Cuencas Subandinas, Cartagena de
Indias, Mem. Tomo I: 167-184.
Kellogg, J., 1984. Cenozoic tectonic history of the Sierra de Perij, Venezuela-Colombia, and adjacent
basins. In : W. E. Bonini, R. B. Hargraves and R. Shagam (eds.), The Caribbean South
American plate boundary and regional tectonics, Geol. Soc. Am., Mem., 162: 239-261.
Kellogg, J. and Duque, H. S., 1994. Crustal scale deformation in the Eastern Cordillera, Colombia. V
Simposio Bolivariano Exploracin Petrolera en las Cuencas Subandinas. Puerto La Cruz,
Memoria: 365-372.
Kellogg, J., Ogujiofor, I. J. and Kansakar, D. R., 1985. Cenozoic tectonics of the Panama and North
Andes blocks, Mem. Sixth Latin American Geological Congress., Bogot, 1: 40-59.
Kellogg, J., Godley, V. M., Ropan, C., Bermdez, A. and Aiken, C. L. V., 1991. Gravity field of
Colombia, eastern Panam and adjacent marine areas. Geol. Soc. Am. Map and Chart Series,
MCH070.
Kellogg, J., Ojeda, G., Duque, H., Roth, S., Scott, M., Zhao, Ji, K., Diaz, D. and Gladdish, D., 1995.
Colombia geophysical project Eastern Cordillera Technical report. An interdisciplinary project
to interpret the geology and geopysics of Colmbia's Eastern Cordillera, foothills and Vichada.
Andean Geophysical Laboratory Univ South Carolina, Columbia S.C., 166 p.
Kerr, A. C., Mariner, G. F., Arndt, N. T., Tarney, J., Nivia, A., Saunders, A. D., Storey, M. and
Duncan, R. A., 1996. The petrogenesis of Gorgona komatiites, picrites and basalts: new field,
petrographic and geochemical constraints. Lithos 37: 245-260.
Kerr, A. C., Mariner, G. F., Tarney, J., Nivia, A., Saunders, A. D., Thirlwall, M. F. and Sinton, C. W.,
1997. Elemental and isotopic constraints on the petrogenesis and origin of Cretaceous basaltic
terranes in Western Colombia. Jour. Petrol. 38:677-702
Khalturin, V. I., Rautian, T. G. and Molnar, P., 1977. The spectral content of Pamir-Hindu Kush
intermediate depth earthquakes; evidence for a high-Q zone in the upper mantle, J. Geophys.
Res. , 82(20): 2.931-2.943.
Kirby, S. H., 1983. Rheology of the lithosphere. Rev. Geophys. Space Phys., 21: 1458-1487.
Kluth, Ch., F., Laad, R., De Armas, M., Gmez, L. and Tilander, N., 1997. Different structural styles
and histories of the Colombian foreland: Castilla and Chichimene oil fields areas, east-central
Colombia. VI Simpolsio Bolivariano Exploracion en las Cuencas Subandinas, Cartagena de
Indias, Mem. Tomo II: 185-197.
Kohn, B. P., Shagam, R., Banks, P. O. and Burkley, E. A., 1984. Mesozoic-Pleistocene fission track
ages on rocks of the Venezuelan Andes and their tectonic implications. In: W.E. Bonini, R.B.,

278
References
Hargraves and R. Shagam (eds.), The Caribbean-South American plate boundary and regional
tectonics. Geol. Soc. Am. Mem., 162: 365-384.
Kooi, H., 1991. Tectonic modelling of extensional basins: the role of lithosphere flexure, intraplate
stresses and relative sea-level change. Ph.D. Thesis, Vrije Universiteit, Amsterdam. 183 p.
Kooi, H., Cloetingh, S. and Burrus, J., 1992. Lithospheric necking and regional isostasy at extensional
basins: part 1, Subsidence and gravity modeling with an application to the Gulf of Lions margin
(SE France). J. Geophys. Res. , 97: 17553-17571.
Kovas, E. J., Rodgers, D. A. and Binger, J. H., 1982. Seismic interpretation of back thrusts and
displacement transfer zone between in chelon thrust faults, Middle Magdalena Valley,
Colombia. V Congreso Latino-Americano de Geologa, Argentina, Actas 1: 565-582.
Kuhry, P., Helmens, K. and Rutter, N. W., 1994. Magnitude and duration of four stadial intervals
between 22 and 8 Ka. In the tropical Andes of Colombia (South America). Geol. Soc. Am.
1994 Annual Meeting Abstracts with programs. Geol. Soc. Am. 26(7), p. 446.
Kusznir, N. J. and Karner, G., 1985. Dependence of the flexural rigidity of the continental lithosphere
on rheology and temperature. Nature, 316: 138-142.
Kusznir, N. J. and Ziegler, P. A., 1992. Mechanics of continental extension and sedimentary basin
formation: a simple-shear/pure-shear flexural cantilever model. In: P.A. Ziegler (ed.),
Geodynamics of rifting, v. III, Thematic discussions, Tectonophysics, 215: 117-131.
Kusznir, N. J., Marsden, G. and Egan, S. S., 1991. A flexural cantilever simple-shear/pure-shear model
of continental lithosphere extension: application to the Jeanne dArc Basin, Grand Banks and
Viking Graben, North Sea. In: A. M. Roberts, G., Yielding and B. Freeman (eds.), The
geometry of normal faults. Geol. Soc., London, Spec. Publ. 56: 41-60.
Lance-Le Cornec S., 1998. 1D and 2D themal modeling of sandstone reservoirs. Incidence of the
circulation of meteoric and hydrothermal water. In: Toro, J., Bordas-LeFloch, N., Le Cornec-
Lance, S., Roure, F., Bnard, F., Lafargue, E., Sassi, W., Robion, P., Aubourg, C., Frizon de La
Motte, D., Guilhaumou, N., Report on the Colombian transects Subtrap Consortium (Subthrust
reservoir appraisal), Paris, Chapter 2, 20p.
Laubscher, H. P., 1987. The kinematic puzzle of the Neogene Northern Andes. In J. P Schaer and J.
Rodgers, (eds.), The anatomy of mountain ranges, Chapter 11, Princeton University Press,
Princeton, New Jersey, p. 211-227.
Laubscher, H., 1990. The regional tectonics of the thrustbelts bordering the Llanos and the Middle and
Upper Magdalena Valley, Colombia. Report to Hocol. Geological Institute of the University
Basel, Switzerland, 35 p.
Laverde, F., 1979. Espesor estratigrafa y facies de la Formacin Guaduas en algunos sitios del
Cuadrngulo K-11. Tesis Pregrado Geologa, Univ. Nacional de Colombia, Bogot, 92 p.
Laverde F., 1985. La Formacin Los Santos: un depsito continental anterior al ingreso marino del
Cretcico. In: F. Etayo-Serna and F. Laverde-Montao (eds.), Proyecto Crtacico,
contribuciones. Chapter XX, Ingeominas Publicacin Geolgica Especial 16, Bogot, 24 p.
Laverde, F., 1989. Stratigraphy of the Tertiary sequence southwest of Bogot, Colombia, northeastern
Upper Magdalena Valley western border of the Cordillera Oriental. M.Sc. Thesis, Univ. of
South Carolina, Columbia, SC., 66 p.
Laverde F. and Clavijo, J., 1985. Anlisis facial de la Formacin Los Santos, segn el corte Yo y Tu
(Zapatoca). In: F. Etayo-Serna and F. Laverde-Montao (eds.), Proyecto Crtacico,
contribuciones. Chapter VI, Ingeominas Publicacin Geolgica Especial 16, Bogot, 9 p.
Laverde, F. and Ramrez, C., 1987. Unpublished stratigraphic column, 1:5000 scale, Area 4. In: R.
Ressetar, and C.E. Macellari). ESRI Tech. Rept. 88-0006, 40 p.
Leeds, A., 1977. Velocidades del manto en la regin Colombia-Ecuador. In: S. J. Ramrez and L.T.
Aldrich (eds.), La Trancisin ocano-continente en el suroeste de Colombia. Proyecto
Cooperativo Internacional Nario. Inst. Geof. Univ. Javeriana, Bogot, p. 237-238.
Linares, R., 1996. Structural styles and kinematics of the Medina area, Eastern Cordillera, Colombia,
M.Sc. Thesis, Univ. of Colorado, 104 p.
Lyon-Caen, H. Armijo, R. and Cifuentes, I., 1990. Recent deformation of the Ecuadorian sub-Andes
(abstract), Eos Trans, AGU, 71, p 1560.

279
References
Macia, C., 1995. El Magmatismo Trasico Superior Jursico en Colombia. Breve sntesis,
contribution to IGCP 322 Jurassic events in South America, Geologa Colombiana, Bogot, 20:
163-164.
Maca, C. and Mojica, J., 1981. Nuevos puntos de vista sobre el magmatismo Trasico Superior (Fm
Saldaa), Valle Superior del Magdalena. Colombia. Zbl. Geol. Palaeont. Stuttgart, 1(3-4): 243-
251.
Maca, C., Mojica, J. and Colmenares, F., 1985. Consideraciones sobre la importancia de la
paleogeografa y las reas de aporte precretcicas en la prospeccin de hidrocarburos en el
Valle Superior del Magdalena, Colombia. Geologa Colombiana, Bogot, 14: 49-70.
Malavielle, J., 1984. Modlisation exprimentale des chevauchements imbriques: Application aux
chanes de montagnes. Bull. Soc. Gol. France, 7, XXVI(1): 129-138.
Mandl, G., 1988. Mechanics of tectonic faulting; models and basic concepts, Developments in
Structural Geology 1. Elsevier, Amsterdam, 407 p.
Mann P. and Burke, K., 1984. Neotectonics of the Caribbean. Rev. Geophys., 22: 309-362.
Mann P., Schubert, C. and Burke, K., 1990. Review of Caribbean neotectonics. In G. Dengo and J Case
(eds.), The Caribbean Region, vol. H, The geology of North America. Geol. Soc. Am. Boulder,
p. 307-338.
Mantilla, M., 2000, Presentacin de resultados interpretacin rea Lisama. In: Modelamiento
estructural de los campos Cretceos del VMM. Unpublished report Instituto Colombiano del
Petroleo, Piedecuesta, Santander. 12 p.
Marksteiner, R. and Aleman, A. M., 1997, Petroleum systems along the fold belt associated to the
Maran Oriente Putumayo (MOP) foreland basins. VI Simposio Bolivariano Exploracin
Petrolera en las Cuencas Subandinas. P. 63-74.
Martnez, J. I., 1990. Estratigrafa de la plancha 227. Ingeominas Informe Interno, Bogot, 120 p.
Martnez W., Rubio, Mndez, A. and Taborda, A., 1998. Ideas sobre la Geologa estructural del
Noreste de Colombia y su relacin con los campos de petrleo. III Simposio Bolivariano
Exploracin Petrolera en las Cuencas Subandinas Caracas, p. 505-522.
Maughan, E. K., Zambrano, F. O., Mojica, G., Abozagalo, M., Pachn, F. P. and Duran, R. R., 1979.
Paleontologic and stratigraphic relations of phosphate beds in Upper Cretaceous rocks of the
Cordillera Oriental, Colombia. U.S. Geol. Surv. Open File Report 79-1525, 97 p.
Mayorga M. and Vargas, M., 1995. Caracterizacin geoqumica y facial de las rocas potencialmente
generadoras de hidrocarburos del Cretceo y Terciario inferior, Cordillera Oriental y
Piedemonte. Tesis Pregrado Geologa, Univ. Nacional de Colombia, Bogot, 150 p.
Maze, W. B., 1984. Jurassic la Quinta Formation in the Sierra de Perij, northwestern Venezuela:
geology and tectonic environment of red beds and volcanic rocks. In: W. E. Bonini, R. B.,
Hargraves and R. Shagam (eds.), The Caribbean-South American plate boundary and regional
tectonics. Geol. Soc. Am. Mem., 162: 263-282.
McCollough,C. N. and Carver, J. A., 1992. The giant Cao Limn field, Llanos Basin, Colombia. In:
M.T. Halbouty (Ed.), Giant oil and gas fields of the decade 1978-1988 Conference. A.A.P.G.
Mem. 54: 175-195.
McCourt, W. J., Feininger, T. and Brook, M., 1984. New geological and geochronological data from
the Colombian Andes: continental growth by multiple accretion. Jour. Geol. Soc. London, 141:
831-845.
McKenzie, D., 1978. Some remarks on the development of sedimentary basins. Earth Planet. Sci. Lett.,
40: 25-32.
McKenzie, D. and Bickle, M. J., 1988. The volume and composition of melts generated by extension of
the lithosphere. Jour. Petrol., 29: 625-679.
McLaughlin Jr, D. H., 1972. Evaporite deposits of Bogot Area, Cordillera Oriental, Colombia.
A.A.P.G. Bull., 56(11): 2240-2259.
McLaughlin, D. H. and Arce Jr., M., 1969. Mapa geolgico del cuadrngulo K-11 Zipaquir, Scale
1:100.000, Ingeominas, Bogot.
Mgard, F., 1987. Cordilleran Andes and marginal Andes: A review of Andean geology north of the
Arica elbow (18 S). In: J.W.H. Monger, J. Francheteau (Eds), Circum-Pacific orogenic belts

280
References
and evolution of the Pacific Ocean Basin. International Lithosphere Program Contribution,
Geodynamic Series, Washignton, D.C., V. 18: 71-95.
Meier, B., Schwander, M. and Laubscher, H. P., 1987. The tectonics of Tchira: a sample of North
Andean Tectonics. In J. P. Schaer and J. Rodgers, (eds.), The anatomy of mountain ranges,
Chapter 12, Princeton University Press, Princeton, New Jersey, p. 229-237.
Meissner, R. O., Fleh, E. R., Stibane, F. and Berg, E., 1973. Dynamics of the active plate boundary in
southwest Colombia according to recent geophysical measurements. In: J. E., Ramrez and L.
T. Aldrich (eds.), La transicin ocano-continente en el suroeste de Colombia, Proyecto
Cooperativo Internacional Nario, Univ. Javeriana, Bogot, p. 157-168.
Meissner, R. O., Flh, E. R., Stibane, F. and Berg, E., 1976. Dynamics of the active plate boundary in
southwest Colombia according to recent geophysical measurements. Tectonophysics 35: 115-
136.
Meja, R. M. V. and Giraldo, R. J., 1993. Estudio petrogrfico de la Formacin Picacho en un rea al
Noreste de Sogamoso, Boyac. Tesis pregrado Geologa, Univ. Nacional de Colombia, Bogot,
80 p.
Mendivelso, L. D., 1982. Aspectos fotogeolgicos y estratigrficos del Cretceo en la regn de Itaibe,
Valle Superior del Magdalena. Tesis Pregrado geologa. Univ. Nacional de Colombia, Depto.
de Geociencias, Bogot, 116 p.
Mendoza H., 1985. La Formacin Cumbre, modelo de transgresin marina rtmica, de comienzos del
Cretcico. In: F. Etayo-Serna and F. Laverde-Montao (eds.), Proyecto Crtacico,
contribuciones. Chapter IX, Ingeominas Publicacin Geolgica Especial 16, Bogot, 17 p.
Meschede, M. and Frisch, W., 1998. A plate tectonic model the Mesozoic and Early Cenozoic history
of the Caribbean Plate. Tectonophysics, 296(3-4): 269-291.
Meyer, R. P., Mooney, W. D., Hales, A. L., Helsley, C. E., Woollard, G. P., Hussong, D. M., Kroenke,
L. W. and Ramrez, J. E., 1973. Refraction observation across a leading edge, Malpelo Island to
the Colombian Cordillera Occidental. In: J.E., Ramirez and L. T.Aldrich (eds.), La transicin
ocano-continente en al suroeste de Colombia, Proyecto Cooperativo Internacional Nario,
Univ. Javeriana, Bogot, p. 83-136.
Miall, A. D., 1992. Exxon global chart; an event for every occasion?. Geol. Soc. Am. Bull., 20(9): 787-
790.
Miall, A. D., 1993. Exxon global chart; an event for every occasion?: Reply. Geol. Soc. Am. Bull.,
21(3): 284-285.
Miller, T. A. and Etayo-Serna, F., 1979. The geology of the Eastern Cordillera, between Aguazul-
Sogamoso-Villa de Leiva, 1972, In Geotec (Ed.) Geological Field Trips, Colombia, 1958-1978.
Colombian Society Petroleum Geologists and Geophysicists, Bogot: 349-396.
Mojica, J., 1982. Observaciones acerca del estado actual del conocimiento de la Formacin Payand
(Trisico superior), Valle superior del Ro Magdalena, Colombia. Geologa Colombiana,
Bogot, 11: 67-91.
Mojica, J., 1983. Direcciones de esfuerzos preteritos y recientes en Colombia. 10a Conferencia
Geolgica del Caribe, 14-22 Aug, 22 p.
Mojica, J., 1985. Estilolitos horizontales en las sedimentitas del Cretcico de la Cordillera Oriental, y
su significado tectnico. Geologa Colombiana, Bogot, 14: 7-24.
Mojica, J. and Llinas, R., 1984. Observaciones recientes sobre las caractersticas del basamento
econmico del Valle superior del Magdalena en la regin de Prado-Rovira (Tolima, Colombia)
y en especial sobre la estratigrafa y petrografa del Miembro Chical (=parte baja de la Fm.
Saldaa). Geologa Colmbiana, Bogot, 13: 81-128.
Mojica, J. and Macia, C., 1981. Caractersticas estratigrficas y edad de la Formacin Yav, Mesozoico
de la regin entre Prado y Dolores, Tolima, Colombia. Geologa Colombiana, Bogot, 12: 7-32.
Mojica, J. and Sheidegger, A. E., 1981. Diaclasas recientes en Colombia y su significado tectnico.
Geologa Colombiana, Bogot, 12: 57-90.
Mojica, J. and Scheidegger, A. E., 1984. Fsiles deformados y otras estructuras microtectnicas en la
Formacin Hil (Albiano). Alrededores de Sasaima, Cundinamarca (Colombia). Geologa
Colombiana, Bogot, 13: 41-54.

281
References
Mojica, J. and Villarroel, C., 1984. Contribucin al conocimiento de las uniddes Paleozicas del rea
de Floresta (Cordillera Oriental Colombiana, Departamento de Boyac) y en especial de la
Formacin Cuche. Geologa Colombiana, Bogota, 13: 55-78.
Mojica, J. and Franco, R., 1990. Estructura y evolucin tectnica del Valle Medio y Superior del
Magdalena, Colombia. Geologa Colombiana, 17: 41-64.
Mojica, J. and Maca, C., 1992. Geological Reconnaissence of the Northern Neiva Basin, Upper
Magdalena Valley, Colombia. XXVII Field Conf. 1980, Geol. Field Trips 1980-1989,
Colombian Society Petroleum Geologists and Geophysicists. Ediciones Geotec, Bogot, p. 91-
120.
Mojica, J. and Kammer, A., 1995. Resumen de Jursico en Colombia, Contribution IGCP 322 Jurassic
Events in South America, Geologa Colombiana, Bogot, 20: 160-162.
Mojica, J., Macia, C. and Colmenares, F., 1985. Consideraciones sobre la importancia de la
paleogeografa y las reas de aporte precretcicas en la prospeccin de hidrocarburos en el
Valle Superior del Magdalena, Colombia . Geologa Colombiana, Bogot, 14: 49-70.
Mojica J., Kammer A. and Ujueta, G., 1996. El Jursico del sector noroccidental de Suramerica y gua
de la excursin al Valle Superior del Magdalena (Nov. 1-4/95), Regiones de Payand y Prado,
Departamento del Tolima, Colombia. Geologa Colombiana, Bogot, 21:
Mooney, W. D., Meyer, R. P., Laurence, J. P., Meyer, H. and Ramirez, J. E., 1979. Seismic refraction
studies of the Western Cordillera, Colombia. Bull. Seism. Soc. Am. 69: 1745-1761.
Mora, A. and Kammer, A., 1999. Comparacin de los estilos estructurales en la seccin entre Bogot y
los Farallones de Medina, Cordillera Oriental de Colombia. Geol.- Colombiana,Bogot, 24: 55-
82.
Mora, C., 1996. Evaluacin geoqumica. In: Torres M.P., Fajardo, G., Mora, C.A., Bedregal, R.P.,
Estudio de evaluacin conjunta estratigrfica, geoqumica y estructural Bloques Laguna
(Occidental), Lanceros (Braspetro) y Tunja (Ecopetrol). Cordillera Oriental.Santaf de Bogot,
100 p.
Mora, C., 1997. Pulsos de generacin y expulsin de hidrocarburos a partir de rocas de edad Cretcica
en el Valle Medio del Magdalena (Colombia). VI Simposio Bolivariano Exploracin Petrolera
en las Cuencas Subandinas, Cartagena, Abs., 22.
Mora, C., 2000. Evaluacin del potencial de los sistemas petrolferos en las Cuencas Cretcicas con
produccin comercial en Colombia (Putumayo, Valle Superior del Magdalena, Valle Medio del
lMagdalena, Catatumbo y Llanos Orientales). Maestra en Sistemas Petrolferos Univ. Federal
de Ro de Janeiro and Ceenpes/Petrobras (Coppe and Cegeq). Ro de Janeiro, 177 p., 4
appendixes.
Mora, C. and Posada, C., 2000. Elementos y procesos de los sistemas petrolferos en las cuencas
Cretceas Colombianas con produccin comercial (Putumayo, Valle Superior del Magdalena,
Valle Medio del Magdalena, Catatumbo y Llanos Orientales). In prep.
Mora, C., Cordoba, F., Luna O., Sarmiento L. F., Bartels H., Rangel A., Giraldo B. N., Reyes J. P. and
Magoon L., 1996. Definicin de los Sistemas Petrolferos del Valle Medio del Magdalena.
Empresa Colombiana de Petrleos Vicepresidencia de Exploracin y Produccin, Gerencia de
Exploracin. Instituto Colombiano del Petrleo. Internal Report, Bogot, 100 p.
Mora, C., Torres, M. P. and Escobar, J., 1997. Potencial generador de hidrocarburos de la Formacin
Chipaque y su relacin estratigrfica secuencial en la zona axial de la Cordillera Oriental
(Colombia). VI Simposio Bolivariano Exploracin Petrolera en las Cuencas Subandinas,
Cartagena, Abs., 23, Tomo I: 217-237.
Morales, L. G., and the Colombian Petroleum Industry., 1956. General geology and oil occurrences of
the Middle Magdalena Valley, Colombia. In: L. G. Weeks (ed.), Habitat of the Middle and
Upper Magdalena basins, Colombia. Oil - a symposium, A.A.P.G. p. 641-695.
Moreno, J. M., 1990a. Stratigraphy of the Lower Cretaceous Rosablanca Formation, west flank,
Eastern Cordillera, Colombia. Geologa Colombiana, Bogot, 17: 65-86.
Moreno, J. M., 1990b. Stratigraphy of the Lower Cretaceous units central part Eastern Cordillera,
Colombia, 13th International Sedimentological Congress, Nothingham, Aug 22-31, 1990.
Abstract.

282
References
Moreno, J. M., 1991. Provenance of the Lower Cretaceous sedimentary sequences, central part, Eastern
Cordillera, Colombia. Revista Academia Colombiana Ciencias Exactas Fsicas y Naturales,
18(69): 159-173.
Moreno, J. M. and Concha, A.E., 1993. Nuevas manifestaciones gneas bsicas en el flanco occidental
de la Cordillera Oriental, Colombia. Geologa Colombiana, 18: 143-150.
Moreno, M. and Rubiano, J., 1987, Unpublished stratigraphic column, 1:5000 scale, Area 2. In: R.
Ressetar, and C.E. Macellari). ESRI Tech. Rept. 88-0006, 40 p.
Moreno J. and Velzquez, M., 1993. Estratigrafa y tectnica en los alrededores del municipio de
Nuncha, Departamento del Casanare, Colombia. Tesis Pregrado Geologa Univ. Nacional de
Colombia, Bogot, 60 p.
Morley, C. K., Nelson, R. A., Patton, T. L. and Munn, S. G., 1990. Transfer zones in the East African
rift system and their relevance to hydrocarbon exploration in rifts. A.A.P.G. Bull., 74: 1234-
1253.
Mueller, S. and Phillips, R. J., 1991. On the initiation of subduction. J. Geophys. Res. , 96: 651-666.
Naar, J. and Coral, M. G., 1993. Modelo tectnico de un rea del Piedemonte Llanero al noroeste de
Yopal-Casanare (Colombia). Tesis Pregrado Univ. Nacional de Colombia, Bogot, 33p.
Namson, J., Cunningham, R. and Wodcock, G., 1994. Structural geology and hydrocarbon potential of
the northern part of the Upper Magdalena Basin, Colombia. V Simposio Bolivariano
Exploracin Petrolera en las Cuencas Subandinas, Puerto La Cruz, Memoria: 356-364.
Nelson, R. A., Patton, T. L. and Morley, C. K., 1992. Rift-segment interaction and its relation to
hydrocarbon exploration in continental rift systems. A.A.P.G. Bull., 76: 1153-1160.
Nivia, A., 1987. The geochemistry and origin of the Amime and volcanic sequences, SW Colombia.
M. Phil. Thesis, Univ. of Leicester, UK, 164 p.
Notestein, F. B., Hubman, C. W. and Bowler, J. W., 1944. Geology of the Barco Concession, Republic
of Colombia. Geol. Soc. Am. Bull., 55: 73-1218.
Nez, A., 1986. Petrognesis del Batolito de Ibagu. Geologa Colombiana, Bogot, 15: 35-45.
Nuttal, C. P., 1990. A review of the Tertiary non-marine molluscan faunas of the Pebasian and other
inland basins of northwestern South America. Bull. Bri. Mus. Nat. Hist. Geol. 45(2): 165-371.
Occidental, Braspetro and Ecopetrol., 1996. Geochemistry and sequence strtigraphy of the source rocks
in the Colombia eastern Cordillera. Field trip. Asociacin Colombiana Gelogos y Geofsicos
del Petrleo, Santaf de Bogot, 25 p.
Occidental, Ecopetrol and Braspetro., 1996. Estudio de evaluacin conjunta estratigrfica, geoqumica
y estructural bloques Laguna (Occidental), Lanceros (Braspetro) y Tunja (Ecopetrol),
Cordillera Oriental. Bogot, 100 p.
Ocola, L., Aldrich, L. T., Gettrust, J. F., Meyer, R. P. and Ramrez, J. E., 1973. Project Nario I:
Crustal structure under southern Colombian-northern Ecuador Andes from seismic refraction
data. In: J. E., Ramirez and L. T.Aldrich (eds.), La transicin ocano-continente en al suroeste
de Colombia, Proyecto Cooperativo Internacional Nario I, Univ. Javeriana, Bogot, p. 47-70.
Ocola, L. C., Aldrich, L. T., Getrust, T. F., Meyer, R. P. and Ramrez, J. E., 1975. Proyecto Nario I:
Crustal structure under southern Colombia-northern Ecuador Andes from seismic refraction
data. Bull. Seism. Soc. Am., 65: 1.685-1.695.
Ochoa, L. E. and Ponguta, S. Y., 1991. Determinacin de esfuerzos tectnicos actuales mediante
diversos mtodos, especialmente alongacin de pozos (breakouts) en el Valle Superior del
Magdalena, Colombia, Mem. IV Simposio Exploracin Petrolera en las Cuencas Subandinas,
Bogot, Trabajo 13.
Ojeda, G. and Pea H., 1994. Interpretacin de la geologa del subsuelo en el rea Ortega Chaparral
Coyaima (Departamento del Tolima). In: F. Etayo Serna (ed.), Estudios geolgicos del Valle
Superior del Magdalena. Chapter VIII, Univ. Nacional de Colombia, Ecopetrol, Bogot, 20 p.
Okaya, N., Cloetingh, S. and Mueller, St., 1996. A lithosphere cross-section through the Swiss Alps
II. Constraints on the mechanical structure of a continent-continent collision zone. Geoph. Jour.
Int. 127, 399-414.
Olaya, I. D.and Serrano M.P., 1998, Anlisis ssmico estratigrfico del rea de Zarzalito en la Cuenca
VMM. Informe final. Instituto Colombiano del Petrleo Divisin de Exploracin y Produccin,

283
References
Area de Estratigrafa, Ecopetrol Vicepresidencia de Adjunta dce Exploracin, Gerencia de
Prospeccin de Hidrocarburos, Piedecuesta, 52 p., 6 appendixes. Compact disc.
Ord, A. and Hobbs, B. E., 1989. The strength of the continental crust, detachment zones, and the
development of plastic instabilities. Tectonophysics, 158: 269-289.
Page, W. D. and James, M. E., 1981. The antiquity of the erosion surfaces and late Cenozoic deposits
near Medelln, Colombia: Implications to tectonics and erosion rates. Mem. 1 Sem. Cuat. Col.,
Revista CIAF, 6: 421-454.
Pardo Casas, F. and Molnar, P., 1987. Relative motion of the Nazca (Farallon) and South American
plates since Late Cretaceous time. Tectonics, 6: 233-248.
Parker, R. L., 1972. The rapid calculation of potential anomalies. Geophys. Jour. Roy. Astr. Soc., 31:
447-455.
Parkinson, N. and Summerhayes, C., 1985. Synchronous global sequence boundaries. A.A.P.G. Bull.,
69, 685-687.
Peel, F. J. and Hossack, J. R., 1987. A structural study of the T.E.A. And adjacent areas in the Eastern
Cordillera, Colombia. Ecopetrol ISN 9737, Informe 2219 (In 11297), 35 p.
Pennington, W. D., 1981. Subduction of the eastern Panam Basin and seismo-tectonics of
northwestern South America, J. Geophys. Res. , 86: 10753-10770.
Pennington, W. D., Mooney, W. D., Van Hissenhoven, R., Meyer, H. J., Ramrez, J. E. and Meyer, R.
P., 1979. Results of a reconnaissance microearthquake survey of Bucaramanga, Colombia.
Geoph. Res. Lett, 6: 65-68.
Prez, O.J., Jaimes, M.A. and Garciacaro, E., 1997, Microseismicity evidence for subduction of the
Caribbean plate beneath the South American plate in northwestern Venezuela. J. Geophys. Res.
, 102(B8): 17.875-17882.
Perez, G. and Salazar, A., 1978. Estratigrafa y facies del Grupo Guadalupe. Geologa Colombiana.
Bogot, 10: 7-85.
Pilger, R. H., Jr., 1984. Cenozoic plate kinematics subduction and magmatism: South American Andes.
Jour. Geol. Soc., London, 141: 793-802.
Pimpirev C. T., Patarroyo, P. and Sarmiento, G., 1992. Stratigraphy and facies analysis of the Caqueza
Group, a sequence of Lower Cretaceous turbidites in the Cordillera Oriental of the Colombian
Andes. Geologa Colombiana, Bogot, 17: 297-308.
Pindell J. L., 1993. Regional synopsys of Gulf of Mexico and Caribbean evolution, In Pindell, J. L.,
and Perkins, B. F., (eds.), GCSSEPM Foundation 13th Annual Research Conference
Proceedings, Mesozoic and early Cenozoic development of the Gulf of Mexico and Caribbean
region. SEPM Gulf Coast Section, p. 251-274.
Pindell, J. and Dewey, J., 1982. Permo-Triassic reconstruction of western Pangaea and the evolution of
the Gulf of Mexico-Caribbean region. Tectonics 1: 179-211.
Pindell, J. L. and Barret, S. F., 1990. Geological evolution of the Caribbean region: A plate tectonic
perspective. In: G. Dengo and J Case (eds.), The Caribbean region, vol. H, The geology of
North America. Geol. Soc. Am. Boulder, p. 405-432.
Pindell, J. and Erikson, J., 1993. The Mesozoic margin of northern South America. In: J. Salfity (ed.),
Cretaceous tectonics of the Andes, Vieweg Germany, p. 1-60.
Pindell J. L. and Tabbutt K. D., 1995. Mesozoic-Cenozoic Andean paleogeography and regional
controls on hydrocarbon systems. In A. J., Tankard, R. Suarez and H. J. Welsink (eds.),
Petreoleum basins of South America: A.A.P.G. Mem., 62: 101-128.
Pindell, J., George, Jr., R. P., Cristancho, J. and Higgs, R., 1997. Clarification of the Late Cretaceous-
Paleogene evolution of Colombia. IV Simposio Bolivariano Exploracin Petrolera en las
Cuencas Subandinas, Bogot, Tomo I: 129-132.
Polana, J. H. and Rodrguez, O. G., 1978. Posibles turbiditas del Cretceo Inferior (Miembro Socot)
en el rea de Anapoima (Cundinamarca); una investigacin sedimentolgica basada en registros
grficos. Geologa Colombiana, Bogot, 10: 87-91.
Porta, J., 1965. La estratigrafa del Cretcico Superior y Terciario en el extremo S del Valle Medio del
Magdalena. Univ. Industrial de Santander, Bucaramanga, Boletn Geolgico (19): 5-30.
Porta, J., 1966. La geologa del extremo sur del Valle Medio del Magdalena. Boletn Geolgico Univ.
Industrial de Santander, Bucaramanga, 22-23: 1-347.

284
References
Posamentier, H, W., Allen, G. P., James, D. P. and Tesson, M., 1992. Forced regressions in a sequence
stratigraphic framework: concepts, examples, and exploration significance. A.A.P.G. Bull., 76:
1687-1709.
Price, R. A., 1973. Large-scale gravitational flow of supracrustal rocks, southern Canadian Rockies. In:
K. de Jong and R. Scholten (eds.), Gravity and tectonics, Wiley, New York. Proceedings of the
7th International conference on basement tectonics, p. 491-502.
Puerto, E. Y., 1982. Gradientes geotermicos en las cuencas de los Valles Inferior, Medio y Sueprior del
Magdalena y de los Llanos Orientales de Colombia. Tesis Ingeniera de Petrleos Fundascin
Universidad de America, Bogot, 154 p.
Pulham, A. J., 1994. The Cusiana field, Llanos Basin, eastern Colombia; high resolution sequence
stratigraphy applied to late Paleocene-early Oligocene estuarine, coastal plain and alluvial
clastic reservoirs. In: S. D. Johnson (ed.), High resolution sequence stratigraphy; innovation
and applications. Univ. of Liverpool, Abstract volume, p. 63-68.
Pulham, A. J., 1995. The sedimentology and stratigraphy of the early Tertiary Cusiana field, Colombia,
Congreso Colombiano del Petrleo, Bogot, p. 107-113
Pulido, O., 1979a. Geologa de las planchas 135 Sa Gil y 151 Charal. Departamento de Santander,
Ingeominas, Boletn Geolgico Ingeominas, Bogot, 23(2): 42-78.
Pulido, O., 1979b. Mapa geolgico preliminar de la plancha 135, San Gil. Scale 1:100.000,
Ingeominas, Bogot.
Pulido, O., 1980. Mapa geolgico preliminar de la plancha 151, Charal. Scale 1:100.000. Ingeominas,
Bogot.
Raasvelt, H. C., 1956. Mapa geolgico de la Plancha L-9 Girardot, Map scale 1:200.000, Servicio
Geolgico Nacional, Bogot.
Raasvelt, H. C., et al., 1957. Mapa geolgico de la Plancha M-8 Ataco, Map scale 1:200.000, Servicio
Geolgico Nacional, Bogot.
Ramirez, J. E. and Aldrich, L. T., (eds.), 1973. La transicin ocano-continente en al suroeste de
Colombia, Proyecto Cooperativo Internacional Nario, Univ. Javeriana, Bogot, 313 p.
Ramrez, N. and Ramrez, H., 1994. Estratigrafa y origen de los carbonatos del Cretcico Superior en
el Valle Superior del Magdalena, Departamento del Huila, Colombia. In: F. Etayo Serna (ed.),
Estudios geolgicos del Valle Superior del Magdalena. Chapter V, Univ. Nacional de
Colombia, Ecopetrol, Bogot, 15 p.
Ranalli, G., 1987. Rheology of the Earth. Allen and Unwin, 366 p. Boston (2nd edition: 1994).
Ranalli, G., 1995. Rheology of the Earth. Chapman & Hall, London, 413 p.
Ranalli, G. and Murphy, D.C., 1987. Rheological stratification of the lihtosphere. Tectonophysics, 132:
281-295.
Rathke, W. W. and Coral, M. G., 1997. Cupiagua field, Colombia: Interpretation case history of a
large, complex thrust belt gas condensate field. VI Simposio Bolivariano Expoloracin
Petrolera en las Cuencas Subandinas, Cartagena, Abs., 9.Tomo I: 119-128.
Renzoni, G., 1965a. Apuntes acerca de la litologa y tectnica de la zona al este y sureste de Bogot.
Boletn Geolgico, Servicio Geolgico Nacional Bogot, X(1-3):59-79.
Renzoni, G., 1965b. Geologa del cuadrngulo L-11 Villavicencio, mapa a escala 1:200.000. Servicio
Geolgico Nacional e Inventario Minero Bogot.
Renzoni, G., 1967. Geologa del Cuadrngulo J-12, Tunja. Boletn Geolgico, Ingeominas, Bogot,
24(2): 32-48.
Renzoni, G., 1968. Geologa del Macizo de Quetame. Geologa Colombiana, Bogot, 5: 75-127.
Renzoni, G., 1985a. Paleoambientes de la Formacin Tambor en la Quebrada Pujamanes. In: F. Etayo-
Serna, F. Laverde-Montano (eds.), Proyecto Crtacico, contribuciones. Chapter III, Ingeominas
Publicacin Geolgica Especial 16, Bogot, 18 p.
Renzoni, G.,1985b. La secuencia facial de la Formacin Los Santos por la Quebrada Piedra Azul:
Registro de una hoya fluvial evanescente. In: F. Etayo-Serna and F. Laverde-Montao (eds.),
Proyecto Cretcico, contribuciones. Chapter IV, Ingeominas Publicacin Geolgica Especial
16, Bogot, 18 p.

285
References
Renzoni, G., 1985c. Paleoambientes de la Formacin Arcabuco y Cumbre de la Cordillera de Los
Cobardes. In: F. Etayo-Serna and F. Laverde-Montao (eds.), Proyecto Crtacico,
contribuciones. Chapter X, Ingeominas Publicacin Geolgica Especial 16, Bogot, 14 p.
Renzoni, G., 1992. Mapa Geolgico de la Plancha 193 de Yopal: Ingeominas, Esc 1:100.000, Bogot.
Renzoni, G., Rosas, H. and Etayo, F., 1967. Mapa geolgico, plancha 191, Tunja, Map scale 1:100.000.
Ingeominas, Bogot.
Ressetar, R. and Schamel S., 1986. The geology and hydrocarbon potential of the Upper and Middle
Magdalena Valleys, Colombia. Ecopetrol, Informe Geol. 1545, Bogot.
Restrepo, J.J., Toussaint, J. F., Gonzalez, H., Gordani, U., Kawashita, K., Linares, E. and Parica, C.,
1991. Precisiones geocronolgicas sobre el occidente Colombiano. Simposio sobre el
magmatismo Andino y su marco tectnico. Programa Internacional de Correlacin Geolgica
UNESCO, Union Internacional de Ciencias Geolgicas. Manizales, Colombia, Julio 3-5 de
1991. Manizales, p. 1-21.
Restrepo-Pace, P. A., 1989. Restauracin de la seccin geeolgica Cqueza Puente Quetame:
Moderna interpretacin estructural de la deformacin del flanco este de la Cordillera Oriental.
Tesis Geologa (Laureada), Univ. Nal. de Colombia, Santaf de Bogot, 100 p.
Restrepo-Pace, P. A., 1995. Late Precambrian to Early Mesozoic tectoic evolution of the Colombian
Andes, based on new geochronological, geochemical and isotopic data. Ph.D. Thesis. Univ.
Arizona, 195 p.
Restrepo-Pace, P. and Villamil, T., 1998. Assesment of structural styles and estimation of stratigraphic
thickness, stretching and crustal shortening in the Eastern Andean Cordillera, Colombia. VI
Simposio Exploracin Petrolera Cuencas Subandinas, Cartagena, Tomo II: 90-94.
Restrepo-Pace, P. A., Colmenares, F., Higuera, C. and Mayorga, M., 1999a. Fold and thrust belt along
the western flank of the Eastern Cordillera of Colombia: style, kinematics and timing contraints
derived from seismic data and detailed surface mapping. In: K. McClay (ed.), A.A.P.G. Spec.
Publ. In press., 25 p.
Restrepo-Pace, P. A., Colmenares, F., Higuera, C., Mayorga, M. and Leal, J., 1999b. Fold and thrust
belt along the western flank of the Eastern Cordillera of Colombia: style, kinematics and timing
contraints derived from seismic data and detailed surface mapping. A.A.P.G. Bull., 82(10), p.
1956.
Reyes, J. P., 1994. Estratigrafa de secuencias y potenciales trampas estratigrficas en el rea del Meta -
Llanos Orientales de Colombia. V Simposio Bolivariano Exploracin Petrolera en las Cuencas
Subandinas, Puerto La Cruz, Memoria: 122-146.
Reyes, I. and Reyes V. M., 1976. Geologa del yacimiento y variabilidad de las caractersticas
geoqumicas del mineral de hierro en la regin de Paz Vieja (Municipio de Paz de Ro,
Departamento de Boyac). Primer Congreso Colombiano de Geologa, Memoria, p. 267-324.
Rivadeneira, M. V., 1996. Cretaceous-Paleogene stratigraphic sequences and the early Andean
orogenic events in the Ecuadorian Oriente Basin. Third ISAG. St-Malo, France, 17-19, Sept./
1996, Extended abstracts, p. 469-472.
Robertson Research Inc., 1983. The northern Llanos of Colombia, hydrocarbon potential and
stratigraphical control. Propietary report. Houston, 250 p.
Robertson Research Inc., 1986. Llanos Basin, Colombia, a sedimentololgical and stratigraphical
correlation study. 290 p. Vol. 1, Text; Vols. 2-4, Appendices, Vols. 5-8, Enclosures.
Robertson Research Inc., 1989. Tectonic and stratigraphical study of the Middle and Upper Magdalena
Valley, Colombia, 4 Vols.
Rodrguez, C. and Rojas, R., 1985. Estratigrafa y tectnica de la serie InfraCretcica en los alrededores
de San Flix, Cordillera Central de Colombia. In: F. Etayo-Serna and F. Laverde-Montao
(eds.), Proyecto Cretcico, contribuciones. Chapter XXI, Ingeominas Publicacin Geolgica
Especial 16, Bogot, 21 p.
Rodrguez, E. and Ulloa, C., 1994a. Mapa Geolgico de la Plancha 169 de Puerto Boyac: Ingeominas,
Esc 1:100.000, Bogot.
Rodrguez, E. and Ulloa, C., 1994b. Mapa Geolgico de la Plancha 189 de La Palma: Ingeominas, Esc
1:100.000 Bogot.

286
References
Roeder, D. and Chamberlain, R. L., 1995. Eastern Cordillera of Colombia: Jurassic-Neogene crustal
evolution. In: A. J. Tankard, R. Surez S., and Welsnik H. J. (eds.), Petroleum basins of South
America: A.A.P.G. Mem., 62: 633-645.
Rohrman, M., 1995. Thermal evolution of the Fennoscandian region from fission track
thermochronology; an integrated approach. Ph.D. Thesis, Vrije Universiteit, Amsterdam, 168 p.
Roln, L. F. and Carrero, M. M., 1995. Anlisis estratigrfico de la seccin Cretcica aflorante al
oriente del anticlinal de Los Cobardes entre los Municipios de Guadalupe-Chima-Contratacin,
Departamento de Santander. Tesis pregrado Geologa, Univ. Nacional de Colombia, Bogot, 80
p.
Romero, F. H. and Rincn, M. A., 1990. Caractersticas petrogrficas y geoqumicas de las rocas
volcnicas de Iza (Departamento de Boyac, Colombia). Geologa Colombiana, Bogot, 17:
159-168.
Romero, F. H., Schultz, R. A. and Kawashita, K., 2000. Geoqumica del rubidio-estroncio y edad de las
esmeraldas Colombianas. Geologa Colombiana, Santafe de Bogot, 25: 221-239.
Rosendahl, B. R., 1987. Architecture of continental rifts with special reference to East Africa. Ann.
Rev. Earth Planet. Sci. 15: 445-503.
Ross M. I. and Scotese C. R., 1988. A hierarchical tectonic model of the Gulf of Mexico and Caribbean
region. Tectonophysics, 155: 139-168.
Rouby, D., 1995. Restauration en carte des domaines faills en extension: Mthode et applications.
Mem. Geosci. Rennes, Rennes, France, 58, 230 p.
Rouby, D., Fossen, H. and Cobbold, P. R., 1996. Extension, displacement, and block rotation in the
Larger Gullfaks Area, northern North Sea, determined from map view restoration. A.A.P.G.
Bull., 80: 875-890.
Rouby, D., Xiao, H. and Suppe, J., 1998. 3D-restoration of complexly folded and faulted surfaces using
multiple unfolding mechanisms. Unpublished Report, 20 p.
Roure, F., 1998. Subduction of the lihtosphere and crustal balance in orogenic belts. In: S. Cloetingh,
G. Ranalli and C.A. Ricci (eds.), Sedimentary basins models and constraints. Proceedings of
the International School Earth and Planetary Sciences, Siena, p. 175-192.
Roure, F., Choukroune, P., Berastegui, X., Muoz, J. A., Matheron, P., Bareyt, M., Sguret, M.,
Camara, P. and Deramond, J., 1989. ECORS deep seismic data and balanced cross-sections;
geometric constraints on the evolution of the Pyrenees, Tectonics, 8(1): 41-50.
Roure, F., Colletta, B., De Toni, B., Loureiro, D., Passalacqua, H. and Gou, Y., 1997. Within-plate
deformations in the Maracaibo and East Zulia basins, western Venezuela. Marine and
Petroleum Geology, 14(2): 139-163.
Royden, L. H., 1993a. The tectonic expression of slab pull at continental convergent boundaries.
Tectonics, 12: 303-325.
Royden, L. H., 1993b. Evolution of retreating subduction boundaries formed during continental
collision. Tectonics, 12: 629-683.
Royden, L. and Keen, C. E., 1980. Rifting process and thermal evolution of the continental margin of
eastern Canada determined from subsidence curves. Earth Planet. Sci. Lett. 51, 343-361.
Royden, L. and Karner, G. D., 1984. Flexure of the continental lithosphere beneath Apennine and
Carpathian foredeep basins: evidence for an insufficient topographic load. A.A.P.G. Bull., 68:
704-712.
Rubiano, J. L., 1989. Petrography and stratigraphy of the Villeta Group, Codillera Oriental, Colombia,
South America. M.Sc. Thesis, Univ. South Carolina, Columbia, SC., 96 p.
Rubiano, J. L.,1998. Dipmeter paleocurrent determination in the Provincia field, Middle Magdalena
Valley. Ciencia Tecnologa y Futuro, Piedecuesta, Santander, 1(4): 33-48.
Sacks, P. E. and Secor, D. T., 1990. Delamination in collisional orogens, Geology 18: 999-1002.
Salazar, A., 1987. Unpublished stratigraphic column, 1:5000 scale, Area 5. In: R. Ressetar, and C.E.
Macellari). ESRI Tech. Rept. 88-0006, 40 p.
Salazar, A., 1992. Depositional and paleotectonic settings of the Cretaceous sequence, Upper
Magdalena Valley, Colombia. PhD. Dissertation, University of South Carolina, Columbia S.C.,
200 p.

287
References
Salvador, M., 1991. Gravity field, topography and crustal structure of the North Andes. M.Sc. Thesis
Univ. of South Carolina, Columbia SC., 100 p.
Santo, T., 1969. Characteristics of seismicity in South America, Bull. Earthquake Res. Inst., Univ.
Tokyo, 47: 635-672.
Sarmiento, G., 1992. Estratigrafa y medios de depsito de la Formacin Guaduas. Ingeominas Bogot,
Boletn Geolgico, 32(1): 3-44.
Sarmiento, G., 1993. Estratigrafa, palinologa y paleoecologa de la Formacin Guaduas
(Maastrichtiano Paleoceno Colombia), Academisch Proefrschrift. Ph.D. Thesis Univ. of
Amsterdasm, Amsterdam, 192 p.
Sarmiento L. F., 1989. Stratigraphy of the Cordillera Oriental west of Bogot, Colombia. M.Sc. Thesis
University of South Carolina, Columbia, SC.,102 p.
Sarmiento, L. F., Cardozo, E., Forero, H. and Ramrez, C., 1985. Importancia de la estratigrafa en la
evaluacin de las anomalas geoqumicas: caso del rea del ro Zumbe, Utica (Cundinamarca).
In: F. Etayo-Serna and F. Laverde-Montao (Eds.) Proyecto Cretceo, contribuciones. Publ.
Geol. Esp. Ingeominas, Chapter XXVIII Aplicacin econmica de los estudios estratigrficos
Part C, Bogot, 10 p.
Sarmiento, L. F., Van Wess, J.D. and Cloetingh, S., in prep. Mesozoic and Cenozoic rheological
evolution of the lithosphere of the Eastern Cordillera, Colombian Andes. Inferences from
tectonic models., 24 p.
Sassi, W., Rudkewickz, J. L. and Divies, R., 1998. New methods for integrated modelling of
deformation and petroleum generation in fold and thrust belts. A.A.P.G. Annual Meeting, Salt
Lake City. Abstracts.
Sawyer, D. S., 1985, Brittle failure in the upper mantle during extension of continental lithosphere. J.
Geophys. Res. , 90 (B4), 3021-3025.
Sawyer, D. S. and Harry, D. L., 1991. Dynamical modeling of divergent margin formation; application
to the U. S. Atlantic margin. Mar. Geol. 102: 29-42.
Sclater, J. G. and Christie, P. A. F., 1980. Continental stretching; an explanation of the post mid-
Cretaceous subsidence of the central North Sea Basin. J. Geophys. Res. , 85: 3711-3739.
Schamel, S., 1991. Middle and Upper Magdalena basins, Colombia. In: K.T. Biddle (ed.), Active
margins. A.A.P.G. Mem., 52: 283-301.
Schelling D., 1994. Stuctural Geology. In: Earth Sciences and Resources Institute (ESRI), 1997,
Technical Report 86-0014, Univ. of South Carolina, Columbia, SC. v. 4, 100 p.
Schneider, J. F., Pennington, W. D. and Meyer, R. P., 1987. Microseismicity and focal mechanisms of
the intermediate-depth Bucaramanga nest, Colombia. J. Geophys. Res. , 92: 13.915-13.926.
Schubert, C., 1982. Neotectonics of the Bocon Fault, western Venezuela. Tectonophysics, 52: 205-
220.
Schwabe, E., Toro, G., Kairuz, Ch. and Ferreira, P., 2001. Edades por trazas de fisin de circones
provenientes de la Formacin Saldaa, Valle Superior del Magdalena. Resumen. Coloquio
Trisico-Jursico en Colombia. Ingeominas, Univ. Indust. Santander, Sociedad Colombiana de
Geologa. Bucaramanga, Febrero 22-23 de 2001. Bucaramanga, 1 p.
Servicio Geolgico Nacional, (Compilator from Ecopetrol and other Petroleum Companies)., 1967a.
Geologa del cuadrngulo H-11, Barrancabermeja, Map scale 1:200.000. Servicio Geolgico
Nacional, Bogot.
Servicio Geolgico Nacional, (Compilator from Ecopetrol and other Petroleum Companies)., 1967b.
Geologa del cuadrngulo I-11, Cimitarra, Map scale 1:200.000. Servicio Geolgico Nacional,
Bogot.
Setoguchi, T. and Rosenberg, A. L., 1985. Some new ceboid primates from the La Venta, Miocene of
Colombia. Mem., VI Congreso Latinoamericano de Geologa, Bogot, v. 1: 287-298.
Setoguchi, T. and Rosenberg, A. L., 1987. A fossil owl monkey from La Venta, Colombia. Nature,
326: 692-694.
Shagam, R., 1975. The Northern termination of the Andes. In: A.E. M., Nairn, and H. Stehli (eds.), The
ocean basins and margins, v. 3, The Gulf of Mexico and the Caribbean. Plenum Press, New
York and London, Chapter 9, p. 325-420.

288
References
Shagam, R., Kohn, B. P., Banks, P. O. Dash, L. E., Vargas, R., Rodriguez, G. I. and Pimentel, N.,
1984. Tectonic implications of Cretaceous-Pliocene fission track ages from rocks of the
circum-Maracaibo Basin region, western Venezuela and Eastern Colombia. In: W. E. Bonini,
R. B. Hargraves and R. Shagam (eds.), The Caribbean South American plate boundary and
regional tectonics, Geol. Soc. Am., Mem. 162, p. 385-412.
Shemenda, A. I., 1993. Subduction of the lithosphere and back-arc dynamics: Insigths from physical
modeling. J. Geophys. Res. , 98: 16.167-16.185.
Sibson, R. H., 1995. Selective fault reactivation during basin inversion: potential for fluid redistribution
through fault-valve action. In: J. G. Buchanan and P. G. Buchanan (eds.), Basin inversion.
Geol. Soc. London, Spec. Publ., 85: 3-19.
Sinton, C. W., Duncan, R. A., Storey, M., Lewis, J. and Estrada, J. J., 1998. An oceanic flood basalt
province within the Caribbean plate. Earth and Planet. Sci. Lett. 155: 221-235.
Sleep, N. H., 1971. Thermal effects of the formation of Atlantic continental margins by continental
break-up. Geophys. J. Roy. Astr. Soc., 24: 325-350.
Smith, G. A. and Landis, C. A., 1995. Intra-arc Basins. In: C. J. Busby and R. V. Ingersoll (eds.),
Tectonics of sedimentary basins, Blackwell Science, Cambridge, Massachusetts, Chapter 7, p.
263-298.
Steckler, M. S. and Watts, A. B., 1978. Subsidence of the Atlantic-type continental margin off New
York. Earth Planet. Sci. Lett., 41:1-13.
Steinmann, M., Seward, D. and Hungerbuehler, D., 1996. Thermotectonic history of the Andes, South
Ecuador: evidence from fission-track dating. Third ISAG. St-Malo, France, 17-19 Sept./ 1996,
Extended abstracts, p. 501-503.
Stephenson, R. A. and Cloetingh, S., 1991. Some examples and mechanical aspects of continental
lihtospheric folding. Tectonophysics, 188: 27-37.
Steuer, M. R., Bentham, P. A. and Latimer, G. H., 1997. Structural style and emplacement of the Opn
anticline, Middle Magdalena Valley, Colombia. VI Simposio Bolivariano Exploracin
Petrolera en las Cuencas Subandinas. Cartagena, Mem. Tomo II: 257-262.
Streckeisen, A., 1976. To each plutonic rock its proper name. Earth Sci. Rev., 12(1): 1-33.
Surez, G., Molnar, P. and Burchfiel, B. C., 1983. Seismicity, fault planes solutions, depth of faulting,
and active tectonics of the Andes of Per, Ecuador, and southern Colombia, J. Geophys. Res. ,
88(B12): 10.403-10428.
Surez M., 1982. Estratigrafa y Facies de la Formacin Arenisca de El Limbo en el Borde Oeste de
la Cuenca de los Llanos Orientales. Tesis Pregrado. Universidad Nacional de Colombia.
Bogot. Pag. 74.
Surez, M., 1997. Tectonoestratigrafa del Terciario y potencial petrolfero de la Formacin La Paz,
Cuenca del Valle Medio del Magdalena (parte norte), Colombia. VI Simposio Exploracin
Petrolera en las Cuencas Subandinas, Tomo II: 95-98.
Suppe, J.,1983. Geometry and kinematics of fault-bend folding. Am. Jour. Sci. 238: 684-721.
Suppe, J. and Medwedeff, D. A., 1990, Geometry and kinematics of fault-propagation folding. In: P.
Jordan, T., Noack, S., Schmid and D. Bernoulli (eds.), The Hans Laubsher volume. Eclog.
Geol. Helv., 83(3): 409-454.
Taboada, A., Rivera, L.A., Fuenzalida, A., Cisternas, A., Philip, H., Bijwaard, H., Olaya, J. and Rivera,
C., 2000, Geodynamics of the northern Andes: Subductions and intracontinental deformation
(Colombia). Tectonics, 19 (5): 787-813.
Taboada, A., Rivera, L. A., Fuenzalida, A., Cisternas, A., Philip, H., Castro, J. E. and Rivera, C., 1999.
Geodynamics of the northern Andes: Intra-continental subduction and the Bucaramanga
seismicity nest (Colombia). Fouth ISAG, Goettingen, Germany, 4-6 Oct/1999, p. 719-724.
Takemura, K. and Danhara, T., 1983. Fission-track age of pumices included in the Gigante Formation,
north of Neiva, Colombia. Overseas Research Reports of New World Monkeys, Kioto Univ,
Japan, v. 3: 13-15.
Tamaki, K. and Honza, E., 1991. Global tectonics and the formation of marginal basins: Role of the
western Pacific. Episodes, 14: 224-230.

289
References
Tellez, N. and Navas, J., 1962. Interferencia de las direcciones de los pliegues cretcico-terciarios entre
Coello y Gualanday (Valle Superrior del Magdalena). Boletn Geolgico Univ. Industrial de
Santander, Bucaramanga, 9: 45-61.
Ter Voorde, M., 1996. Tectonic modelling of lithospheric extension along faults, implications for
thermal and mechanical structure and basin stratigraphy. Ph.D. Thesis, Vrije Universiteit,
Amsterdam, 197 p.
Testamarck, J. S., Perdomo, J. L. and Serrano, I., 1994. Estilos estructurales del flanco suroriental de la
Sierra de Perij, Venezuela. V Simposio Exploracin Petrolera en las Cuencas Subandinas,
Puerto La Cruz, Memoria: 314-316.
Thouret, J. C., 1988. La Cordillre Centrale des Andes de Colombie et ses bordures: Morphogense
Plio-Quaternaire et dynamique actuelle et rcente dune cordillre volcanique englace. Ph.D.
Thesis, Univ. Joseph Fourier, Grenoble, France.
Thouret, J. C., Murcia, A., Salinas, R. and Vatin Prignon, N., 1985. Cronoestratigrafa mediante
dataciones K-Ar y 14C de los volcanes compuestos del complejo Ruz-Tolima y aspectos
volcano-estructurales del Nevado del Ruz (Cordillera Central, Colombia). VI Congreso
Latinoamericano de Geologa, Bogot, 1: 385-454.
Toro, G., Popeau, G., Hermelin, M. and Schwabe, E., 1999. Chronology of the volcanic activity and
regional thermal events: A contribution from the tephrochronology in the north of the Central
Cordillera, Colombia. Fourth ISAG. Geottingen, Germany, 4-6 Oct./ 1999, Extended Abstracts,
p. 761-763.
Toro, J., 1990. The termination of the Bucaramanga Fault in the Cordillera Oriental, Colombia. M.Sc.
Thesis Univ. of Arizona, Arizona, 60 p.
Toro, J., 1992. Interpretacin estructural del rea Gualanday, Valle Superior del Magdalena. Ecopetrol
Vicepresidencia de Exploracin y Produccin, Gerencia de Exploracin. Proyecto Gualanday
No 20, Ecopetrol, Informe Geolgico, Santaf de Bogot, 40 p.
Toro, J., 1998. Geodynamic evolution of the Eastern Cordillera. In: Toro, J., Bordas-LeFloch, N., Le
Cornec-Lance, S., Roure, F., Bnard, F., Lafargue, E., Sassi, W., Robion, P., Aubourg, C.,
Frizon de La Motte, D., Guilhaumou, N., Report on the Colmbian transects Subtrap Consortium
(Subthrust reservoir appraisal), Paris, Chapter 1, 12p. .
Torres, C., 1992. In situ. stress in the North Andes from borehole breakout analysis. M.Sc. Thesis
University of South Carolina, Columbia S.C., 153 p.
Toussaint, J. F., 1978. Grandes rasgos geolgicos de la parte septentrional del noroccidente
Colombiano. Boletn Ciencias de la Tierra. Univ. Nal. Medelln, 1: 1-147.
Toussaint, J. F., 1993. Evolucin geolgica de Colombia, 1, Precmbrico, Paleozico. Univ. Nacional
de Colombia, Medelln, 229 p.
Toussaint, J. F., 1995a. Hiptesis sobre el marco geodinmico de Colombia durante el Mesozico
temprano, Contribution to IGCP 322 Jurassic events in South America, Geologa Colombiana,
Bogot, 20: 150-155.
Toussaint, J. F., 1995b. Evolucin geolgica de Colombia 2. Trisico Jursico. Contribucin al IGCP
322 Correlation of Jurassic events in South America International Geological Correlation
Programme Unesco IUGS. Univ. Nacional de Colombia. Medelln, 94 p.
Toussaint, J. F. and Restrepo, J. J., 1974a. Algunas consideraciones sobre la evolucin estructural de
los Andes Colombianos. Informe parcial. Publicacin Geolgica Especial Geol., Univ.
Nacional de Colombia, Medelln, 4: 1-17.
Toussaint, J. F. and Restrepo, J. J., 1974b. La Formacin Abejorral y sus implicaciones sobre la
evolucin de la Cordillera Central de Colombia durante el Cretceo. Anales de la Facultad de
Minas, Medelln, 58: 13-29.
Toussaint, J. F. and Restrepo, J. J., 1989. Acresiones sucesivas en Colombia; un nuevo modelo de
evolucin geolgica. V Congreso Colombiano de Geologa, Bucaramanga, I: 127-146.
Toussaint, J. F. and Restrepo, J. J., 1991. El magmatismo en el marco de la evolucin geotectnica de
Colombia. Simposio sobre el magmatismo Andino y su marco tectnico. Programa
Internacional de Correlacin Geolgica, UNESCO. Unin Internacional de Ciencias
Geolgicas. Manizales, Colombia, Julio 3-5 de 1991. Manizales, p. 135-150.

290
References
Toussaint, J. F. and Restrepo, J. J., 1994. The Colombian Andes during Cretaceous times. In: J.A.
Salfity (ed.), Cretaceous tectonics of the Andes. Views Brunwick, Germany, p. 61-100.
Tron, V. and Brun, J. P., 1991. Experiments on oblique rifting in brittle-ductile systems.
Tectonophysics, 188: 71-84.
Trygvason, E. and Lawson, J. E. Jr., 1970. The intermediate earthquake source near Bucaramanga,
Colombia. Bull. Seism. Soc. Am., 60: 269-276.
Turcotte, D. L. and Schubert, G., 1982. Geodynamics, applications of continuum physics to geological
problems, John Wiley New York., 450 p.
Ujueta, G., 1961. Geologa del Noreste de Bogot. Boletn Geolgico Servicio Geolgico Nacional
Bogot, 9(1-3): 23-46.
Ujueta, G., 1993. Lineamientos Muzo, Tunja y Paipa en los Departamentos de Boyac y Casanare,
Colombia. Geologa Colombiana, 18: 65-73.
Ulloa, C. and Rodrguez, E., 1976a. Geologa del cuadrngulo K-12, Guateque. Boletn Geolgico,
Ingeominas, Bogot, 22(1): 4-55.
Ulloa, C. Rodrguez, E., et al., 1976b. Geological map Cuadrngulo K-13. Tauramena . Cuenca de los
Farallones. Scale 1:100.000, Ingeominas, Bogot.
Ulloa, C. and Roddrguez, E., 1979a. Geologa de las planchas 170 velez y 190 Chiquinquir,
Colombia. Ingeominas Informe 1794, Sogamoso, 45 p.
Ulloa, C. and Rodrguez, E., 1979b. Geologa del Cuadrngulo K-13, Tauramena. Boletin Geolgico,
Ingeominas, Bogot, 24(2): 3-30.
Ulloa, C. and Rodrguez, E., 1987. Mapa geolgico de la plancha 190, chiquinquir, Map Scale
1:100.000, ingeominas, Bogot.
Ulloa, C. and Rodrguez, E., 1991. Mapa geolgico de Colombia, plancha 190-Chiquinquira
(Departamento de Boyac), Ingeominas, Santafe de Bogot, Scale 1:100.000, 26 p.
Ulloa, C., Rodrguez, E. and Guerra, A., 1973. Mapa geolgico de la plancha 172 Paz de Ro. Scale
1:100.000, Ingeominas Bogot.
Ulloa, C. Camacho, R. and Escovar, R., 1975. Mapa Geolgico de la Plancha K-12 de Guateque:
Ingeominas, Esc 1:100.000, Bogot.
Ulloa, C.; Escobar, R. and Pacheco, A., 1976c. Mapa Geolgico de la Plancha 230 de Monterrey:
Ingeominas, Esc 1:100.000, Bogot.
United States Geological Survey, National Earthquake Information Center. World Data Center for
Seismology, (http://www.neic.cr.usgs.gov/neis/epic/epic.html).
Ureta, J.E. and Du Toit, Ch., 1997. Prospeccin de hidrocarburos en el piedemonte de la Cuenca del
Putumayo. VI Simposio Exploracin petrolera en las Cuencas Subandinas. Cartagena de Indias,
Mem. Tomo I: 263-273.
Uyeda, S. and McCabe, R., 1983. A possible mechanism of episodic spreading of the Philippine Sea.
In: M. Hashimoto and S. Uyeda, (eds.), Accretion tectonics in the Circum-Pacific regions.
Terra, Tokyo, p. 291-306.
Van Andel, T. H., 1958. Origin and classification of Cretaceous, Paleocene and Eocene sandstones of
western Venezuela. A.A.P.G. Bull., 42(4): 734-763.
Van der Beek, P., 1995. Tectonic evolution of continental rifts, inferences from numerical modelling
and fission track thermochronology. Ph.D. Thesis, Vrije Universitteit, Amsterdam, 232 p.
Van der Beek, P. A. and Cloetingh, S., 1992. Lithospheric flexure and the tectonic evolution of the
Betic Cordilleras (SE Spain). Tectonophysics, 203: 325-344.
Van der Hammen, T., 1958. Estratigrafa del Terciario y Maestrichtiano continentales y tectognesis de
los Andes Colombianos. Boletin Geolgico, Ingeominas, Bogot, 6(1-3): 67-128.
Van der Hamen T. H., 1961. Late Cretaceous and Tertiary stratigraphy and tectogenesis of the
Colombian Andes. Geologie en Mijnbouw, 40: 181-188.
Van der Hammen T. and Hooghiemstra H., 1997. Chronostratigraphy and correlation of the Pliocene
and Quaternary of Colombia, Quaternary International, 40: 81-91.
Van der Hammen, T., Werner, J. H. and Van Dommelen, H., 1973. Palynological record of the
upheaval of the northern Andes: a study of Pliocene and lower Quaternary of the Colombian
Eastern Cordillera and the early evolution of its high-Andean biota. Paleobotany and
Palynology, 16: 1-122.

291
References
Van der Wiel, A. M., 1991. Uplift and volcanism of the SE Colombian Andes in relation to Neogene
sedimentation of the Upper Magdlena Valley. Ph.D. Thesis, Wageningen, 208 p.
Van der Wiel, A. M. and Andriessen, P. A. M., 1991. Precambrian to recent thermotectonic history of
the Garzn Massif (Eastern Cordillera of the Colombian Andes) as revealed by fission track
analysis. In: A.M. Van der Wiel., Uplift and volcanism of the SE Colombian Andes in relation
to Neogene sedimentation of the Upper Magdalena Valley. Ph.D. Thesis, Wageningen, 208 p.
Van der Wiel, A. M. and Van den Bergh., 1992a. Uplift, subsidence, and volcanism in the southern
Neiva Basin, Colombia, Part 1: Influence on fluvial deposition in the Miocene Honda
Formation. Jour. South Am. Earth Sciences, 5(2): 153-175.
Van der Wiel, A. M. and Van den Bergh., 1992b. Uplift, subsidence, and volcanism in the southern
Neiva Basin, Colombia, Part 2: Influence on fluvial deposition in the Miocene Gigante
Formation. Jour. South Am. Earth Sciences, 5(2): 175-196.
Van Houten, F. B., 1976. Late Cenozoic volcaniclastic deposits, Andean foredeep, Colombia. Geol.
Soc. Am. Bull., 87(4): 481-495.
Van Houten, F. B. and Travis, R. B., 1968. Cenozoic deposits, Upper Magdalena Valley, Colombia.
A.A.P.G. Bull., 52: 675-702.
Van Wees, J. D., 1993. Flex3D, version 1.01, Dec. 1993. User manual. Netherlands Research School of
Sedimentary Geology, Faculty of Earth Sciences, Vrije Universiteit, Amsterdam, 13 p.
Van Wees, J. D., 1994. Tectonic modelling of basin deformation and inversion dynamics. The role of
pre-existing faults and continental lithosphere rheology in basin evolution. PhD thesis Vrije
Universiteit, Amsterdam, 164 p.
Van Wees, J. D., 2000. Basin deformation partitioning and the role of pre-existing weak zones.
Geophysical Research Abstracts, 25th General Assembly European Geophysical Society.
Abstract in Volume 2, Compact Disc.
Van Wees, J. D. and Cloetingh, S., 1993. Role of pre-existing weak zones during basin extension and
inversion. Terra Abstracts. J. Suppl. 1, p. 211.
Van Wees, J. D. and Cloetingh, S., 1994. A finite difference technique to incorporate spatial variations
in rigidity and planar faults into 3-D models for lithospheric flexure. Geoph. Jour. Int. 117(1):
179-195.
Van Wees, J. D. and Stephenson, R. A., 1995. Quantitative modelling of basin and rheological
evolution of the Iberian Basin (central Spain); implications for lithospheric dynamics of
intraplate extension and inversion. In: S. Cloetingh, D. Argenio, B., Catalano, R., Horvath, F.,
Sassi, W. (eds.), Interplay of extension and compression in basin formation. Tectonophysics,
252 (1-4): 163-178.
Van Wees, J. D. and Beekman, F., 2000. Lithosphere rheology during intraplate basin extension and
inversion Inferences from automated modelling of four basins in western Europe.
Tectonophysics, 320: 219-242.
Van Wees, J. D., De Jong, K. and Cloetingh, S., 1992. Two dimensional modeling and the dynamics of
extension and inversion in the Betic zone (SE Spain). Tectonophysics, 203: 305-324.
Van Wees, J. D., Cloetingh, S. and De-Vicente, G., 1996a. The role of pre-existing faults in basin
evolution; constraints from 2-D finite element and 3-D flexural models. In: P. G. Buchanan and
D. A. Nieuwland, (eds.), Modern developments in structural interpretation valildation and
modelling. Geol. Soc. Spec. Publ. London, 99: 297-320.
Van Wees, J. D., Stephenson, R. A., Stovba, S. M. and Shymanovskyi, V. A., 1996b. Tectonic
variation in the Dnieper-Donets Basin from automated modelling of backstripped subsidence
curves. Tectonophysics, 268(1-4): 257-280.
Van Wees, J. D., Arche, A., Beijdorff, C. G., Lpez-Gmez, J. and Cloetinhg, S., 1998. Temporal and
spatial variations in tectonic subsidence in the Iberian Basin (eastern Spain): inferences from
automated forward modelling of high resolution stratigraphy (Permian-Mesozoic).
Tectonophysics, 300:285-310.
Vanegas, D. and Arango, F., 1994. Anlisis estratigrfico de una regin al Occidente de la poblacin de
Paz de Ariporo (Casanare). Tesis Pregrado, Univ. Nacional de Colombia, Bogot, 70 p.
Vargas, R. and Arias, A., 1981a. Mapa geolgico de la plancha 86, Abrego, Scale 1:100.000,
Ingeominas, Bogot.

292
References
Vargas, R. and Arias, A., 1981b. Mapa geolgico de la plancha 97, Cchira, Scale 1:100.000,
Ingeominas, Bogot.
Vargas, R., Arias, L., Jaramillo, L. and Tellez, N., 1984a. Geologa del cuadrngulo I-13, Malaga,
Boletn Geolgico Ingeominas, Bogot, 24(3): 1.76.
Vargas, R., Arias, L., Jaramillo, L. and Tellez, N., 1984b. Mapa geolgico preliminar de la plancha
136, Malaga, Scale 1:100.000, Ingeominas, Bogot.
Vargas, R., Etayo, Renzoni, G. and Tllez, N., 1985. Corte estratigrfico panormico de la Formacin
Los Santos, Carretera Quebrada El Medio El Boquern Santander. In: F. Etayo-Serna and F.
Laverde-Montao (eds.), Proyecto Crtacico, contribuciones. Chapter V, Ingeominas
Publicacin Geolgica Especial 16, Bogot, 12 p.
Vargas, R.; Arias, A.; Jaramillo, L and Tellez, N., 1987. Mapa Geolgico de la Plancha 152 de Soata:
Ingeominas, Esc 1:100.000.
Vsquez, C. A., 1983. Geologa del Paleoceno Superior en la margen occidental de la Cuenca de los
Llanos Orientales. Tesis Pregrado Geologa Univ. Nacional de Colombia, Bogot, 91 p.
Vsquez, C. A., 1988. Mapa estructural del tope del Cretceo, Cuenca Llanos Orientales de Colombia.
III Simposio Bolivariano Exploracin Petrolera en las Cuencas Subandinas, Caracas, 1: 580-
609.
Vsquez, L. E. and Ros, O., 1979. Geologa y geoqumica de sedimentos activos de la regin situada al
norte de dolores (Tolima). Tesis pregrado Univ. Nacional de Colombia, Depto. Geociencias,
Bogot, 80 p.
Vergara, L. and Prssl , K. F., 1994. Dating the Yav Formation (Aptian Upper Magdalena Valley,
Colombia). In: F. Etayo Serna (ed.), Estudios geolgicos del Valle Superior del Magdalena.
Chapter XVIII, Univ. Nacional de Colombia, Ecopetrol, Bogot, 14 p.
Villamil, T.,1993. Relative sea level, chronology, and a new sequence stratigraphy model for distal
offshore facies, Albian to Santonian, Colombia. In Pindell, J. A. and Drake, C. D. (eds),
Mesozoic-Cenozoic stratigraphy and tectonic evolution of the Caribbean Region/Northern
South America: Implications for eustasy from exposed sections of a Cretaceous-Eocene passive
margin setting. Geol. Soc. Am. Memoir, paper C-8.
Villamil, T., 1994. High-resolultion stratigraphy, chronology and relaitve sea level of the Albiian-
Santonian (Cretaceous) of Colombia. Ph.D. Thesis Univ. Of Colorado at Boulder, 446 p.
Villamil, T., 1999. Campanian-Miocene tectonostratigraphy, depocenter evolution and basin
development of Colombia and western Venezuela. Palaeogeography, Palaeoclimatology,
Palaeoecology, 153(1-4): 239-275.
Villamil, T. and Restrepo, P., 1997. Paleocene-Miocene paleogeographic evolution of Colombia.
Memorias VI Simposio Bolivariano Exploracin Petrolera en las Cuencas Subandinas.,
Cartagena, Sept.14-17, 1997. Tomo I: 275-302.
Villamil, T. and Arango, C., 1998. Integrated stratigraphy of latest Cenomanian and early Turonian
facies of Colombia. Paleogeographic evolution and non-glacial eustasy, Northern South
America, SEPM Spec. Publ. 58: 129-159.
Villamil, T. and Pindell, J. L., 1998. Mesozoic paleogeographic evolution of northern South America:
Foundations for sequence stratigraphy studies in passive margin strata deposited during non-
glacial times. Paleogeographic evolution and non-glacial eustasy, Northern South America,
SEMP. Spec. Publ. 58: 283-318.
Villarroel, C. and Guerrero, J., 1984. Un nuevo y singular representante de la familia leontiniidase?
(notoungulata, Mammalia) en el Mioceno de La Venta, Colombia. Geologa Norandina,
Bogot, 9: 35-40.
Waddell, M., 1982. Depositional model for the Monserrate Formation of the Neiva Basin, Colombia,
South America. MS. Thesis, Univ. South Carolina, Columbia, SC. 83 p.
Walcott, R. I., 1970. Flexural rigidity, thickness and viscosity of the lithosphere. J. Geophys. Res. , 75:
3941-3955.
Walthall, B. H. and Berry, D. W., 1979. Paipa-Belencito-Paz de Ro area. In Geotec (Ed.), Geological
Field Trips, Colombia. Colombian Society of Petroleum Geologists and Geophysicists.,
Ediciones Geotec, Bogot,: 65-87.

293
References
Ward, D., Golsmith, R., Cruz, J. and Restrepo, H., 1973. Geologa de los cuadrngulos H-12,
Bucaramanga y H-13, Pamplona, Depto. de Santander. Boletn Geolgico Ingeominas, Bogot,
21 (1-3): 1-132.
Watts, A. B., Karner, G. D. and Steckler, M. S., 1982. Lithospheric flexure and the evolution of
sedimentary basins. Phil. Trans. Roy. Soc. London, 305: 249-281.
Wellman, S. S., 1967. Stratigraphy, petrology and sedimentology of the nonmarine Honda Formation
(Miocene), Upper Magdalena Valley, Colombia. Ph.D. Thesis, Princeton Univ., Princeton, 250
p.
Wellman, S. S., 1970. Stratigraphy and petrology of the nonmarine Honda Group (Miocene), Upper
Magdalena Valley, Colombia. Geol. Soc. Am. Bull., 81:2358-2374.
Wiggins, R. A., 1972. The general linear inverse problem: Implication of surface waves and free
occillations for earth structure. J. Geophys. Res. , 80: 1053-1064.
Wijninga, V. M. and Kuhry P., 1993. Late Piocene paleoecology of the Guasca Valley (Cordillera
Oriental, Colombia). Rev. Paleobotany and Palynology 78(1-2): 69-127.
Wilson, M., 1993. Magmatism and the geodynamics of basin formation. Sediment. Geol., 86: 5-29.
Wilson, M. and Guiraud, R., 1992. Magmatism and rifting in western and central Africa, from Late
Jurassic to recent time. In: P. Ziegler (ed.), Geodynamics of rifting Volume II Case history
studies, North, South America and Africa. Tectonophysics, 213(1-2): 203-225.
Willingshofer, E., 2000. Extension in collisional orogenic belts: the Late Cretaceous evolution of the
Alps and Carpathians. PhD thesis Vrije Universiteit, Amsterdam, 146 p.
Woerner, G., Gerdes, A., Henk, A., Tanner, D., and Finger, F., 1999. Magmatism of thickened
continental crust; examples from the Variscan Orogen (Southern Bohemian Massif) and Central
Andes. In: P. Dietrich, W. Frank and B. J. Merkel, 89th annual meeting of the Geologische
Vereinigung: Old crust, new problems, geeodynamics and utilization. Abstracts and
programme. Orogenic processes, quantification and modeling in the Variscan Belt. Terra
Nostra, Bonn, 99(1): 212-213.
Wortel, R. and Cloetingh, S., 1981. On the origin of the Cocos-Nazca spreading center. Geology, 9:
425-430.
Wortel, M. J. R., and Spakman, W., 1992, Structure and dynamics of subducted lithosphere in the
Mediterranean region. Proc. Royal Netherlands Academy of Sciences, 95: 325-347.
Yurewicz, D. A., Advocate, D. M., Lo, H. B. and Hernandez, E. A., 1998. Source rocks and oil
families, southwest Maracaibo Basin (Catatumbo sub-basin), Colombia. A.A.P.G. Bull, 82(7):
1329-1352.
Zamarreo de Julivert, I., 1963. Estudio petrogrfico de las calizas de la Formacin Rosablanca en la
regin de la Mesa de Los Santos (Cordillera Oriental, Colombia). Boletn Geolgico Univ.
Industrial de Santander, Bucaramanga, 15: 5-34.
Ziegler, P. A., 1988. Evolution of the Artic-North Atlantic and Western Tethys. A.A.P.G., Mem., 43,
198p.
Ziegler, P. A., 1990. Geological atlas of Western and Central Europe. 2nd ed. Shell Internationale
Petroleum Mij. B.V., Geol. Soc., London, 239 p.
Ziegler, P. A.,1993. Plate-moving mechanisms: their relative importance. Jour. Geol. Soc., London
150: 927-940.
Ziegler, P. A.,1994. Geodynamic processes governing development of rifted basins. In: Roure, F.,
Ellouz, N., Shein, V. S. and Skvortsov, I. (eds.), Geodynamic evolution of sedimentary basins.
Institut Franais du Ptrole, Technip, Paris, p. 16-97.
Ziegler, P. A., Cloetingh, S. and Van Wees, J. D., 1995. Dynamics of intra-plate compressional
deformation: The Alpine foreland and other examples. Tectonophysics, 252: 7-59.
Ziegler, P. A., Van Wees, J. D. and Cloetingh, S., 1998. Mechanical controls on collision-related
compressional intraplate deformation. Tectonophysics, 300: 103-129.
Zoetemeijer, R., 1993. Tectonic modelling of foreland basins. Thin skinned thrusting, syntectonic
sedimentation and lithosphere flexure. Ph.D. Thesis, Vrije Universiteit, Amsterdam, 148 p.
Zoetemeijer, R., 1998. Flexure and gravity modelling of foreland basins. User manual Cobra version 5.
Netherlands Research School of Sedimentary Geology, Faculty of Earth Sciences, Vrije
Universiteit, Amsterdam, 17 p.

294
References
Zoetemeijer, R., Cloetingh, S., Sassi, W. and Roure, F., 1993. Modelling of piggyback-basin
stratigraphy, record of tectonic evolution. Tectonophysics, 226(1-4): 253-269.
Zoetemeijer, R., Tomek, . and Cloetingh, S., 1999. Flexural expression of European continental
lithosphere under the western outer Carpathians. Tectonics, 18 (5): 843-861.

295

You might also like