maximiliano j. bezada a thesis submitted doctor of philosophy · adeyemi arogunmati, jeniffer masy...

144
RICE UNIVERSITY Crustal and Upper Mantle Investigations of the Caribbean - South American Plate Boundary by Maximiliano J. Bezada A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE Doctor of Philosophy APPROVED, THESIS COMMITTEE: / V\ <_> Lbu-^ J JL Alan Levander, Chair, Carey Croneis Professor, Earth Science Colin Zelt, Professor, Earth Science. 'A^ u. Jullg Morgan, Professor, Earth Science c r. UJJ~ Thomas Killian, Associate Professor, Physics and Astronomy HOUSTON, TEXAS JUNE 2010

Upload: others

Post on 28-Jul-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

RICE UNIVERSITY

Crustal and Upper Mantle Investigations of the Caribbean - South American Plate Boundary

by

Maximiliano J. Bezada

A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE

Doctor of Philosophy

APPROVED, THESIS COMMITTEE:

/ V\ <_> L b u - ^ J JL

Alan Levander, Chair, Carey Croneis Professor, Earth Science

Colin Zelt, Professor, Earth Science.

'A^ u. Jullg Morgan, Professor, Earth Science

c r. UJJ~ Thomas Killian, Associate Professor, Physics and Astronomy

HOUSTON, TEXAS

JUNE 2010

Page 2: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

UMI Number: 3425205

All rights reserved

INFORMATION TO ALL USERS The quality of this reproduction is dependent upon the quality of the copy submitted.

In the unlikely event that the author did not send a complete manuscript and there are missing pages, these will be noted. Also, if material had to be removed,

a note will indicate the deletion.

UMT Dissertation Publishing

UMI 3425205 Copyright 2010 by ProQuest LLC.

All rights reserved. This edition of the work is protected against unauthorized copying under Title 17, United States Code.

uest ProQuest LLC

789 East Eisenhower Parkway P.O. Box 1346

Ann Arbor, Ml 48106-1346

Page 3: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

i

ABSTRACT

Crustal and Upper Mantle Investigations of the

Caribbean - South American Plate Boundary

by

Maximiliano J. Bezada

The evolution of the Caribbean - South America plate boundary has been a matter of

vigorous debate for decades and many questions remain unresolved. In this work, and in

the framework of the BOLIVAR project, we shed light on some aspects of the present

state and the tectonic history of the margin by using different types of geophysical data

sets and techniques. An analysis of controlled-source traveltime data collected along a

boundary-normal profile at ~65°W was used to build a 2D P-wave velocity model. The

model shows that the Caribbean Large Igenous Province is present offshore eastern

Venezuela and confirms the uniformity of the velocity structure along the Leeward

Antilles volcanic belt. In contrast with neighboring profiles, at this longitude we see no

change in velocity structure or crustal thickness across the San Sebastian - El Pilar fault

system. A 2D gravity modeling methodology that uses seismically derived initial density

models was developed as part of this research. The application of this new method to four

of the BOLIVAR boundary-normal profiles suggests that the uppermost mantle is denser

under the South American continental crust and the island arc terranes than under the

Caribbean oceanic crust. Crustal rocks of the island arc and extended island arc terranes

of the Leeward Antilles have a relatively low density, given their P-wave velocity. This

Page 4: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

ii

may be caused by low iron content, relative to average magmatic arc rocks. Finally, an

analysis of teleseismic traveltimes with frequency-dependent kernels produced a 3D P-

wave velocity perturbation model. The model shows the structure of the mantle

lithosphere under the study area and clearly images the subduction of the Atlantic slab

and associated partial removal of the lower lithosphere under northern South America.

We also image the subduction of a section of the Caribbean plate under South America

with an east-southeast direction. Both the Atlantic and Caribbean subducting slabs

penetrate the mantle transition zone, affecting the topography of the 410-km and 660-km

discontinuities.

Page 5: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

i i i

ACKNOWLEDGEMENTS

I would like to thank my parents who always supported me, nurtured me and provided all

the tools I ever needed to advance my education. My father deserves special mention for

encouraging me to pursue a doctoral degree from about the time I learned to read. I am

indebted to FUNVISIS and its people as it was there and with them that I first got

involved with this project. I am especially grateful to Dr. Michael Schmitz who always

pushed me to work hard at FUNVISIS and continued to support and challenge me during

my post-graduate studies. I thank my fellow students: Stoney Clark, Amanda Beardsley,

Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may

not have survived my first year at Rice. This research would not have been possible

without my advisors: Dr. Alan Levander and Dr. Colin Zelt, from whom I have learnt a

great deal about how to think, work and write like a scientist. I thank Dr. Steve Danbom

for trusting me with his course on two occasions, and for being a friend. I am also

grateful to other faculty, students and post-docs at Rice with whom I had many positive

interactions over the years, and to the administrative staff for managing all the things that

we so often take for granted, and doing so with a warm smile and friendly disposition.

Finally, I would like to thank Rachel, my companion and partner in life, for her support

and her unwavering confidence in my scientific ability; and my son, the star that lights up

my life and gives me the most powerful reason to be the best man I can be. This work and

the broader BOLIVAR and GEODINOS projects were made possible by funding from

the NSF Continental Dynamics program (grant EAR0003572 and EAR0607801) as well

as FONACIT (grant G-2002000478) and PDVSA-INTEVEP-FUNVIS (project 04-141).

Page 6: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

iv

TABLE OF CONTENTS

ABSTRACT i

ACKNOWLEDGEMENTS iii

LIST OF FIGURES v

Introduction 1

Chapter 1: The Caribbean - South American plate boundary at 65°W: Results from wide-angle seismic data 6 1.1. Introduction 7 1.2. Tectonic Setting 10 1.3. Wide-angle Seismic Data Acquisition 14 1.4. Wide-angle Seismic Data 15 1.5. Wide-angle Traveltime Analysis 17 1.6. Results 30 1.7. Discussion 34 1.8. Summary of conclusions 45

Chapter 2: Gravity inversion using seismically derived crustal density models and genetic algorithms: An application to the Caribbean - south American plate boundary 47 2.1. Introduction 48 2.2. Description of the algorithm 52 2.3. Synthetic test 56 2.4. Application to real data 69 2.5. Discussion and conclusions 79

Chapter 3: Subduction in the Southern Caribbean: Images from finite-frequency P-wave tomography 83 3.1. Introduction 84 3.2. Data 88 3.3. Method 89 3.4. Model 96 3.5. Model Assessment 101 3.6. Discussion 108 3.7. Conclusions 120

Appendix A: Details of GA implementation 122

REFERENCES 127

Page 7: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

V

LIST OF FIGURES

Introduction Figure 1: Tectonic setting of the study area 2

Chapter 1: Figure 1.1: Location and gravity map of study area 8 Figure 1.2: Example record sections 16 Figure 1.3: First arrival tomography models 19 Figure 1.4: First arrival tomography ray coverage 20 Figure 1.5: Starting model for layer-based inversion 22 Figure 1.6: Final velocity model 23 Figure 1.7: Ray coverage for layer-based inversion 24 Figure 1.8: Resolution test results 25 Figure 1.9: Gravity modeling results 28 Figure 1.10: Comparison of Shallow velocity model and MCS line 32 Figure 1.11: One-dimensional velocity profiles 37 Figure 1.12: Schematic interpretation 43

Chapter 2: Figure 2.1: Wp-p conversion curves and empirical data 50 Figure 2.2: Known model for synthetic tests 57 Figure 2.3: Noise-free synthetic test results 59 Figure 2.4: Assessment of noise-free synthetic test results 62 Figure 2.5: Noise added to synthetic data 64 Figure 2.6: Noisy synthetic test results 66 Figure 2.7: Assessment of noisy synthetic test results 69 Figure 2.8: Gravity map of study area 71 Figure 2.9: Seismically derived initial density models 72 Figure 2.10: Real data application results 73 Figure 2.11: Interpretation of real data application results 74 Figure 2.12: Island arc Vp-pcurves and empirical data 77 Figure 2.13: Absolute density models 79

Chapter 3: Figure 3.1: Tectonic setting, station distribution and seismicity 85 Figure 3.2: Event locations 90 Figure 3.3: Assessment of free parameters 93 Figure 3.4: Effect of different free parameters 94 Figure 3.5: Velocity model depth slices (depth < 245 km) 98 Figure 3.6: Velocity model depth slices, cross sections and 3D view 100 Figure 3.7: Atlantic slab lateral resolution synthetic test results 103 Figure 3.8: Caribbean slab lateral resolution synthetic test results 105 Figure 3.9: Vertical smearing synthetic test results 106 Figure 3.10: Shallow southward dipping slab synthetic test results 107 Figure 3.11: Cartoon interpretation of removal of South American lithosphere.... 112 Figure 3.12: Caribbean slab and Bucaramanga nest 115 Figure 3.13: Caribbean slab cross section with elevation profile and seismicity... 116 Figure 3.14: Subducted slabs and transition zone topography 119

Page 8: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

Introduction

1

The margin between the Caribbean and South American plates has a history of

oblique collision that has resulted in a variety of geologic features observed on both sides

of the margin. These include accretionary wedges, fold and thrust belts, and extensional

as well as foreland basins [e.g. Ostos et al, 2005]. Additionally, the Caribbean plate

underwent extensive volcanism in the Cretaceous that resulted in the formation of the

Caribbean Large Igneous Province and a thickening of large areas of the Caribbean crust

[e.g. Hoernle et al, 2004]. Current relative plate motions make the boundary

predominantly strike-slip, bounded on the east by the Atlantic subduction and in the west

by partial subduction of the Caribbean plate under South America (Figure 1).

For the past six years, the BOLIVAR (Broadband Offshore-Onshore Lithospheric

Investigation of the Venezuela and Antilles Arc Region [Levander et al, 2006]) project

in the United States and its Venezuelan counterpart GEODINOS (Geodinamica Reciente

del Limite Norte de la Placa Suramericana) have been studying this margin in a

multidisciplinary way. The goals of the projects are to understand the current state and

the past evolution of the margin and to evaluate island arc accretion as a mechanism for

continental crust growth. The collection, modeling and interpretation of geophysical data

on a lithospheric and sublithospheric level constitute one of the main pillars of the

projects. This dissertation is composed of slightly modified versions of three manuscripts

in different stages of the publication process that describe the geophysical

characterization of the plate boundary at different levels and utilizing different types of

data and techniques.

Page 9: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

Figure 1: Tectonic setting of the study area. Red arrows show GPS vectors from a compilation by Calais and P. Mann [2009], plate boundaries from the UTIG plates database. SA - San Andres island; LA - Leeward Antilles; SS-EP - San Sebastian-El Pilar fault system

The first chapter describes the analysis of the controlled-source, wide-angle

seismic data along BOLIVAR boundary-normal profile 65W and interprets the results in

the context of velocity models derived for the other BOLIVAR profiles. At the northern

end of profile 65W we find that the Venezuela Basin is underlain by Caribbean Large

Igneous Province crust, unlike the normal oceanic crust found on profile 67W ~ 250 km

to the west. The profile crosses the trend of the Leeward Antilles islands at the La

Blanquilla High which has a velocity structure that is consistent with that found in the

other BOLIVAR profiles for the Leeward Antilles, Aves Ridge and Lesser Antilles. In

contrast with BOLIVAR profiles 67W and 64W, where the San Sebastian - El Pilar

strike-slip fault system was found to separate two distinct types of crust; along profile

Page 10: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

3

65W we see no significant differences in crustal velocity or thickness across the fault

zone. This suggests that the crustal blocks that meet at the margin have similar

characteristics and a fragment of Caribbean island arc crust may have been accreted to

South America at this location. Our findings show that the evolution of the margin is not

a simple progression of events occurring in the same manner and diachronously from

west to east, but that the geologic processes resulting from the oblique collision were

likely significantly affected by the paleogeography and pre-existing structures of the

South American side of the margin.

In chapter two I present a new methodology for 2D gravity modeling using

seismically derived initial density models and apply the algorithm to four of the

BOLIVAR boundary-normal profiles. Crustal density adjustments to the initial models in

the crystalline crust found to fit the gravity data are of a different nature on either side of

the strike slip margin for profiles 67W and 64W (negative for the island arc crust and

weakly positive for continental South American crust). On profile 65W negative density

adjustments were also found for the island arc crust, but there is no change across the

strike slip margin and the step to positive values that we associate with continental South

American crust occurs further inland. This analysis of the gravity data supports the

conclusion drawn from the seismic velocity models regarding the nature of the crust to

the north and south of the strike-slip fault zone: Along profiles 67W and 64W the two

crustal blocks that meet at the fault zone have different characteristics, while along

profile 65W they are of a similar nature. The negative density adjustments in the island

arc crust suggest a lower iron content with respect to typical island arc rocks. Island arc

crust must loose some of its iron in order to eventually become continental crust [e.g. Lee

Page 11: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

4

et al, 2007], and our analysis suggests that this has indeed occurred at the Leeward

Antilles. However, more work from a geophysical and geochemical perspective is

required to further explore this hypothesis.

In the third and last chapter, I use teleseismic data recorded by the BOLIVAR

broadband seismometer array to build a 3D tomographic P-wave velocity model of the

upper mantle under the margin, revealing lithospheric features and imaging the present

configuration of Atlantic and Caribbean subduction. Shallow subduction of a section of

the Caribbean plate under northern Colombia is followed by a stage of steep subduction

that is imaged in our model. Limitations due to station coverage do not allow us to image

the down going Caribbean slab at the location of the Bucaramanga seismic nest. The

work presented in this dissertation suggests that efforts should be made to extend the

mantle image westward in order to investigate the relationship between the slab and this

seismically active zone.

Our image of the Atlantic slab suggests that while subduction terminates at the

strike slip boundary for the upper lithosphere, part of the lower South American

lithosphere subducts along with the Atlantic slab. This partial removal of the lithosphere

would potentially cause isostatic rebound which may have played a role in the uplift of

the coastal mountain ranges in northern Venezuela. Future work will examine

lithosphere-asthenosphere boundary depths in northern Venezuela and this information

should be analyzed in conjunction with uplift and erosion rate data for the coastal ranges

to test this hypothesis.

The Atlantic as well as the Caribbean slab we have imaged reach depths greater

than 700 km, affecting mantle transition zone topography. In the near future we can

Page 12: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

5

expand on this study by updating the transition zone topography results using our mantle

velocity model and building a tomographic S-wave velocity model for the study area.

Combining all these results we can try to place constraints on the temperature and water

content of the two slabs.

The research presented here adds to the growing body of geophysical information

about the Caribbean - South America plate boundary, adds new constraints, and helps

refine our understanding of this much debated margin, while pointing towards new

directions for future research.

Page 13: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

6

Chapter 1

The Caribbean - South American plate boundary at 65°W: Results from

wide-angle seismic data

M. J. Bezada, M. B. Magnani, C. A. Zelt, M. Schmitz, A. Levander

Abstract

We present the results of the analysis of new wide-angle seismic data across the

Caribbean - South American plate boundary in eastern Venezuela at about 65°W. The

~500 km long profile crosses the boundary in one of the few regions dominated by

extensional structures, as most of the southeastern Caribbean margin is characterized by

the presence of fold and thrust belts. A combination of first-arrival traveltime inversion

and simultaneous inversion of PmP and Pn arrivals was used to develop a P-wave

velocity model of the crust and the uppermost mantle. At the main strike-slip fault

system, we image the Cariaco Trough; a major pull-apart basin along the plate boundary.

The crust under the Southern Caribbean Deformed Belt exhibits a thickness of -15 km,

suggesting that the Caribbean Large Igneous Province extends to this part of the

Caribbean plate. The velocity structures of basement highs and offshore sedimentary

basins imaged by the profile are comparable to those of features found in other parts of

the margin, suggesting similarities in their tectonic history. We do not image an abrupt

change in Moho depth or velocity structure across the main strike-slip system, as has

been observed elsewhere along the margin. It is possible that a terrane of Caribbean

island arc origin was accreted to South America at this site, and was subsequently

bisected by the strike-slip fault system. The crust under the continental portion of the

This chapter is a reformatted version of the article "The Caribbean - South American plate boundary at 65°W: Results from wide-angle seismic data" J. Geophys. Res., 115, doi:10.1029/2009JB007070, 2010. Copyright (2010) American Geophysical Union. Reproduced by permission of the American Geophysical Union.

Page 14: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

7

profile is thinner than observed elsewhere along the margin, possibly as a result of

thinning during Jurassic rifting.

1.1. Introduction

The Caribbean-South American (CAR-SA) plate boundary in the southeast

Caribbean is a -300 km wide diffuse zone of deformation that accommodates the ~ 20

mm/yr eastwards displacement of the Caribbean plate relative to South America. Along

the central and eastern Venezuelan coast most of this motion is accommodated along the

right-lateral strike-slip San Sebastian-El Pilar (SS-EP) fault system. The main tectonic

elements of this area include an accretionary wedge, active and extinct island arcs, a large

dextral strike-slip system, and coastal thrust belts with associated foreland basins (Figure

1.1).

Many factors contribute to making this an area of broad geological interest

including 1) the potential accretion of the Leeward Antilles to the South American

continent as a mechanism for creating new continental crust, 2) the relationship between

margin processes and the formation and accumulation of large oil and gas reserves, and

3) the high population density in northern Venezuela that is at risk from geohazards,

particularly earthquakes and landslides, resulting from margin dynamics. Due to all these

factors, the boundary has been the subject of geological and geophysical studies for

decades [e.g. Hess andEwing, 1940; Hess, 1966; Silver etal, 1975; Dieboldet al, 1981;

Burke, 1988; van der Hilst and Mann, 1994]. Despite the abundance of studies and

extensive seismic exploration for petroleum in this region, data that image the

lithospheric structure at the plate boundary is limited [e.g. Schmitz et al, 2005].

Page 15: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

8

Figure 1.1: (a) Shaded relief topography of the study area showing the BOLIVAR wide angle seismic profiles (70W,67W,65W and TRIN): Offshore MCS (red lines), OBS stations (white circles), Texan stations (black dots) and land shots (stars). Black box shows the area of our focus, shown in (c) EG - Espino Graben; CMS - Caribbean Mountain System; SI - Serrania del Interior, (b) Gravity map of the study area, offshore portions of wide-angle profiles shown in black; CAR - Caribbean; SA - South America, (c) Detail of topographic map in area surrounding profile 65W, symbols as in (b). LBI -La Blanquilla Island; MI - Margarita Island; MB - Margarita Basin; LTI - La Tortuga Island; LRC - Los Roques Canyon; MF - Moron Fault; ECT, WCT - East and West Cariaco Trough; SSF - San Sebastian Fault; EPF - El Pilar Fault; UF - Urica Fault; GTF, PTF - Guarico and Pirital Thrust Fronts.

The BOLIVAR (Broadband Ocean-Land Investigation of Venezuela and the

Antilles arc Region) project was designed to image the structure of the lithosphere along

the entire southeastern Caribbean margin [Levander et ah, 2006]. The active-source

component of BOLIVAR provided wide-angle and reflection seismic data along a series

of 5 boundary-normal profiles (Figure 1.1) intended to image the margin at different

stages of evolution after the early Cenozoic CAR-SA collision. This paper will focus on

Page 16: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

9

the profile at ~65°W longitude, hererafter referred to as 65W. We present a P-wave

velocity model along the profile, derived from a hybrid inversion technique that

combines first-arrival tomography and layer-based inversion of deeper seismic phases.

The 65W profile approximately bisects the plate boundary zone at the strike-slip

fault system, extending from the South Caribbean Deformed Belt (SCDB) in the north, to

the Oriental Basin and Espino Graben to the south, crossing the La Blanquilla High, the

Los Roques Canyon, the Margarita Basin, and the Cariaco pull-apart basin in the

Barcelona Bay (Figure 1.1). Understanding of the margin in this area has evolved with

the increasing availability of data through time. In the 1990's, several studies were

published that focused on producing balanced cross-sections constrained by the abundant

exploration-scale reflection seismic profiles that have imaged the region's sedimentary

basins. These balanced cross-sections were extended to a crustal scale and structures at

depth were constrained by potential field data [e.g. Passalacqua et ah, 1995]. Schmitz et

ah [2005] presents the first wide-angle seismic survey to study a north-south profile

across the Oriental Basin (OB). The study proposed a depth to basement of ~13 km and

placed some constraints on the depth of the Moho. The profile was extended offshore,

where no wide-angle seismic information was available at the time, using gravity data.

The profile presented here differs from its companion BOLIVAR profiles

[Christeson et ah, 2008; Clark et ah, 2008; Guedez, 2007; Magnani et ah, 2009] in that

it meets the strike-slip system not at the coastal mountains of the Cordillera de la Costa,

but rather at a bay that encompasses a pull-apart basin offshore, and a broad lowland

onshore. Some other distinctive characteristics of the profile include crossing the Los

Roques Canyon, and the eastern end of the SCDB offshore and the Espino Graben

Page 17: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

10

onshore. Additionally, the profile was less well instrumented offshore than the other

BOLIVAR profiles with one sixth as many ocean bottom seismometers (OBSs) deployed.

The smaller number of OBSs results in decreased coverage, particularly towards the

northern end of the profile.

1.2. Tectonic Setting

A review of the tectonic history of the study area begins at the Mid-Jurassic, long

before this region hosted any interactions between the CAR-SA plates, indeed before a

Caribbean plate, per se, existed. Opening of the Atlantic Ocean in the Mesozoic resulted

in the formation of many grabens located today around the Atlantic's margins [e.g.

Burke, 1976]. In northeastern South America, a series of such grabens with a SW-NE

orientation include the Takutu Graben in Guyana [Crawford et al, 1985] and the Espino

Graben in Venezuela [Moticska, 1985]. Continued rifting between North and South

America in the Late Jurassic - Early Cretaceous resulted in the opening of the proto-

Caribbean seaway [Pindell, 1985] while, to the west, the eastward subduction of the

Farallon plate beneath the Americas created a volcanic arc known as the Great Arc of the

Caribbean [Burke, 1988; Pindell and Dewey, 1982]. In the Early Cretaceous (~110 Ma) a

modification in plate configurations resulted in a change in subduction polarity beneath

the Great Arc, which caused a fragment of the Farallon plate to start overriding the proto-

Caribbean; this fragment became the Caribbean plate. During the same time, the

Caribbean Large Igneous Province (CLIP) was formed by processes that remain unclear

[Hoernle et al, 2004], covering much of what is now the Caribbean plate with 5 to 15 km

of additional basaltic crust. As it advanced over the proto-Caribbean, the Caribbean plate

Page 18: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

11

migrated northeastwards relative to SA until it collided with the Bahamas platform in the

Eocene. The collision caused clockwise rotation of the plate and modified its trajectory to

establish a generally eastward motion relative to stable SA [Pindell and Kennan, 2001,

2007] with a small component of convergence that led to highly oblique collision

between CAR and SA. The collision caused the diachronous (eastward younging)

emplacement of the fold and thrust belts onshore Venezuela (known as the Caribbean

Mountain System to the west of the study area and as the Serrania del Interior to the east)

(Figure 1.1), as well as associated foreland basins onshore [Ysaccis, 1997; E. J. Hung,

1997; Ostos et al, 2005]. Emplacement of the onshore fold and thrust belts was

progressively followed by backthrusting north of the Leeward Antilles which resulted in

the formation of the SCDB [Mann et al, 2007]. Fission track data show that shortening

in the eastern Serrania del Interior ceased at ~12 Ma [Locke and Garver, 2005] possibly

reflecting the end of collision. After the end of oblique convergence between CAR and

SA in the Late Miocene the CAR plate acquired its present motion and the plate

boundary evolved to the nearly purely strike-slip margin observed today [Pindell et al.,

2005; Ysaccis, 1997]. Motion along the El Pilar fault system seems to have initiated at

this time coinciding with the start of deposition in the Cariaco-Tuy Basin [Jaimes, 2003].

Recently, Higgs [2009] suggested that onset of motion along the El Pilar fault occurred in

the Pliocene and could be as young as 2.5 Ma. Today, the plate boundary is a 300 km

wide zone of deformation. It is most clearly defined in central and eastern Venezuela

where the SS-EP fault systems account for 80% of the 20 mm/yr of right-lateral

displacement [Perez et al., 2001; Weber et al, 2001].

Page 19: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

12

Profile 65W begins to the north in the Venezuela Basin floored by undeformed

oceanic crust (Figure 1.1) and continues south crossing the SCDB near its eastern end

(only a small expression of the SCDB is observed in a companion profile ~120 km to the

east, suggesting minor underthrusting of CAR under SA [Clark et al, 2008]).

Accordingly, the accretionary prism at this longitude is thinner than in the west, where

the onset of deformation is older. South of the SCDB the profile crosses a bathymetric

high, the La Blanquilla High, part of a tectonic belt revealed by gravity and bathymetry

data containing the Leeward Antilles arc and the Aves Ridge (Figure 1.1b, [Mann,

1999]). Further south, the profile crosses the Los Roques Canyon, a NW-SE trending

basin that exhibits a bathymetric and gravity signature (Figure 1.1). Little has been

published about the tectonic origin of this basin. The Los Roques fault runs along the

southwestern flank of this basin. It has been hypothesized that, prior to the onset of pure

strike-slip conditions; this fault was connected to the Urica fault onshore. This continuous

system would have served as a transfer zone between the underthrusting of CAR under

the SCDB and the compressional structures of the Serrama del Interior [Higgs, 2009]. On

its path between the Margarita and La Tortuga islands, profile 65W crosses the Margarita

Basin (Figure 1.1), an extensional structure that opened in the Paleogene as a result of

rifting between the two islands. Sediment deposition in this basin remained modest until

the Pliocene, when the transtensive regime responsible for the opening of the Cariaco

Trough renewed fault-controlled subsidence [Jaimes, 2003]. Profile 65W encounters the

Cariaco Trough further south, where the seafloor drops to 1400m below sea level,

making this the deepest basin in the southeast Caribbean (Figure 1.1). The Cariaco

Trough has an area of 20 km2 and is composed of two rhomboidal features, the West and

Page 20: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

13

East Cariaco Trough. The profile crosses the smaller East Cariaco Trough. The Cariaco

trough has been interpreted as a pull-apart basin occurring due to transfer of strike-slip

motion from the San Sebastiam to the El Pilar fault zones [Schubert, 1982; Jaimes,

2003]. Opening of the Cariaco Trough initiated in the Late Miocene and fully developed

in the Pliocene through a transtensive regime that has since weakened [Jaimes, 2003].

Jaimes [2003] hypothesized that the releasing bend formed as a consequence of

misalignment of the SS-EP fault system by motion along the Urica fault; a lateral ramp

connecting the Guarico and Pirital thrust fronts.

As the profile enters land, it crosses a thick sedimentary sequence with a multi­

stage subsidence history. The bottom of the sequence is comprised of Paleozoic

sediments that predate the rifting associated with the opening of the Atlantic. The pre-rift

sequence is followed by Jurassic syn-rift and post-rift sediments that filled the Espino

Graben and that are sealed by a Cretaceous and Paleogene passive margin sequence that

rests on Paleozoic pre-rift sediments in areas not affected by the Jurassic rifting episode.

The Neogene foreland basin sediments of the ESE-WNW trending Oriental Basin were

deposited over the passive margin wedge and cap the sedimentary sequence [Salazar,

2006]. A depth to basement of 6-8 km at the axis of the Espino graben has been estimated

from potential field methods [Feo-Codecido et at, 1984]; however, more recent depth

estimates based on reflection seismic data place the basement of the graben (bottom of

Jurassic sediments) at 10-11 km [Salazar, 2006].

Page 21: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

14

1.3. Wide-angle Seismic Data Acquisition

The seismic wide-angle data were acquired using sources and receivers located

both onshore and offshore along the profile. Offshore, 7 OBSs were deployed by the R/V

Seward Johnson II. The OBSs were LC2000 instruments from the Institute of Geophysics

and Planetary Physics at Scripps Institute of Oceanography, with one vertical component

and one pressure component. They were deployed at intervals of ~ 27 km in water depths

ranging from 250 to 2000 m and used a sampling rate of 8 ms. Locations for these

offshore receivers were obtained by inverting the travel times of the water-wave arrivals.

On land, 514 recording stations consisting of one (48 hr recording capacity) or two (24 hr

recording capacity each) RefTek 125-01 ("Texans") single-channel digitizer/recorders

with 4.5 Hz geophones as sensors were deployed at intervals of ~ 500 m. Land stations

were located using handheld GPS units with errors ranging from 5 to tens of meters

depending on the site and atmospheric conditions.

All receivers recorded shots from the airguns of the R/V Maurice Ewing. The R/V

Maurice Ewing's source consisted of a tuned 20-element airgun array with a combined

capacity of 114 1 (6400 in ). The profile was shot twice, once at 50 m shot spacing to

achieve a high fold for multi-channel seismic reflection recordings and once at 150 m

shot spacing to reduce noise from the wave train of previous shots on OBS records for the

wide-angle analysis. In addition to the offshore sources, two land shots (600 kg of

pentolite) located 60 km and 115 km from the coast, provided some reversed coverage of

the land portion of the profile (Figure 1.1).

Page 22: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

15

1.4. Wide-angle Seismic Data

The quality of the OBS data used for wide-angle analysis is generally good. Data

were examined in both the 50 m and 150 m shot interval records as well as both the

hydrophone and vertical component records. Signal processing applied consisted of an

Ormsby bandpass filter with corner frequencies typically at 2 and 15 Hz (varying slightly

on a case-by-case basis). Picks were made in whichever of the records showed the

clearest arrivals. First arrivals are clearly visible in OBS records typically to offsets of

50-70 km. Occasionally, arrivals are identifiable at larger offsets reaching, in one case,

110 km. Moho reflections (PmP) were identified and picked in records from five of the

seven OBSs at offsets ranging from 35 to 95 km (Figure 1.2a).

The quality of the Texan recordings of the offshore airgun shots was variable, but

generally poor. Only a handful of the recording stations (a fraction of those closest to the

coast, up to 35 km inland) yielded record sections with clearly visible arrivals. First

arrivals were picked for 12 of these at offsets as large as 115 km, but more commonly up

to 50-80 km. PmP arrivals were also picked for 10 of these stations (Figures 1.2b, 1.2c).

Unlike airgun shots, the land shots were recorded very well by the Texan stations (Figure

1.2d). First arrivals were clearly visible to offsets as large as 160 km. The PmP phase was

identified and picked in the Texan record sections of both land shots. An additional

reflected phase was also picked in the Texan records of the northern land shot (Figure

1.2d), we interpret this phase as a reflection from the base of the sedimentary sequence

(i.e. the base of the Espino Graben). OBS recordings of the land shots were poor and no

picks were made. This is most likely due to the large offsets (> 90 km) involved as well

as the large spacing between OBSs which makes it difficult to correlate arrivals between

Page 23: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

16

instruments. At a later stage of picking, arrivals corresponding to waves traveling through

the uppermost mantle (Pn) were identified with the aid of arrival times predicted from

preliminary models. Larger uncertainties were assigned to these picks due to the low

signal-to-noise ratio (SNR).

-100 -80 -60 -40 -20 0 20 -200 -150 -100 -50 Distance (km) Distance (km)

-200 -150 -100 -50 -50 0 50 100 150 Distance (km) Distance (km)

Figure 1.2: Record sections of wide-angle data for (a) OBS station 07 located at model distance 164 km, (b) Texan station 004 located at model distance 186 km, (c) Texan station 039 located at model distance 221 km and (c) the northern landshot, located at model distance 257 km. Data are plotted with a reduction velocity of 8 km/s and band­pass filtered from 3 to 15 Hz. First arrivals shown in red, PmP arrivals are shown in green and an intracrustal phase identified in the northern landshot is shown in blue. Offsets are plotted positive to the south.

Uncertainties were assigned to each pick based on the SNR using a 500 ms time

window centered on the pick [Zelt and Forsyth, 1994]. Uncertainty is proportional to

SNR within pre-defined bounds. The lower and upper bounds were 50 ms and 250 ms for

Page 24: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

17

the first arrivals used in the tomographic inversion and 150 ms and 350 ms for the PmP,

Pn and intracrustal arrivals used in the layer-based inversion.

1.5. Wide-angle Traveltime Analysis

The wide-angle data were analyzed using a hybrid, layer-stripping approach. In

the first stage, first arrival traveltime tomography was used to constrain the velocity

structure of the upper ~20 km of the model. This was followed by the inversion of PmP

and Pn arrivals in a layer-based algorithm to constrain the location of the Moho and the

velocity of the lower crust and uppermost mantle. Model assessment was performed as a

final step, to provide an estimate of the reliability of the results. This modeling strategy

was applied to all BOLIVAR wide-angle profiles and a detailed description is available

in Magnani et al. [2009]. For this reason, it will only be broadly described here and we

will focus on the aspects that are different for this profile.

1.5.1. First-arrival inversion

To obtain the tomographic model we use the regularized, iterative, first arrival

traveltime inversion algorithm of Zelt and Barton [1998]. For the forward traveltime

calculations, a finite difference method to solve the eikonal equation based on that of

Vidale [1990] as modified by [Hole and Zelt, 1995] is used. The model is parameterized

in terms of nodes for the forward calculation of traveltimes and as cells for the inversion

of slowness values. The inversion algorithm solves for the slowness values that minimize

an objective function that includes norms for model roughness and data misfit [Zelt and

Barton, 1998]. If the inversion is successful, the resulting model will fit the observed

Page 25: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

18

traveltime data appropriately (x2 of 1, RMS data misfit equivalent to data uncertainty)

while having minimum roughness in the vertical and horizontal directions.

The model that we present here was parameterized with a node spacing and cell

size of 0.5 km in both the X and Z directions for the forward and inverse step,

respectively. With a total model length of 560 km and a depth of 60 km this results in

137,883 nodes and 136,640 cells. The 0 km horizontal reference point on the model was

established at 12°N in keeping with the convention for all BOLIVAR profiles. The

reference depth of 0 km corresponds to the mean sea level with positive values indicating

subsurface depths.

Choosing an appropriate starting model proved to be essential to the success of

the inversion. The first starting model was a simple, pseudo-1-D model consisting of the

following velocity values: 3km/s at the surface/seafloor, 6 km/s at a depth of 8 km and 8

km/s at every depth equal to or greater than 40 km (Figure 1.3). When using this starting

model, the inversion failed to converge to an acceptable % value. However, the inverted

models showed low velocity structures that matched the known location of sedimentary

basins. Therefore, a new starting model was constructed by incorporating the laterally

varying distribution of sedimentary (2.0 -5.5 km/s) velocities (Figure 1.3). The

traveltimes calculated through this new starting velocity model had an RMS misfit of 227

ms or x2 of 4.24. Using this starting model, a final model was reached after 9 iterations

of the inversion algorithm (Figure 1.3). The final model achieved a x2 of 1.14,

corresponding to an RMS traveltime misfit of 90 ms. Ray coverage from the first arrival

inversion is only adequate for the uppermost -17 km of the model with only a few rays

Page 26: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

Dis

tanc

e (k

m)

Dis

tanc

e (k

m)

50

100

150

200

250

300

350

-0' -'

50

100

150

200

Distance (km

) -100

-50

50

100

150

200

Distance (km

) 300

350

1.4

2.0

2.5

3.0

3.5

4.0

4.5

5.0

5.5

6.0

6.5

7.0

7.5

Vp(km/s)

Figu

re 1

.3: P

seud

o-ID

sta

rting

mod

el (t

op le

ft) a

nd th

e co

rres

pond

ing

best

mod

el (b

otto

m le

ft) w

hich

ach

ieve

d a

norm

aliz

ed m

isfit

x2

of 4

.24.

The

mod

el s

hallo

w v

eloc

ity s

truct

ure

in th

is m

odel

was

use

d as

a te

mpl

ate

to p

rodu

ce a

new

sta

rting

mod

el (

top

right

) whi

ch

resu

lted

in th

e fi

nal f

irst

arr

ival

mod

el (

botto

m r

ight

) w

ith a

x2 o

f 1.

14.

Reg

ions

of

the

final

fir

st a

rriv

al m

odel

(bo

ttom

rig

ht)

not

cove

red

by ra

ys a

re b

lank

ed o

ut in

the

Figu

re.

Page 27: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

20

reaching greater depths. The distribution of rays suggests that constraint on the model

from first arrivals is strictly limited to the upper crust (Figure 1.4).

Figure 1.4: Ray coverage from first arrival inversion. Every fifth ray is shown. Arrow indicates location of the El Pilar fault. Dots indicate the location of the Moho resulting from ray-tracing inversion (gray) and tomographic inversion (black).

1.5.2. Layer-based inversion

The velocity model obtained through first arrival traveltime inversion was re-

parameterized to be used as starting model for a new, layer-based inversion (Figure 1.5).

This second inversion method is based on ray tracing and allows for inverting the vertical

position of boundary nodes and the velocities of the nodes that define each layer [Zelt

and Smith, 1992]. The shallowest 20 km of the model (shallowest 10 km at the northern

end) were represented by 10 layers with laterally varying velocities to replicate the final

first-arrival inversion model, including a first layer that represents the seawater. The 11th

layer represents the mid to lower crust, down to the Moho, and the 12 layer represents

the uppermost mantle. The first ten layer boundaries were fixed throughout the inversion

Page 28: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

21

process and only the depth of the nodes representing the Moho and the velocity values

directly above and below the Moho were inverted.

The data used for the inversion consists of PmP traveltime picks from the OBS

and Texan records of airgun shots, as well as the record sections from the two land shots.

A total of 7920 PmP picks were included in the inversion. Additionally, Pn arrivals were

included for a total of 11,670 traveltime picks. Uncertainties in these arrivals are higher

than those assigned to the Pg arrivals used in the tomographic inversion. Larger

uncertainties were given to these arrivals because the larger offsets and smaller SNR

make it more difficult to pick the arrival time and, in the case of PmP, since these events

are not first arrivals and are mixed in the coda of earlier phases.

The Moho in the starting model is mostly flat (Figure 1.5) and originally

parameterized by 24 nodes located at 25 km intervals. Where lateral variability across

the model was required to fit the data, the spacing of the nodes was reduced to 12.5 km

resulting in a total of 26 nodes. The inversion process was guided by alternatively fixing

or inverting the positions and velocities of different nodes until a satisfactory model was

reached. The final model resulting from this inversion will be referred to as the preferred

model henceforth (Figure 1.6). The final misfit of this model has a % value of 1.228 and

an RMS of 231 ms. More specifically, the PmP arrivals are better fit with a %2 of 1.089

and RMS of 213 ms, while the Pn arrivals were fit with an RMS of 295 ms and a %2 of

1.821. Given the PmP and Pn ray coverage, the Moho in this model is constrained

mostly under the center of the profile and only a weak constraint on the velocity of the

uppermost mantle is achieved, also near the center of the profile (Figure 1.7). To model

the reflected phase picked in the northern land shot record, a floating reflector was

Page 29: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

22

introduced. Since this phase is only constrained by one shot, there are no multiple

observations that need to be reconciled making it easy to fit the data very accurately.

Thus, the RMS misfit for this phase is 87 ms, much smaller than the assigned

uncertainties, resulting in % of 0.15.

1.5.3. Model Assessment

Before any interpretations based on the velocity model are made, it is important to

have a clear notion of the robustness and reliability of the results. We do this by

following two procedures: 1) Checkerboard tests provide insight into the resolution of the

first arrival traveltime tomography, and 2) A semi-independent inversion procedure

estimates the depth of the Moho interface. This is the same assessment carried out by

Magnani et al. [2009] where a more detailed description can be found.

-100 -50 0 50 100 150 200 250 300 350 400

Distance (km)

1 I 1 1 i ; 1 1 i i , i I m 1.4 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0 7.5 8.2

Vp (km/s) Figure 1.5: Starting velocity model for the layer-based inversion. Velocity structure above the dotted white line is taken from the first arrival inversion model and held fixed. Starting velocities directly above and below the Moho are indicated in white.

Page 30: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

Distance

-100

OH

10

J 20 .c Q.

^ 3 0

40 H

• i ' ' > ' i • i i i i i i i i i i i i i i • » • • i •

VB SCDB LBH I LRC I MB CT EG/OB

JL -I

-100 -50 • i • > • • i •

150 200

1 i ' ' ' ' i ' i • ' i

300 350 400 50 100 250

1.4 2.0 7.5 8.2 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0

Vp (km/s) Figure 1.6: Final Velocity model for profile 65W shown with a 5X vertical exaggeration (top) and no vertical exaggeration (bottom). Velocities in the lower crust and upper mantle labeled in white in km/s. Isovelocity contour interval is 0.5 km/s; with thicker contours labeled. Green dots indicate the location of OBSs, white dots indicate the location of Texan stations with picked records, red dots indicate the location of land shots. Yellow dots in the bottom of the model correspond to bounce points of modeled PmP reflections referred to in the text as ray-trace Moho, the tomographic Moho is indicated in black. Orange dots at a depth of-10 km between model distances 260 and 280 km represent bounce points of the intracrustal phase interpreted as a reflection from the base of the Espino Graben. VB: Venezuela Basin, SCDB: Southern Caribbean Deformed Belt, LBH: La Blanquilla High, LRC: Los Roques Canyon, MB: Margarita Basin, CT: Cariaco Trough, EG/OB Espino Graben / Oriental Basin. Black arrow shows the location of the San Sebastian Fault, white arrow shows the location of the El Pilar Fault and yellow arrow shows the location of the coastline.

Page 31: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

24

H DISTANCE (km)

Figure 1.7: Ray coverage from ray-tracing inversion of PmP (green), Pn (red) and intracrustal reflections (blue). Every twentieth ray is shown. Dashed lines show the location of the model layer boundaries.

1.5.3.1. First-arrival Checkerboard Tests

We applied a 2-D adaptation of the 3-D checkerboard tests described in Zelt

[1998] to the first arrival inversion process as a mean of assessing model resolution. A

grid of alternating positive and negative velocity anomalies was superimposed on the

starting velocity model, synthetic traveltimes through this perturbed model were

calculated using the actual experiment geometry, and Gaussian noise with a standard

deviation equal to the uncertainties of the real data was added. This was followed by an

attempt to recover the anomaly pattern with our inversion algorithm, keeping the same

values for the free parameters used in the inversion of the real data. The success of the

inversion was measured by calculating the semblance of the recovered and known

perturbed models. Any point in the model with a semblance value greater than 0.7 was

considered to be resolved, meaning that resolution in that point was better than the

anomaly size. To obtain results independent of the polarity and registration of the

Page 32: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

25

anomaly pattern we switched the polarity of the anomalies and spatially shifted the

patterns for each grid size and averaged the results of the semblance calculations for each

case. The procedure was then repeated for different grid sizes (25x2.5 km, 50x5 km,

75x7.5 km, 100x10 km and 150x15 km) and a resolution estimate plot was generated by

interpolating the resolved areas at the grid sizes tested (Figure 1.8).

1

P 10-

-^ r 2 0 -

•*-»

S"30-n

H

-100

._ -— --

— 1

— —

,— — i -

—'— —i—

_;__ ; '

Distance (km) 0 100 200

^&--<Xv

'-— — — — — _

... — —

_ L _ —

_.

cr

t t

' , -!

... • S ^

1 ' =

" ...

_

• - h -

1

» ^ V

- • i

...

--

— -— N~

300

J^.

- f - -

• • • • -

' !

s s r = •

— —

... ...

400 — 1 - ,

1

...,.,. .

— M

N

— -100

i

Distance (km) ) 100 200 300

s 400

. ..^*...

-

0

10 £ -

20 JZ

30 §" O

40

0-

E io-

^ 2 0 -

§ " 3 0 -

Q 40-

r*^«—

— i —

i •" 1 , 1

=5^=n

1 —,

rs-fc— —k>

1

!

--<

: -""*-—

i ^O^O—° ' p" ° - y ^ - v rU

10 E

4-*

30 §•

40

o-_ E io-.* ^ 2 0 4 J

Q-,„ O; 3 0 -

Q 40-

^~o-^a^o-°~~°~~°^ -

f--

'<

— —

•!•» UM r U

! i

\ ? ^ ^ i .

^X : ^*

-100

-20

300 400

Distance (km)

10 20 25 35

Distance (km)

45 55 65 75 85 100

perturbation (%) lateral resolution (km) Figure 1.8: Results of checkerboard resolution tests at different grid sizes for the first arrival inversion and estimated lateral resolution plot (bottom right panel). Areas colored in black in the lateral resolution plot have an estimated lateral resolution better than 25 x 2.5 km, while regions colored in white have an estimated lateral resolution poorer than 150x50 km. Moho as determined by the layer-based inversion is indicated by the bold black line for reference, white dots mark the location of OBS stations, Texan stations with offshore arrivals and land shots. See text for a description of the checkerboard grid sizes and testing procedure.

Page 33: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

26

The checkerboard test results indicate that the shallowest 15 km of the first arrival

model are well constrained, whereas constraint at depths greater than 20 km is poor. The

poor constraint provided by first arrivals at mid and lower crustal depths is not of great

concern however, since this area of the model is constrained by the PmP and Pn arrivals

used in the layer-based inversion. The results of the checkerboard tests give us

confidence that the velocity structure in the shallowest 20 km of the model, which is held

fixed in the second (layer-based) inversion, is reasonably well constrained by the first-

arrival tomography.

1.5.3.2. Moho Reflection Tomography

To test the results of the layer-based inversion we employ a separate regularized

inversion to find the depth to Moho using PmP traveltimes exclusively. The Moho was

modeled as a floating reflector while the entire velocity model was fixed. The velocity

model is a slightly modified version of the final first arrival model, in which we allowed

a maximum velocity of 7.0 km/s by resetting all higher velocities in the model to that

value. Taking the velocity model as a priori information, the inversion algorithm then

seeks to find the flattest (minimum first-spatial-derivative) Moho interface that minimizes

the PmP traveltime residuals to a value consistent with their assigned uncertainties. The

method works iteratively and is analogous to the one described earlier for the first arrival

tomography. This approach is different form the layer-based inversion in that it uses a

finite difference method for calculating the traveltimes [Hole and Zelt, 1995], it is

regularized, and does not include Pn traveltimes. The method is not entirely independent

however, since the PmP traveltimes and upper crustal structure are identical to those used

Page 34: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

27

in the layer-based ray tracing approach. After six iterations, the final interface was

reached, achieving a normalized misfit % of 1.01 and an RMS misfit of 205 ms.

The result of the tomographic PmP inversion agrees well with the ray-trace

inversion Moho, as the two interfaces are within ~2km of each other throughout the

profile (Figure 1.6). Differences are to be expected since the lower crustal velocities

determined by the layer-based inversion are different from the homogeneous 7.0 km/s

value assumed in this method. The largest absolute differences are in the south of the

profile where lower crustal velocities as found by the layer-based inversion are slowest

(6.7 km/s).

1.5.4. Gravity Modeling

Gravity modeling served as an additional test on the feasibility of the P-wave

velocity model. An observed gravity curve (Figure 1.9) was obtained from a 3 minute

(~5.5 km) grid constructed from the Simon Bolivar University gravity database with free-

air anomaly values offshore and Bouguer anomaly values on land [C. Izarra et al, 2005].

The curve represents an average of the observed anomaly over a 30 km wide swath

centered on the profile. A density model was generated by mapping the P-wave

velocities (Vp) of the preferred model into density values (p) using a fifth-order

polynomial fit to experimental data [Zelt, 1989]. For the gravity calculations the model is

parameterized as a series of trapezoidal blocks of constant density following the method

of Talwani et al. [1959].

The calculated gravity for the density model matches the broad characteristics of

the observed anomalies, with an RMS misfit of 23.098 mgal (Figure 1.9). The position

Page 35: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

28

of the Cariaco Trough corresponds to a narrow gravity low where the anomaly reaches =

§7 rnigai The SCDB, as well as the Los Roques Canyon, correspond to broader local

lows in the observed gravity anomaly. The La Blanquilla High, with its high seismic

velocities, and correspondingly high densities, is aligned with a local high in the observed

gravity anomaly (Figure 1.9). On land, the geometry of the sedimentary basins and the

Moho topography in the model result in a 200 km wide, low gravity anomaly. Although

the calculated lateral extent of this anomaly agrees well with the observations, its

absolute amplitude is too large by ~75 mgal.

Distance (km) * " Distance (km)

•100 -50 O 50 100 150 200 250 300 350 400

Distance (km)

Figure 1.9: Results of gravity modeling, (a) Comparison of the observed gravity anomaly (black crosses) with gravity anomalies calculated from a density model derived from the preferred P-wave velocity model without any adjustments (b) and three models adjusted to fit the observed data. In each of the three adjusted models, density adjustments where restricted to one of three geologic domains: sedimentary basins (c), igneous-metamorphic crust (d) or upper mantle (e). VB - Venezuela Basin; SCDB - Southern Caribbean Deformed Belt; LBH - La Blanquilla High; LRC - Los Roques Canyon; MB - Margarita Basin; CT - Cariaco Trough; EG/OB - Espino Graben / Oriental Basin. Non-linear (top) color bar used for (b) and red-white-blue (bottom) color bar used for (c), (d) and (e).

Page 36: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

29

We attribute the misfit between the observed and calculated gravity anomalies

chiefly to the non-uniqueness of the Vp-p relationship. Given the scatter of the empirical

data on which the relationship is based, a range of density values is permissible for any

given velocity value [P. J. Barton, 1986; Ludwig et al, 1970]. To investigate how the

data misfit could be reduced by adjusting the densities of the blocks that make up the

model, we used a forward modeling approach. We explored three different scenarios

considering density adjustments restricted to three geologic domains: sedimentary basins

(original densities between 1100 and 2550 kg/m ), igneous-metamorphic crust (original

densities between 2550 and 3000 kg/m3), and upper mantle (original densities larger than

3000 kg/m3).

The modeling approach is based on defining regularly spaced density adjustment

(Ap) nodes for each domain. We obtain the Ap for each model block within the domain

by linearly interpolating between the nearest corresponding nodes. Through trial-and-

error forward modeling we found sets of Ap for each of the three domains that

substantially reduced the data misfit. The spacing of the Ap nodes increased from 10 km

for the sedimentary basins to 20 km for the igneous-metamorphic crust and 40 km for the

upper mantle, following the assumption that lateral variability of densities is highest in

the shallow subsurface and decreases with depth. The three resulting models show an

improvement in the RMS data misfit, from the original 23.10 mgal to 14.27 mgal when

adjustments were restricted to the upper mantle, 10.21 mgal when they were restricted to

the igneous-metamorphic crust and 10.51 mgal when they were restricted to the

sedimentary basins. Because of the shallower depth and smaller size of the blocks that

make up the sedimentary basins, adjustments in this domain can better fit small scale

Page 37: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

30

structure but require larger amplitude adjustments to reproduce the observed gravity

anomalies. The crystalline crustal adjustments achieve the maximum misfit reduction (a

fraction of 1 mgal better than the sedimentary adjustments); while the mantle adjustments

achieve a smaller, but still significant reduction of RMS misfit. The difficulty in fitting

the data adjusting only the densities of mantle blocks is due to their depth and size that

make them least capable of fitting small-scale anomalies. In all cases, the density

adjustments range between -130 kg/m and 220 kg/m and in the case of the crust and

mantle they range between -130 and 150 kg/m .

We acknowledge two major drawbacks to this forward modeling approach: 1) It

does not amount to a thorough exploration of the Ap model space, and 2) it is intrinsically

limited due to the restrictions on the distribution of density adjustments imposed. It is

possible, even likely, that the real densities of the sedimentary basins, the igneous-

metamorphic crust, and the upper mantle, all vary from the values we infer from their P-

wave velocities; instead of these variations being restricted to only one of these domains

as we have assumed for the sake of simplicity. In spite of these limitations, the forward

modeling we have conducted shows that relatively minor adjustments in the densities

derived from the P-wave velocity model, that are well within the scatter of experimental

Vp-p determinations, can achieve a good fit to the gravity data.

1.6. Results

In this section we present a description of the preferred P-wave velocity model

(Figure 1.6). Like all of the BOLIVAR velocity models [Christeson et al, 2008; Clark et

al, 2008; Guedez, 2007; Magnani et al, 2009], the 65W model has a high degree of

Page 38: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

31

lateral heterogeneity and shows considerable Moho topography. Offshore, we image a

series of sedimentary basins separated by basement highs. The SCDB, the Los Roques

Canyon, Margarita Basin and the Cariaco Trough sedimentary basins are characterized by

Vp values ranging between 1.9 and 5.4 km/s. The coincident MCS data reveal that these

basins are bounded by faults that put the sedimentary sequences in contact with igneous-

metamorphic rocks (Figure 1.10), resulting in a laterally alternating pattern of high and

low velocities in our model. Similar patterns have been observed in the neighboring

BOLIVAR profiles [Clark et al, 2008; Magnani et al, 2009]. Onshore, there is good

correspondence between the location of the Espino Graben/Oriental Basin and Vp values

between 3.0-5.5 km/s. The deeper crustal velocities vary laterally and are consistent

within tectonic domains. In the following paragraphs we will discuss the characteristics

of each of the structures imaged starting in the north with the SCDB and continuing

southwards.

At the north of the profile, at model distances -50 to 0 km we image the SCDB

near its eastern end. Velocities at the seafloor are 1.9 km/s and increase to 5.4 km/s at a

maximum depth of ~11 km. Lower crustal velocities appear to be high compared to the

rest of the model reaching 6.9 km/s at the base of the crust; however this result may not

be significant due to the relatively poor constraint at the northern edge of the model. The

Moho under the SCDB is modeled at a depth of 25 km and average crystalline crustal

velocities are 6.3 km/s. Due to lack of ray coverage, the Venezuela Basin, north of the

SCDB, is not imaged. Immediately south of the SCDB we image the La Blanquilla High

which shows very high crustal velocities even at shallow depths (6.35 km/s at 4.5 km

depth) as a result of the very thin or absent sedimentary cover blanketing the igneous

Page 39: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

N

Sou

th

Car

ibbe

an

Def

orm

ed B

elt

Los

Rog

ues

Can

yon

Mar

garit

a H

igh

• i

n •

Car

iaco

B

asin

§

o ..<

Vp

(km

/s)

Fig

ure

1.10

: T

ime

mig

rate

d M

CS

line

alon

g of

fsho

re p

ortio

n of

pro

file

65

W (

top)

with

lin

e dr

awin

g in

terp

reta

tion

of m

ain

si

(mid

dle)

and

lin

e dr

awin

g in

terp

reta

tion

over

lain

on

the

offs

hore

por

tion

of t

he p

refe

rred

vel

ocity

mod

el c

onve

rted

to

time

The

re i

s go

od c

orre

latio

n of

the

shal

low

vel

ocity

str

uctu

re w

ith th

e se

dim

enta

ry b

asin

s an

d ba

sem

ent

high

s ob

serv

ed i

n th

e M

C

Gre

en c

ircl

es i

ndic

ate

the

loca

tion

of O

BS

stat

ions

.

Page 40: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

33

basement. The average crystalline crustal velocity is 6.4 km/s, the highest along the

profile. The Los Roques Canyon between 40 and 75 km in model coordinates is imaged

as a 30 km wide basin that deepens steeply towards the south, where it reaches a

sediment thickness of 8 km. The average velocity of the sedimentary column is 3.8 km/s,

the slowest along the profile. The Moho follows a generally gentle, southward deepening

trend throughout the offshore portion of the profile, going from ~26 km depth at model

distance 25 km to ~28 km at the coastline (model distance 200 km). The Margarita Basin

(model distances 90 to 150 km) also exhibits velocities of 1.9 km/s at the sea floor and

shows a sedimentary thickness of- 6.5 km. This basin is flanked to the north and south

by blocks of high velocity, slightly above 6.0 km/s, that reach shallow depths.

Between 150 and 185 km, the Cariaco Trough is clearly imaged as a narrow,

deep, V-shaped basin. At model distance 163 km the basin reaches its deepest point

where the contact between the sedimentary sequence (7.6 km thick) and the igneous

basement is at 9 km depth and the sedimentary column in 7.6 km thick. The average

velocity of the sedimentary column both on the Cariaco Trough and the Margarita Basin

is 4.0 km/s and crystalline crustal velocities under the two basins average 6.10 and 6.17

km/s respectively. South of the Cariaco Trough, crystalline crustal velocities greater than

5.5 km/s reach shallow depths near the coastline. As the profile enters mainland, Moho

deepens southwards from 27.5 km at the coastline to a maximum depth of 39.8 km at

model distance 300 km, about 100 km inland. From its deepest point, the Moho shoals to

the south to a depth of ~34 km at the southern end of PmP ray coverage (model distance

375 km). The upper crustal velocity structure on land shows a broad basin that

encompasses the Espino Graben as well as the younger overlying sediments of the

Page 41: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

34

Oriental Basin. The boundary between the two is not conspicuous in the velocity

structure. The average velocity of the sedimentary rocks on land (4.3 km/s) is greater

than the average velocity of the sedimentary cover of any of the offshore basins;

reflecting older, more consolidated sedimentary units onshore, and young, less

consolidated sediments offshore. The reflection identified as the base of the Espino

Graben (Figure 1.2d) is modeled at 10.6 km depth between the 5 km/s and 5.5 km/s

contours (Figure 1.6). The average velocity of the SA plate crystalline crust is 6.3 km/s

1.7. Discussion

1.7.1. Southern Caribbean Deformed Belt and Venezuela Basin

The profile images the SCDB near its eastern end, where the sedimentary prism is

youngest and least deformed. As noted above, only a small expression of this structure is

observed further east [Clark et al, 2008a]. Crustal thickness is high, as there are -15 km

of crystalline crustal velocities beneath the SCDB sediments (Figure 1.6). The unusually

large thickness of the oceanic crust could immediately be attributed to the presence of the

Caribbean Large Igenous Province [Coffin and Eldholm, 1994]. However; previous

seismic reflections studies have concluded that the southeast edge of the Caribbean

(including our study area) is floored by normal oceanic crust [Diebold et al, 1981], and a

companion BOLIVAR profile at 67°W longitude found oceanic crust of normal thickness

-240 km west of our study area [Magnani et al, 2009]. Although we must acknowledge

that data coverage is relatively poor in the northern edge of our profile, our observations

are inconsistent with oceanic crust of normal thickness. Additionally, the bathymetry of

the Venezuela Basin shows that while the depth of the sea floor is -5 km where the crust

Page 42: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

35

has been found to be of normal thickness, it is shallower (4.0-4.7 km) along our profile

(Figure 1.1). The shallower seafloor suggests the presence of a thicker, more buoyant,

crust, in accordance with our observations. Relatively thick oceanic crust (10-11 km) is

also found -125 km east of our study area, There, the excessive thickness of the crust has

been interpreted to result from the proximity to the Aves Ridge instead of the presence of

the CLIP [Clark et al, 2008a]. The wide-angle seismic observations presented here

suggest that the CLIP is present in the eastern edge of the southeast Caribbean.

1.7.2. Island Arc Region

Between the South Caribbean Deformed Belt and the Cariaco Trough, the Moho

depth is relatively stable, dipping gently towards the south. In this region, the profile

crosses basement highs as well as deep basins. We will discuss this portion of the

velocity model in the context of observations across the other BOLIVAR wide-angle

seismic profiles.

Immediately south of the SCDB, our profile crosses the extinct volcanic arc of the

Leeward and Venezuelan Antilles at the bathymetrically high La Blanquilla High. Based

on the lateral continuity of gravity anomalies, Mann [1999] proposed that the Leeward

Antilles arc, the La Blanquilla High and the Aves Ridge are part of one continuous

volcanic island arc, an hypothesis strongly supported by the similarities in the average 1-

D velocity curves of these features as determined by wide-angle seismic data along

BOLIVAR profiles 64W, 67W, 70W and TRTN [Clark et al, 2008a; Magnani et al,

2009; Guedez, 2007; Christeson et al, 2008]. The average 1-D velocity profile at the La

Blanquilla High along profile 65W is remarkably similar to the velocity structure of the

Page 43: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

36

belt sampled at different locations (Figure 1.11a), lending additional support to the

hypothesis of a continuous tectonic belt that includes all the aforementioned structures.

South of the La Blanquilla High, the profile crosses a NW-SE trending trough

known as the Los Roques Canyon (Figure 1.1). Little information is available about this

basin from the scientific literature. The geometry of the basin, as inferred from the

distribution of sedimentary (<5.5 km/s) velocities in the model, appears to be distinctly

asymmetric; with the sedimentary cover deepening and thickening to the southwest. The

geometry of the sediments filling the narrow trough suggests the presence of a NE-

dipping steep fault (Los Roques Fault), bounding the NE edge of the high velocity

basement beneath the Margarita Basin. The location of the canyon between two

bathymetric highs and its NW-SE orientation suggest a comparison with the West

Curacao Basin, which has been imaged by BOLIVAR profile 70W [Guedez, 2007]. The

West Curacao Basin formed in the Early-Middle Miocene along with a series of basins

that dissected the Leeward Antilles ridge offshore western Venezuela, whose subsidence

was controlled by NW-SE trending faults [Gorney et al, 2007]. A comparison of the

average 1-D velocity structure of the two basins (Figure 1.11b) reveals great similarity,

specially if one accounts for the difference in bathymetry that makes the West Curasao

Basin ~2 km shallower than the Los Roques Canyon. One important difference is the

depth to Moho, which is significantly greater beneath the West Curasao Basin. The

similar sedimentary velocities suggest that the two basins may have been formed by

similar processes, and the Los Roques Canyon may be a product of rifting between

fragments of the Leeward and Venezuelan Antilles Arc. Extension at the Los Roques

Canyon probably occurred later than at the West Curasao Basin, given the generally

Page 44: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

10

20

E

a a>

a 30

40

50

LBH

65W

LBH

64W

Less

er A

ntill

es

Ave

s R

idge

Leew

ard

Ant

illes

lot

20

Q.

o

30

40

50

Los

Roq

ues

Can

yon

Wes

t C

urac

ao B

asin

M

arga

rita

Bas

in

Bon

aire

Bas

in

02

46

8 0

24

68

Vel

ocity

(km

/s)

Vel

ocity

(km

/s)

Vel

ocity

(km

/s)

Fig

ure

1.11

: O

ne-d

imen

sion

al v

eloc

ity p

rofi

les,

(a)

Com

pari

son

of 1

-D v

eloc

ity p

rofi

les

aver

aged

acr

oss

the

Lee

war

d A

ntill

es A

rc,

the

La

Bla

nqui

lla H

igh

as i

mag

ed b

y B

OL

IVA

R p

rofi

les

64W

and

65W

, the

Les

ser

Ant

illes

Arc

and

the

Ave

s R

idge

, (b)

Com

pari

son

of 1

-D v

eloc

ity p

rofi

les

aver

aged

acr

oss

the

Los

Roq

ues

Can

yon

and

the

Wes

t C

urac

ao B

asin

as

wel

l as

the

Mar

gari

ta B

asin

and

B

onai

re B

asin

, (c

) C

ompa

riso

n of

1-D

vel

ocity

ave

rage

d ov

er 5

0 km

sec

tions

nor

th a

nd s

outh

of

the

mai

n st

rike

-slip

bou

ndar

y al

ong

BO

LIV

AR

pro

file

s 64

W, 6

5W a

nd 6

7W.

Page 45: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

38

eastward younging progression of tectonics in the southeast Caribbean. The difference in

crustal thickness between the two basins may explain the difference in bathymetry and

thus, the shallower position of the West Curacao Basin with respect to the Los Roques

Canyon. A different interpretation of the nature of the Los Roques Canyon is given by

Higgs [2009]. This author proposes that the Los Roques Fault, running along the western

flank of the canyon, is the offshore continuation of the Urica Fault. This fault system

would have formed part of the CAR-SA plate boundary before the Pliocene by linking

the SCDB and the thrust and fold belts of the Serrania del Interior. After the Pliocene, he

suggests conditions switched from oblique convergence to pure strike slip; displacing the

Los Roques Canyon towards the east and causing the current misalignment with the

Urica Fault. This interpretation fails to explain the relationship between the fault and the

basin and whether the fault has controlled the subsidence of the basin. The latter seems

unlikely, if the fault was indeed a lateral ramp connecting two thrust fronts.

Further south, our profile crosses the Margarita Basin, where deposition began in

the Paleogene and resumed in greater volumes in the Neogene after a period of tectonic

inversion in the Middle to Late Miocene [Jaimes, 2003]. In the case of the Margarita

Basin we look for a reference in the Bonaire basin, located at similar latitudes 250-400

km to the west. The velocities of the sediment column in the Margarita Basin are

significantly higher (-30-40%) than in the Bonaire Basin, but the velocities in the

igneous-metamorphic crust (Vp > 5.5 km/s) as well as Moho depth, are very similar

(Figure 1.11b). This observation indicates that the two basins may be underlain by a

similar kind of crust but are probably not genetically linked.

Page 46: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

39

In general, we observe that the crust in this portion of the profile is comparable to

the corresponding areas on other BOLIVAR profiles and has been interpreted as a rifted

island arc crust [Magnani et al, 2009]. The sedimentary basins overlying the crust

appear to be the result of a complex tectonic history involving periods of extension in

different directions, as well as episodes of compression, and clockwise rotation

[Beardsley and Ave Lallemant, 2007]. In the case of the Los Roques Canyon we find it

analogous to other geometrically similar basins along the margin.

1.7.3. South American Continent

As noted above, the Moho deepens southwards from ~28 km at the coastline,

reaches a maximum depth of 39.8 km -100 km inland under the Espino Graben/Oriental

Basin, and then shoals again towards the south following a trend that is similar to what is

observed -300 km to the west in BOLIVAR profile 67W [Magnani et al, 2009]. This

characterization of the Moho topography differs from the geometry modeled by Schmitz

et al. [2005] and Schmitz et al [2008] along a coincident profile, where the Moho depth

monotonically decreases towards the north. Average crustal velocities in the continental

portion of the profile are higher than those occurring north of the coastline with the

exception of those found in the La Blanquilla High.

The extension that formed the Espino Graben likely weakened and thinned the

continental crust that we image along our profile. Crustal thinning has been observed

along similar grabens also formed during the opening of the Atlantic (eg. Benue Trough

[Fairhead and Okereke, 1990]) and 3D gravity inversion along the Espino Graben itself

also suggested the presence of a thinned crust [Durdn, 2007]. If we interpret the top of

Page 47: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

40

the crystalline basement to coincide with the 5.5 km/s contour, then crustal thickness in

the continental section of this profile is -25 km, much lower than the -35 km observed

along profile 67W [Magnani et al, 2009]. The implications of this difference in crustal

thickness shall be discussed below. Our estimate of the thickness of the sediments

onshore is constrained by the velocity structure as well as the modeled depths of the

intracrustal reflections observed in the record of the northern land shot. Where the

sedimentary column is thickest, the contact with the basement is at -11 km depth; this is

consistent with the 5.5 km/s velocity boundary in the Schmitz et al. [2005] model, as well

as the most recent, reflection-seismic derived, estimates for the base of the Jurassic

sequence in the Espino Graben of 11 km [Salazar, 2006].

1.7.4. Cariaco Trough and the Strike-Slip Fault System

Lastly, we will address what is perhaps the most provocative result of this study,

the nature of the Moho and the crust across the EP strike-slip fault system that constitutes

the current plate boundary between the Caribbean and South American plates.

BOLIVAR research led to the development of a shear-tear hypothesis for the rupture of

subducting Atlantic (oceanic South American) lithosphere from continental South

American lithosphere. This hypothesis is based primarily on observations along profile

64W, -100 km east of our study area, and predicts the existence of the following features

[Clarke/a/., 2008b]:

1) A step in the depth of the Moho (deeper in the south) at the strike-slip boundary

that juxtaposes Caribbean and South American crust. This step should become

more gradual westwards as the crust "relaxes".

Page 48: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

41

2) A downward flexure of the northern edge of South American lithosphere which

should be most pronounced at the site of the tear and decrease progressively

westwards.

Observations along profile 67W are all consistent with these predictions [Magnani et al.,

2009]. Therefore, the expected result along profile 65W would be an intermediate state

between what is observed at 64W and 67W; i.e., a significant step in the Moho at the site

of the strike-slip boundary and a relatively deep Moho south of the boundary. The results

presented in this work differ greatly from those expectations, as our seismic observations

are clearly incompatible with the predicted crustal features: A step in the Moho is not

observed at the strike-slip boundary and the maximum Moho depth along the profile of

39.8 km is shallower than the 45 km depth observed along profile 67W.

The base of the crust under the major strike-slip system (coincident with the

Cariaco Trough) is densely covered by PmP reflections (Figure 1.7) recorded in one OBS

station (Figure 1.2a) as well as several Texan stations on land (Figures 1.2b, 1.2c),

suggesting that our experiment geometry is sensitive to variations in Moho depth at the

strike-slip boundary. If there was indeed a step in Moho depth, the observed travel times

for PmP reflections would reflect it. First arrival ray coverage is densest slightly to the

north of the strike-slip boundary (Figure 1.6), which allows us to have a high level of

confidence in the upper crustal velocities in that portion of the model as well.

The velocity models, and locally averaged ID velocity profiles on the 64W and

67W models clearly show a step in Moho depth, as well as slightly lower crustal

velocities at any given depth on the southern side of the strike-slip boundary, in what is

interpreted as South American continental crust (Figure 1.11c). In the model presented in

Page 49: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

42

this study, not only is there no difference in Moho depth across the boundary but the

velocity profiles on both sides of the boundary are practically indistinguishable from each

other and from the northern ID profile at 67W (Figure 1.11c). This suggests that the

crustal blocks that meet at the El Pilar fault zone at this longitude are similar. We note

that a step in crustal thickness occurs about 50 km to the south of the surface expression

of the strike-slip fault system.

One possibility is that the Cariaco-Tuy Basin is underlain by a crustal block of

Caribbean origin similar to what is interpreted in the 67W profile to be extended island

arc crust. This idea in itself is not new. Wells drilled in the Cariaco-Tuy basin revealed

that its basement is composed of Cretaceous metavolcanics interpreted to have developed

in an island arc setting [Talukdar and Bolivar, 1982]. Passalacqua et al. [1995]

incorporate a high density "crustal indenter" into their crustal model based on gravity and

magnetic data that reaches south of the El Pilar fault. In a recent overview of the margin

Ostos et al. [2005] classify the Tuy-Cariaco basin as part of the "Venezuelan Platform"

Cenozoic allochthon along with the Triste Gulf, north of the strike-slip system along

profile 67W.

What our observations suggest, is that the basement of the Cariaco Basin could

indeed be a whole crust block that has been separated from the main body of Caribbean

Island Arc crust and effectively accreted to South America (Figure 1.12) given the

current plate motions as measured by GPS that show that the San Sebastian-El Pilar fault

system is the main boundary between the two plates [Perez et al, 2001]. The mechanism

by which this accretion would have occurred is difficult to deduce given the 2D nature of

this study and the fact that our profile crosses only the eastern end of this hypothetical

Page 50: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

N

SS-E

P

Str

ike

Slip

Fa

ult S

yste

m

OH

10H

J 20

Q.

S3

0H

40

H

50

Ori

enta

l Bas

in

and

Es

pino

Gra

ben

5-

Car

ibbe

an

Larg

e Ig

enou

s P

rovi

nce

Vol

cani

c 7

Isla

nd A

rc

Cru

st

Ext

ende

d

Vol

cani

c Is

land

Arc

C

rust

C

ontin

enta

l SA

Cru

st

Mon

o

Moh

o

-100

-5

0 0

—[—

r—

i—i-

50

100

150

200

250

300

350

400

Dis

tanc

e (k

m)

Fig

ure

1.12

: Sc

hem

atic

int

erpr

etat

ion

of t

he S

outh

Am

eric

an -

C

arib

bean

pla

te b

ound

ary

alon

g pr

ofile

65W

bas

ed o

n th

e ve

loci

ty

mod

el o

f Fi

gure

1.6

. Se

dim

enta

ry b

asin

s ar

e sh

own

in w

hite

, unc

onst

rain

ed M

oho

dept

hs a

re s

how

n as

thic

k da

shed

lin

es. S

trip

ed a

rea

sout

h of

the

str

ike

slip

fau

lt sy

stem

is

inte

rpre

ted

as e

ither

an

accr

eted

ter

rane

of

Car

ibbe

an i

slan

d ar

c or

igin

or

Sout

h A

mer

ican

pa

ssiv

e m

argi

n cr

ust.

VB

-

Ven

ezue

la B

asin

; SC

DB

-

Sout

h C

arib

bean

Def

orm

ed B

elt;

LB

H -

L

a B

lanq

uilla

Hig

h; L

RC

-

Los

R

oque

s C

anyo

n; M

B -

Mar

gari

ta B

asin

; SS

-EP

- Sa

n Se

bast

ian-

El

Pila

r.

4^

u>

Page 51: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

44

crustal terrane. Here we merely present this as a hypothesis that will have to be further

developed and tested by future work.

A second possible explanation for the similarity of the velocity and crustal

structure across the strike-slip system is that the crust along the northern edge of the

South American plate is heterogeneous and exhibits different characteristics in this area

than elsewhere along the CAR-SA boundary. The SA continental crust at this location

may simply resemble Caribbean Island Arc crust. In any case, it is important to consider

that the mismatch of our observations and the predictions of the shear-tear hypothesis do

not negate said hypothesis, but rather are most likely the result of the impact of the

heterogeneity of the northern South American margin on the dynamic of the shear tear

process. The Espino Graben is a first order crustal heterogneity and, therefore, we are

intuitively inclined to assume it must have played some role on the dynamics of the

response to lithospheric tearing. On a first glance it is clear that it may help explain why

the Moho depth is shallower than expected; since the crust seems to be thinner under the

graben than in areas not affected by Jurassic rifting, as outlined above.

The answer to which one, if any, of these two scenarios is correct will have to

result from future work that takes into account the distribution of depocenters onshore

and offshore, the morphology of the fold and thrust belts of the Caribbean Mountain

System and Serrania del Interior, the timing of their uplift and, perhaps, physical and/or

digital modeling that explores the possible effects of the Espino Graben on the advancing

thrust fronts during the convergence period. Physical modeling has already shown that

pre-existing basin morphology affects the geometry of fold and thrust belts [Macedo and

Page 52: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

45

Marshak, 1999]. These experiences could provide a starting point for modeling the

problem at hand.

1.8. Summary of conclusions

The crust under the SCDB is ~ 15 km thick, suggesting that the Venezuela Basin

in our study area is floored by Caribbean Large Igneous Province crust, and that the CLIP

extends to parts of the southeast Caribbean that were previously thought to be floored by

normal oceanic crust. The Moho under the La Blanquilla High is -25 km deep, and the

average ID P-wave velocity profile across this structure is similar to that observed at the

Leeward Antilles Arc, and at the Aves Ridge, providing further support for Mann's

[1999] hypothesis that these features form a single volcanic belt. The northwest trending

Los Roques Canyon has a velocity structure that resembles that of the West Curacao

Basin -430 km to the west. The similar orientation and velocity structure of the two

basins suggests that they might be the result of analogous processes, and the Los Roques

Canyon may have resulted from rifting within the Venezuelan Antilles - Aves Ridge arc.

The crustal structure between the Margarita Basin and the shoreline shows little variation,

with a gently dipping Moho from 26.6 km depth under the northern edge of the Margarita

basin, to a depth of 27.5 km under coast. We do not observe a sharp contrast in Moho

depth and crustal velocities across the El Pilar fault zone as was expected from

BOLIVAR work in other areas of the margin. The crustal block immediately south of the

El Pilar Fault exhibits a Moho depth and a velocity structure similar to the Bonaire basin

at longitude 67°W, interpreted as extended Caribbean island arc crust [Magnani et ah,

2009], and it may represent an accreted terrane of Caribbean origin. We observe an

Page 53: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

46

increase in crustal thickness at the southern edge of the Cariaco trough, i.e. the coastline,

from -28 km to -40 km. Onshore, the thick sediment column that comprises deposits of

the Espino Graben and Oriental Basin has a maximum thickness of 11 km. The contact

between the Jurassic graben fill and the younger deposits is not evident in the velocity

model. The Moho under the continental SA plate has a concave downwards shape and

reaches a maximum depth of 39.8 km -100 km inland. The continental crust is thinner

along our profile than in other areas of the margin, possibly due to crustal thinning

associated to the Jurassic rifting of the Espino Graben.

Page 54: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

47

Chapter 2

Gravity inversion using seismically derived crustal density models and

genetic algorithms: An application to the Caribbean - South American

plate boundary

M. J. Bezada and C. Zelt

Abstract

Crustal density models derived from seismic velocity models by means of velocity-

density conversions typically reproduce the main features of the observed gravity

anomaly over the area but often show significant misfits. Given the uncertainty in the

relationship between velocity and density, seismically derived density models should be

regarded as an initial estimate of the true subsurface density structure. In this paper we

present a method for estimating the adjustments necessary to a seismically derived

density model to improve the fit to gravity data. The method combines the Genetic

Algorithm paradigm with linear inversion as a way to approach the non-linear and linear

aspects of the problem. The models are divided into three layers representing the

sedimentary column, the crystalline crust and the lithospheric mantle; the depths of these

layers are determined from the seismic velocity model. Each of the layers is divided into

a number of provinces and a density adjustment (A/?) value is found for each province so

that the residual gravity (difference between the observed gravity anomaly and the

anomaly calculated for the seismically derived model) is minimized while keeping Ap

between predefined bounds. The preferred position of the province boundaries is found

through the artificial evolution of a population of solutions. Given the stochastic nature of

the algorithm and the non-uniqueness of the problem, different realizations can yield

Page 55: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

48

different solutions. By performing multiple realizations we can analyze a set of solutions

by taking their mean and standard deviation, providing not only an estimate of the Ap

distribution in the subsurface but also an estimate of the associated uncertainty. Synthetic

tests prove the ability of the algorithm to accurately recover the location of province

boundaries and the Ap values for a known model when using noise-free synthetic data.

When noise is added to the data, the algorithm broadly recovers the features that define

the known model despite greater standard deviations of the solutions and the occurrence

of artifacts in the mean solutions. The algorithm was applied to four profiles across the

Caribbean - South America plate boundary. Some general patterns in the distribution of

Ap were observed consistently in the profiles and are correlated with the interpretations

of the velocity models: Positive Ap values in the sedimentary layer, negative Ap values

for island arc and extended island arc crust with an abrupt change to positive values in

South American crust, and positive values in the mantle under the continent and island

arc with a transition to negative values under the Caribbean oceanic crust.

2.1. Introduction

A powerful and widely used means of studying lithospheric structure and

composition is the construction of seismic velocity models from earthquake and

controlled source traveltime observations. Given the correlation between P-wave

velocities (Vp) and densities (p) it is common practice to convert the velocity models

obtained from seismic methods into density models in order to test their compatibility

with gravity data. The conversion can be carried out with the many empirical and

theoretical Vp-p relations that are available in the literature, some of the most widely

Page 56: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

49

used ones being the Nafe-Drake curve (polynomial regressions to this curve by Zelt

[1989] and Brocher [2005]), relations derived specifically for sedimentary rocks [e.g.

Gardner et al, 1974; Hamilton, 1978] and relations specific to crystalline crust and upper

mantle rocks [e.g. Christensen & Mooney, 1995; Godfrey et al,. 1997; Birch, 1961;

Sobolev & Babeyko, 1994] (Figure 2.1). Researchers can choose between these formulas,

or a combination of them, according to a priori geologic information and subjective

preference. Density models thus derived often produce gravity anomalies that correlate

well with the observations, but commonly have significant misfits, even when lithology-

specific Vp-p relations are used and careful measures are taken to account for the effects

of pressure and temperature [e.g. Ravat et al, 1999; Korenaga et al, 2001]. This is to be

expected, since the data on which these relations are based show significant scatter

(Figure 2.1). The variability of the Vp-p relation is caused by its dependence on a

number of factors including mineralogy, composition, presence of melt, and porosity

structure [Korenaga et al, 2001]. The effects of variations in these parameters can be so

large as to call into question the usefulness of seismic velocities as effective constraints

on densities [Barton, 1986]. Therefore, although the Vp-p relations are certainly

valuable, any density model derived using them must be regarded as an initial

approximation and the need for subsequent model refinement should be expected. Site-

specific Vp-/7 relations will be found as a result of refining the density models to fit the

observed gravity anomalies, which may help estimate or constrain some of the

parameters that control the relation. Additionally, a spatially systematic variation of the

Vp-/? relation could help define otherwise inconspicuous boundaries of tectonic units.

Page 57: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

Vp (

km/s

) F

igur

e 2.

1: N

afe

and

Dra

ke d

atas

et (

grey

dot

s) a

nd c

omm

only

use

d W

p-p

rela

tions

. Thi

n bl

ack

lines

rep

rese

nt t

he Z

89 c

urve

± 0

.25

Mg/

m3. R

efer

ence

s: Z

89 -

Zel

t, 19

89; G

74 -

Gar

dner

et a

l, 19

74; B

05 -

Bro

cher

, 200

5; G

97 -

God

frey

et a

l, 19

97; B

61 -

Bir

ch, 1

961;

C

M95

- C

hris

tens

en &

Moo

ney,

199

5.

o

Page 58: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

51

Forward modeling is commonly used to refine the density models derived from

seismic velocities [e.g. Horsefleld et ah, 1994; Barton & White, 1997; Hirsch et ah,

2009; Klingelhoefer et ah, 2009; Sumanovac et ah, 2009], and is sometimes guided by

assumptions concerning the lateral variations in chemical composition, lithology,

mineralogy and/or temperature [e.g. Ravat et ah, 1999; Korenaga et ah, 2001]. While

certainly a valid and widely used approach, forward modeling has significant drawbacks

when applied to gravity modeling, given the inherent non-uniqueness of the problem.

This issue can be addressed by producing a set of different models that incorporate

different assumptions. As an alternative to forward modeling, inversion schemes can be

applied. However, in order to make the problem suitable for inversion techniques,

restrictions are usually imposed on the distribution of the density adjustments [e.g.

limited to a single layer: Kauahikaua et ah, 2000; Korenaga et ah, 2001]. Both of these

approaches result in a limited exploration of model space and cannot provide a

statistically valid estimate of the range of parameters that can fit the data adequately.

Therefore, when using these methods, it is difficult to estimate the errors on the model

parameters.

In this paper, we present an alternative approach to this problem based on the

Genetic Algortihm (GA) paradigm, which aims to find suitable solutions to a problem by

mimicking biological evolution [Vose, 1999]. GAs are a powerful and versatile tool for

global optimization and are an efficient method for the exploration of model spaces with

local minima. They have been shown to be useful in the analysis and modeling of gravity

data both in 2D and 3D [e.g. Boschetti et ah, 1997; Roy et ah, 2002; Montesinos et ah,

Page 59: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

52

2005; King, 2006; Chen et al, 2006]. The methodology we present here is based on a

multi-realization GA that also incorporates linear inversion as part of the optimization

process. The goal of this algorithm is to recover the density adjustments Ap required to

improve the fit of a seismically-derived 2D crustal density model. We have applied this

method to four active-source wide-angle profiles of the BOLIVAR project [Levander et

al., 2006] across the Caribbean - South America plate boundary.

2.2. Description of the algorithm

2.2.1. Residual anomaly

A 2D initial density model is obtained from a 2D seismic velocity model by using

a given Vp-p relation. In the examples presented in the following sections, we have used

a polynomial regression to the Nafe and Drake curve [Zelt, 1989] hereafter referred to as

Z89. To calculate the initial gravity anomaly we parameterize the density model as a

series of trapezoidal blocks of constant density and use the method of Talwani et al.

[1959]. In order to avoid edge effects, the models are padded on either end by laterally

extrapolating the densities at the edges of the model for one model length. We obtain the

residual anomaly by subtracting this initial calculated anomaly from the observations and

removing the mean.

Page 60: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

53

2.2.2. Parameterization

While the initial model can be arbitrarily complicated and can contain an arbitrary

number of polygons distributed in any number of layers, in order to make the

implementation of the GA feasible we simplify the problem as follows:

a) We divide the model into three layers representing the sedimentary cover, the

crystalline crust and the upper mantle. We obtain the depths of the boundaries

between these layers from the velocity model and assume these boundaries to be

well constrained, and thus, fixed.

b) We assume that each of these layers can be divided laterally into a small number

of "provinces" and that the provinces are separated by vertical boundaries. The

number of provinces and the location of the boundaries in each layer are

independent of the other layers.

c) We assume that Ap is constant within each province and limited to a certain

range (AyC n <Ap < A / w ) .

By making these simplifications, the inversion process effectively results in static

shifts in the model densities relative to the values in the initial model, with the shift being

constant within a province. The question of the optimal number of provinces per layer

will be discussed in section 3.2. The values of AyOfoax and Apbin were fixed in this

implementation to ±0.25 Mg/m3 based on the distribution of empirical data points about

the Nafe-Drake curve (Figure 2.1).

Page 61: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

54

2.2.3. Solution representation and evaluation

Since the general architecture of the GA method has been extensively described

elsewhere [e.g. Goldberg 1989; Vose 1999; Sivanandam and Deepa, 2007 ], we will

leave the details of our implementation to appendix A. In this section we will discuss the

evaluation operator, which is the central part of the algorithm. Each density adjustment

solution can be viewed as consisting of two parts: the geometry of the blocks (provinces)

and the value of the density adjustment (Ap) in each block. Although the problem of

calculating the gravity anomaly resulting from a 2D body has a non-linear dependence on

the geometry of the body, it depends linearly on its density. Therefore, we approach the

two parts of the problem in different ways. Since the layer boundaries are assumed to be

fixed, the geometry of the blocks is determined by the set of parameters defining the

province boundaries. It is these parameters that will represent a particular solution

competing for survival in the GA. Using GA terminology, this can be thought of as the

solution genotype. Due to the linear dependence of the gravity anomaly on the block's

density, once we have defined the geometry of the blocks we can use the formulation of

Talwani et al. [1959] to calculate a matrix G that relates the Ap values to the residual

gravity anomaly they produce by Gm=deS( where m is a vector containing a given set of

Ap and dest is the corresponding residual gravity anomaly. Conversely, given the values

of the residual anomaly dres and the matrix G we can perform a linear inversion to obtain

the values of m that minimize ||Gm - dres ||, thereby finding the set of density adjustments

that would best explain (in a least squares sense) the observed gravity residual, given the

predefined block geometry. Since we do not place constraints on the inversion, the

Page 62: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

55

elements of m are free to take values outside the permitted range, so we obtain the vector

m' by imposing minimum and maximum values (Aphnn and A / ^ ) on the elements of m'.

The vector m' can be thought of as the solution phenotype as it is the density

distribution that results from the province geometry and the bounds placed on A/?. We

then calculate the gravity anomaly generated by the reset model by d'est = Gm', and

finally obtain a measure / of the fitness of the solution by calculating the RMS misfit

between dres and d'^,. The fittest solutions are then those with the smallest/values, and

we will refer to the fittest solution (minimum/) in a population as the a solution. We let

the solutions compete and evolve for a fixed number of generations and retrieve the a

solution at the end of the run, which we consider to be one realization. We note that,

when applying inverse methods, it is often desirable not to minimize data misfit but to

obtain a normalized data misfit of unity (misfit equivalent to data uncertainty;

[Bevington, 1969]). However in the case of crustal scale 2D gravity modeling,

uncorrelated noise stemming from observational error is negligible and other factors,

difficult to estimate before the inversion, play a more important role as discussed in

section 3.3. It is for this reason that the concept of normalized misfit is not used in our

evaluation function.

2.2.4. Analysis of the results

Since the GA method is based on a guided but random process, after a set number

of generations, it converges to a different a solution in each run depending on the seed of

the random number generator. Although GA's are designed to avoid local minima, there

is such a high level of non-uniqueness in the problem at hand that many different

Page 63: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

56

combinations of model parameters can fit the data almost equally well. We speculate that

if the process was allowed to run indefinitely, the global minimum might be found each

time, independent of the random seed used. However, given the slow pace of

convergence after a given number of generations we find it more practical and efficient to

stop the convergence at a given point and repeat the process a number of times to obtain

different approximations to the global minimum. There is an added benefit to this

approach in that it allows us to examine the variability in the set of model parameters that

fit the data at a given level.

The representation of each solution based on the location of the province

boundaries does not allow for the direct comparison of two solutions (two significantly

different sets of boundaries could result in very similar density distributions), so we

resample the density distributions resulting from the a solutions emerging out of each

realization into 1 km wide cells. We can then perform an elementary statistical analysis

on the resampled density distributions to obtain the mean Ap value and standard

deviation for each 1 km wide cell. We will refer to this average as the mean a solution.

2.3. Synthetic test

In designing a synthetic model to test the algorithm, we used the geometry of the

P-wave velocity model from BOLIVAR profile 67W [Magnani et al, 2009] as a

template. The Moho interface was taken directly from the velocity model and the

crystalline basement interface was constructed from the 5 km/s contour in the velocity

model. We prescribed a series of density anomalies on each of the three layers as follows:

We placed positive (0.10 Mg/m) and negative (-0.17 Mg/m ) density anomalies in the

Page 64: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

57

north and south of the sedimentary layer respectively, a negative density anomaly (-0.13

Mg/m3) in the crystalline crustal layer near the center of the model and we divided the

upper mantle layer into two zones of negative (-0.15 Mg/m ) and positive (0.15 Mg/m )

density anomalies in the north and south respectively, with a transition in three steps

between the two zones near the center of the model (Figure 2.2). This results in four

provinces in the sedimentary layer, three in the crystalline crustal layer and five in the

uppermost mantle layer. This distribution of density anomalies was meant to pose a

challenge to the methodology since negative and positive density anomalies in different

layers of the model coincide laterally, producing competing effects on the residual

gravity curve. The synthetic residual anomaly has a peak-to-trough amplitude of ~ 350

mgal and is clearly dominated by the long-wavelength effect of the upper mantle

anomalies (Figure 2.2).

ApflVlg/nrn

N -200 -150 -100 -SO

100

_. 0

~i 1 r-0 50 100 150 200 250 300

- i 1 r -

/^v^^n

-150 -100 0 50 100 Distance (km) SMi

0.25

0.20

0.15

0.10

0.05

0

-0.05

•0.10

-0.15

-0.20

-0.25

Figure 2.2: Known model for synthetic tests (bottom) and associated residual gravity (top)

Page 65: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

58

2.3.1. Noise-free residual gravity

For a first test, we use the synthetic residual gravity as input to the algorithm and

we set the number of provinces to match those of the known model. We ran the algorithm

for 150 generations per realization and 50 realizations. The results are satisfactory in that

the mean a solution recovers the location and amplitude of most of the anomalies

accurately (Figure 2.3). The distribution of the individual a solutions and the values of

the standard deviations indicate that the level of constraint on the anomalies is variable

(Figure 2.3). The density anomalies seem to be much better constrained in the south of

the model (less variability in the individual a solutions and smaller standard deviations),

where there are shorter wavelength lateral variations in the topography of the layer

boundaries. The northern half of the sedimentary layer and the northernmost third of the

igneous-metamorphic layer show the least constraint on Ap. This is likely due to their

small thicknesses that make it easy to compensate a Ap there with a laterally coincident,

small change in the Ap value of another layer. The RMS misfits of each of the fifty a

solutions vary from 0.5 to 1.3 mgals with the mean being 0.8 mgal (Figure 2.4).

Meanwhile, the RMS misfit of the mean a solution is 0.4 mgal, better than any of the

individual a solutions. Another noteworthy result is that the areas where the anomalies

are not recovered correctly are those where the uncertainty is greatest (Figure 2.3).

Regarding the lateral boundaries of the anomalies, the northern boundaries of the

negative sedimentary anomaly and the positive mantle anomaly, as well as the southern

boundary of the crustal anomaly are recovered successfully. In contrast, the boundaries of

the positive sedimentary anomaly and the northern boundary of the crustal anomaly are

not recovered very well, although sharp gradients or small steps are observed in the mean

Page 66: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

59

a solution at the location of the prescribed boundaries (Figure 2.3). The magnitudes of

the mantle anomalies, the negative sedimentary anomaly and even the poorly laterally

constrained positive sedimentary anomaly are recovered accurately, while the crustal

anomaly is underestimated in the mean a solution by -20%. The stepwise transition

between the two mantle zones is not recovered in the mean a solution. Instead there is a

zero anomaly zone that is partially in spatial agreement with the prescribed zero anomaly

block, connected to the high and low density anomalies by gradients (Figure 2.3).

ISO JOU 2SO 300

Ap(Mg/rrT)

0.25

a

Figure 2.3a: Results of 50 realizations of the algorithm, using the residual gravity from the known model shown in figure 2.2 and the same number of provinces as the known model in each layer. From top to bottom: Observed (black crosses) and calculated (red line) residual gravity curves, RMS misfit value shown is for mean a solution; mean a solution with the boundaries of the known density anomalies superimposed in gray for reference, and plots of Ap vs model distance for the sedimentary, crustal and mantle layers. In the 3 bottom plots each of the individual a solutions are shown as thin gray lines, the mean a solution is shown by the red line, the green shaded area shows one standard deviation above and below the mean and the known model is shown for references as the dashed blue line.

Page 67: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

60

N 100

i s -ioo

»?*

-200

••jtso -05ft - I & J -y

RMS- l.Sau)tl

rt M IOC

/ • '

t w j

n"'\^ 10 250 ilKJ

J^^-SK^'^'-I

1 '

/v

•200 -ISO -IOO -50

-200 150 -100 -SO

100 150 200 350 MO

SO 100 150 2QQ 2S0

*- - - - 7

• •

! ' i ^ ^ ^ -

• .Jt 1 •

-200 -150 -100 -SO

•)0Q -150 -1U0

RMS • 0.2 mqa:

-SO 0 SO 100 IS!) /GO >W 100

, . ; J ^ C •»*-**"-*

0 2

0.1

I » •0.1

•03

<u

0.1

# " -0.1

fl.2

-200

-200

-150

-ISO

'100

-100

50

SO

0 50 100 Dlftwwe [km)

0 50 IOO

150

iw

200

200

250

250

300

300

• • -

P ,

f . ,

.

-100 -150 -100 -50 0 50 100 ISO 200 250 300 - . D I M M K C [km) •

N 5 — -ISO -100 SO 0 50 IOO 150 200 350 300

Ap(Mg/nrn

10.25

0.20

0.15

0.10

0.05

0

-0.05

-0.10

-0.15

0.20

•0.25

j | j |

-200 ISO -100 -SO SO 100 ISC •100 ISC 100 -50 0 « IOO ISO I M ISO 300

Figure 2.3b-e: Same as figure 2.3a but using one fewer province per layer with respect to known model (b), one additional province per layer (c), three additional provinces per layer (d) and five additional provinces per layer (e).

Page 68: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

61

2.3.2. Incorrect number of provinces

This set of tests takes into account the fact that the number of provinces used to

subdivide each layer is generally an unknown. It is therefore important to test the

robustness of the algorithm with respect to this parameter. We consider four cases: The

first one uses one fewer province in each layer with respect to the true model, the second

one uses one additional province in each layer and the third and fourth use three and five

additional provinces in each layer, respectively.

The results show that, for a fixed number of realizations, the ability of the

algorithm to fit the data increases with the number of provinces used. As the number of

provinces increases, we see an improvement of the RMS misfit of the mean a solution

and, for numbers of provinces matching or exceeding those in the true model, a general

reduction of the variability of the sets of solutions. In the case where we assumed one

fewer province per layer with respect to the true model, the variability of the a solutions

is very low, with the distribution of solutions seemingly bimodal in the sedimentary

layer. This low variability is misleading since it gives the impression of a tight constraint

even in areas where the true anomaly is not recovered well and results from the

limitations imposed by the small number of provinces. In all cases, the RMS misfit of the

mean solution is at least as small as the smallest RMS misfit of the individual a solutions

(Figure 2.4). Overall, the mean a solutions for all cases contain the main features of the

true model and the RMS fit is excellent (0.1-1.0 mgal). The northern boundary of the

positive sedimentary anomaly and the magnitude and northern boundary of the crystalline

crustal anomaly are the most difficult features to recover. These features become

increasingly well resolved as the number of provinces increases and they are recovered

Page 69: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

62

a

0 1 2 3 4 5 # of additional provinces

0 1 2 3 4 5 # of additional provinces

Figure 2.4: RMS misfit (a) and difference with respect to known model (b) as a function of the number of additional provinces used in the algorithm for the noise-free synthetic test. Vertical lines extend from the minimum to the maximum value of the individual a solutions, open circles indicate the mean of those values and diamonds show the values corresponding to the mean a solution.

accurately in the case with five additional provinces per layer. In that case, the

northernmost ~50 km of the sedimentary and crystalline crustal layer (where these

difficult to resolve features are located) show the largest standard deviations of Ap,

reflecting the fact that a variety of combinations of density anomalies in these two layers

can explain the synthetic gravity values in this part of the model. The algorithm is unable

to recover the stair-step transition between the low and high density mantle anomalies

when more than three additional provinces per layer are added. Instead, the mean a

solution shows a density gradient connecting the two anomalies. A quantitative

assessment of the success of the inversions is possible by calculating the L2 norm of the

difference between vectors containing the resampled Ap values from the true model and

each of the individual a solutions as well as the mean a solution for each of the cases

considered. Figure 2.4 shows that the difference between the mean a solution and the

Page 70: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

63

true model decreases monotonically with increasing number of provinces per layer. The

trends seen in the graphs presented in Figure 2.4 suggest that the addition of a greater

number of additional provinces per layer would produce only incremental improvements

in the RMS misfit and difference with the true model of the mean a solution.

We conclude that the algorithm can successfully recover a set of density

anomalies if the key assumptions we have made are met (i.e. the layer boundaries are

known accurately from the velocity model, the residual gravity is caused by static shifts

in the Vp-p conversion and these shifts are constant within a province). It is not necessary

to know a priori the number of provinces that each layer should be divided into, as the

results are robust with respect to this parameter and satisfactory solutions are found while

using a number of provinces that differs from that of the true model. Moreover, it is

desirable to use a greater number of provinces than is strictly needed, as this helps the

algorithm converge to the correct solution.

2.3.3. Residual gravity with added noise

In the next set of tests we explore the robustness of the algorithm with respect to

noise to the residual gravity curve. Due to the precision and accuracy of gravity meters,

noise in gravity data is practically negligible in crustal scale studies. However, we have

assumed that the layer boundaries that we define from the velocity model are accurate,

but we should expect some degree of error depending on the quality of the seismic data.

Our formulation also ignores 3D effects that could be locally significant, and the simple

parameterization we use cannot represent small bodies with anomalous densities within a

province. We added noise to the synthetic residual anomaly to simulate randomly

Page 71: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

64

distributed density contrasts not accounted for by our parameterization. Since the intent is

not to simulate uncorrelated noise, instead of adding a random value to each residual

gravity point, we create a smooth noise curve by placing Gaussian random values with a

standard deviation of 8 mgal at equal intervals of 20 km and resample them to match the

observation interval of 5 km (Figure 2.5). We repeated the set of tests performed with the

noise-free synthetic data set to gauge the effect of the number of tectonic provinces

assumed on the set of solutions found. The results show that even after adding noise to

the residual gravity curve, the algorithm is able to broadly recover the main features of

the prescribed model (Figure 2.6).

N

100

=• o n>

£ •"„ -100

Si

-200

•3f\r\

-200 -150 -100 -50 0 50 i i i i i i

^ - . - - ^ - ' • " • - a

• i i i i i

100 150 200 250 300 i i i i i

eL-~'m'*^'~'***. —

i i i i i

-200 -150 -100 -50 0 50 100 150 200 250 300 Distance (km)

Figure 2.5: Noisy synthetic residual gravity curve (crosses) resulting from adding random noise described in the text (dashed gray line) to original synthetic residual gravity curve (solid gray line)

As was the case for the noise-free synthetic test, we consider five cases where we

varied the number of provinces per layer with respect to the true model from -1 to +5. In

all of the five cases, we recovered the negative and positive density anomalies in the

mantle. The magnitude of the negative mantle anomaly is more accurate and robust

(smaller standard deviations), but that of the positive anomaly is generally adequate

Page 72: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

65

except in the case with five additional provinces per layer. The stepwise transition is not

recovered well, but it is suggested by the mean a solution in the cases with the correct

number of provinces and with one additional province per layer. The negative crystalline

crustal anomaly is also present in the mean a solution in all the cases considered. The

southern boundary of this anomaly is a very robust feature and is recovered accurately in

all cases, while the northern boundary is only apparent in the results when fewer than

three additional provinces per layer are used. The magnitude of the crystalline crustal

anomaly is underestimated by 33 to 69%. The recovery of the anomaly magnitude

improves as the number of provinces per layer with respect to the true model increases

from -1 to 1, but no additional improvement is seen when a larger number of provinces

per layer is used (Figure 2.6). In the sedimentary layer, the northern boundary of the

negative anomaly is recovered accurately in all cases but the magnitude is severely

underestimated. Only 50% of the magnitude of this anomaly is recovered when using the

correct number of provinces, recovery worsens slightly when one additional province per

layer is added and is very poor (-25% of the true value) in the cases with a larger number

of additional provinces per layer. The positive sedimentary anomaly is generally

overestimated (up to 100%). In this case, the estimate does improve consistently with the

addition of more provinces per layer. The lateral boundaries of this anomaly are

recovered only approximately, but within -25 km of the corresponding boundaries in the

true model (Figure 2.6). Additional, laterally small anomalies not present in the true

model appear in the mean a solutions (e.g. at -200 km, Figure 2.6) as a result of the

noise. These artifacts become more pronounced with the addition of provinces

Page 73: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

66

Overall, although the fit to the data improves, the quality of the mean solution

decreases if the number of additional provinces with respect to the true model is greater

than one (Figure 2.7). Quantitatively, the norm of the vector difference between the mean

a solution and the true model is smallest when one additional province per layer is

considered and increases for a larger number of additional provinces (Figure 2.7). While

the additional provinces do not improve the estimate of the true model, they do produce a

significant increase in the standard deviation of the individual a solutions, making the

interpretation of the set of solutions more difficult.

-200 -ISO -100 -SO O SO 100 ISO 200 MO 300

DliUnw (km I

Figure 2.6a: Results of 50 realizations of the algorithm, using the noisy residual gravity from the known model shown in figure 2.2 and the same number of provinces as the known model in each layer. Panels as in figure 2.3a.

Page 74: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

67

N

« : /; -30!

100 150

R.V.S - 1.1

• _ _ ^

-IOC

TCJd

, ^ ^

^

so

0 »

/

——~-_

:oo m ;DD

/ £ ^

r ^

250 3C0 • _ . . ^

w

-7 -200 -150 -100 -SO 0 SO 100 ISO 200 2S0

Distance (km)

•200 ISO -100 -50 0 SO COO 150 2 0 0 3 5 0

-IDO -ISO

RMS - 2 J or:

_ . ^ t ^ ^

rco

si

'->

-SO

.__

0 SO 103

y ^

^SO 200 250 100

\ c

•200

•200

*7~

•ISO

-ISO

_PTT

~

too

100

s

-50 0 SO 100 DcHwt te (hml

•50 0 SO 100

, -mm- ;M--iifc ™ ~ ' w

150 200

8S0 200

W ; i ' ;

1 •f" •*

2S0

__ KO

<m' MO

300

300

•* .

Ap(Mg/m3)

02

01

1 « -0.1

•0.2

0.2

at

"i o -0.1

•02

»•

"

-._

. :l . — —. i . _ _ .

— M • M -Ml <M "

. >...- _.

.•m-m~*r-m::mrm

. .

.11

."' \ .""".

m m-'^ wt~~ m"

• ; ?

. N -200 -ISO -100 -SO

RMS ' M S m j a l

• ^ ^ ^ ^

0

-y^

SO 100 350

Z^' 200 250 300

^~~~. J - = - - =

-200 -ISO -DQO -50

-200 -ISO -100 -50

; - « " " > ' i i : ' V

0 50 100

D f s u n t o f c m )

0 SO 100

• J . - - ' " ' . - - - - . .

tmwm - -

ISO

ISO

-V : '1,

200

2O0

ISO 1 0 0

2 5 0 3 0 0

'- * - - - .

fciirlii.""*»- •>• m^im V

0.25

0.20

0.15

0.10

0.05

0

-0.05

-0.10

-0.15

-0.20

-0.25

Figure 2.6b-e: Same as figure 2.6a but using one fewer province per layer with respect to known model (b), one additional province per layer (c), three additional provinces per layer (d) and five additional provinces per layer (e).

Page 75: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

68

We can conclude that using a large number of provinces when applying the

algorithm to real data degrades the quality of the results and leads to the strengthening of

artifacts in the solutions. Our results suggest that many additional provinces do not result

in better approximations to the true model and complicate the interpretation of the sets of

solutions given the larger scatter produced. These tests suggest that, when using real data,

this method would be most successful when the number of provinces assumed in the

algorithm exceeds the number required or expected by one or two.

An encouraging result that arises from this set of tests is that by looking at the

mean a solution and the distribution of the solutions it is possible to discern what the

main features of the true model are and what the level of uncertainty is. As elaborated

above, the mean a solution in all cases contains the transition from low to high density in

the mantle, the low density anomaly in the crystalline crust near the center of the model

and the high and low density anomalies in the north and south of the sedimentary layer

respectively. These are all the features that define the true model, and although their

magnitudes are not always recovered well, the polarities are correct and the recovered

locations of their lateral boundaries are fairly accurate. In these tests, additional density

anomalies not present in the true model are seen in the mean a solutions. This implies

that, as with most geophysical techniques, attention should be paid to the possible

presence of artifacts in the solutions. Laterally small anomalies that occur in areas of the

model otherwise known to be geologically homogenous should be considered potential

artifacts; meaning not that they should be ignored but that their interpretation should be

approached with caution and alternative explanations for the gravity residual they cause

should be considered.

Page 76: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

69

0 1 2 3 4 5 # of additional provinces

-1 0 1 2 3 4 5 # of additional provinces

Figure 2.7: RMS misfit (a) and difference with respect to known model (b) as a function of the number of additional provinces used in the algorithm for the noise-free synthetic test. Vertical lines extend from the minimum to the maximum value of the individual a solutions, open circles indicate the mean of those values, diamonds show the values corresponding to the mean a solution.

2.4. Application to real data

In this section we describe the application of our procedure to 2D velocity models

of four transects across the Caribbean-South American plate boundary produced as part

of the BOLIVAR project [Levander et al, 2006]. The four boundary-normal profiles we

will consider resulted from the analyses of first-arrival and reflected traveltimes from

controlled-source seismic data. The profiles were named according to their approximate

longitude: 70W [Guedez, 2007], 67W [Magnani et al, 2009], 65W [M. J. Bezada et al,

2010] and 64W [S A Clark et al, 2008]. The four profiles cross a series of tectonic

provinces roughly perpendicularly. These provinces are, from north to south, the

Venezuela basin over the Caribbean seafloor, the Southern Caribbean Deformed Belt, the

Leeward Antilles extinct volcanic island arc and extended island arc, the strike slip fault

Page 77: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

70

system, the coastal ranges (with the exception of profile 65 W) and sedimentary basins on

land (Figure 2.8).

The gravity anomaly values over the profiles were interpolated from a 3' grid of

land, ship and satellite measurements [C. Izarra et al, 2005](Figure 2.8). The data

represent Bouger anomaly on land and free-air anomaly offshore. For the initial gravity

calculations, the models were parameterized as a set of trapezoids with constant density

(Figure 2.9) and padded at the edges as described in section 2.1. The density of each

trapezoidal block was determined from the block's average velocity using the Z89 curve.

The location of relative high and low gravity anomalies calculated from the initial models

correlate well with the observations (Figure 2.9) but the level of misfit is significant, with

original RMS misfits ranging from 23 to 69 mgal (average 51 mgal). Previous attempts to

fit the gravity data for two of the profiles relied on forward modeling and constrained the

density adjustments to laterally varying depth intervals to avoid the problem of tradeoffs

between density anomalies at different depths. These modeling efforts reduced the RMS

misfit to a minimum of 11 mgal [Magnani et al., 2009; Bezada et al. ,2010].

Based on the experience with the noisy synthetic data set described in the

previous section, and the number of tectonic provinces observed on the surface and

interpreted from the seismic velocity models, we chose to use six provinces per layer

when applying our algorithm to these data. As was the case for the synthetic models

described above, we ran the algorithm for 150 generations and 50 realizations. The mean

a solutions achieve a significant reduction of the misfit with respect to the original

density models, reaching RMS values of 4.5 to 7.7 mgal (Figure 2.10).

Page 78: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

Mga

l

250

200

150

100 50

0

-50

-100

-150

-200

-250

-300

-350

Fig

ure

2.8:

Gra

vity

map

of

the

sout

heas

t C

arib

bean

sho

win

g th

e lo

catio

n of

BO

LIV

AR

pro

file

s. D

ata

from

Iza

rra

et a

l. (2

005)

sho

ws

Bou

guer

ano

mal

y on

land

and

fre

e-ai

r an

omal

y of

fsho

re.

CA

R -

Car

ibbe

an P

late

; SA

- So

uth

Am

eric

an P

late

; VB

- V

enez

uela

Bas

in;

SCD

B -

Sout

hern

Car

ibbe

an D

efor

med

Bel

t; L

A -

Lee

war

d A

ntil

les;

SSF

Z -

Stri

ke S

lip F

ault

Zon

e.

Page 79: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

72

ISO -100 -50 0 SO 100 150 200 2S0 300 -150 -100 -50 0 50 100 150 200 250 300 350 400

<1.433 1.886 2.127 2.279 2.431 2.592 2.800 2.946 3.115 3.373

Density (Mg/m3)

Figure 2.9: Initial density models for four BOLIVAR profiles across the Caribbean -South American plate boundary: 64W (a), 65W (b), 67W (c) and 70W (d). The density models were derived from P-wave seismic velocity models using the Z89 curve. Observed gravity anomaly shown as black crosses, calculated anomaly as red line. RMS misfits are indicated in the figure.

As expected, the character of the sets of solutions is similar to what is seen in the

noisy synthetic test, and there is significant variation in the individual a solutions (Figure

2.10). Based on the noisy synthetic results, we expect to encounter artifacts in the mean a

solutions and will focus the analysis of the solutions on the polarity and the location of

lateral changes in Ap. In general, in the mean a solutions for each of the four profiles we

observe a good correlation between lateral changes in Ap and the boundaries of tectonic

provinces interpreted in the velocity models (Figure 2.11). We will describe the mean a

Page 80: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

73

-SO 0 SO 103 ISO 200 250. _ 3C0_ 300 400

641V ISO

ISO

•100

•100

-so

-so

0

0

50 100 Diitartce (km)

50 100

ISO

ISO

230

200

250

250

100

300

— „ _ ! '_

H--,[J

, —.—r—| [.-..!'• 'r

' ,

"1

__J-™-="

—".

1.

•r..s • __

, .1

'1

'~n

± j .

l i ••

, + > - . . • - -

...

02

0.1

"I ' •0.1

-0.2

:.,_

ffi

1 '

.'Jl

?

~ .'_>:

* <

••

• h t:i_ .

1 " - T -

7;J i 7 - - r -^ ' -

.. '.?.: • a_. - .

-150 -100 -50 0

N -200

^

-150 -tOO -50 0 SO t 10 ISO 200

, *AW 3S0

<% wo

I;:^

67W -200 -ISO

•230 -150 •50 0 59 ICO ISO 2 0 0 3<0 3OT

--

:-

_.!

- 4 —

• — —

i

I

—. -

. —..._ h . - • • - - • * . • •••

W T • .. ;-

1 , • : - - | i

. !.' . J . . .

i • -, • "

_.......

~:

' .—

'

' — ~ ~ ' " • . • ' . - . . • . ~ : . ' -

; • - " - - , - 4 - - " •

F " - I - ' - • — •' - i :

w : v ; • J . .

-. • 1 - -

,

"--"!' •200 -ISO -tOO -SO

100 if- RilS:. 7.9 mgd

^ A.:>- //"•"' s ^ - - " ^ -" ^ « ! j .

j b

6SW

QJ

0.1

- 0 1

-0.2

•100

, . . _ _ . . £ _ r -----

0

0

V- -' -; ;;

) ( » 200 KManee(km)

100 200

, ! ! •

wo

300

400

400

•fl

J

Ap(Mg/rri )

,0.25

w -2S0 -200 -ISO -tOO -50

200 300

0 SO 100 ISO JOU

70W •250 -200 -ISO 50 100 150 200

SO 100 IW Jtl»

0.20

0.15

0.10

0.05

0

-0.05

-0.10

-0.15

-0.20

-0.25

Figure 2.10: Results of 50 realizations of the algorithm for the residual gravity and layer geometries from BOLIVAR profiles 64W (a), 65W (b), 67W (b) and 70W (d). Panels as in figure 2.3a.

Page 81: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

64W -60.

-150 -100 -50 50 100 150 200 250 300 Distance (km)

Ap (Mg/nrO

10.25

0.20

0.15

0.10

0.05

0

•0.05

-0.10

-0.15

-0.20

-0.25

Distance (km)

Figure 2.11: Mean a solution for BOLIVAR profiles (from top to bottom) 64W, 65W, 67W, and 70W with the corresponding interpretation of the velocity model by Clark et al. (2008), Bezada et al. (2010), Magnani et al. (2009) and Guedez (2007) respectively. S.S.F.Z - Strike Slip Fault Zone; L.I.P - Large Igneous Province; V.I.A -Volcanic Island Arc; E.V.I.A. - Extended Volcanic Island Arc; SCDB - Southern Caribbean Deformed Belt.

Page 82: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

75

solutions in terms of those tectonic provinces and will make some general

observations based on the features that are roughly consistent across the different

profiles.

In the mean a solution for each of the four profiles, mantle under the volcanic

island arc, extended island arc and continent shows consistently slightly higher Ap than

the mantle north of the Southern Caribbean Deformed Belt (Figure 2.11). The transition

between the two blocks occurs over a distance of 50 to 100 km and the magnitude of the

difference in Ap is 0.05 to 0.10 Mg/m3 (Figure 2.10). This pattern seems counterintuitive

from the perspective of the known variations in p and Vp as a function of mantle

depletion. We would expect the mantle under the extinct island arc and the continent to

be more depleted than that under the oceanic plate. Depletion of mantle peridotite

(increase in the Mg#) does not produce significant variation in the P-wace velocity but

produces a systematic decrease in density [Lee, 2003]. Therefore, if composition was the

only contributing factor we would expect that, relative to a reference Vp-p curve,

depleted peridotite would have a lower density for any given Vp, the opposite of what we

observe. This suggests that other factors influence the relationship between Vp and p in

this area. In interpreting this spatial pattern in Ap we must take into account that, in the

seismic models, constraint on upper mantle velocities is limited to the shallowest depths,

given the penetration of the Pn phase. It is therefore possible that variations in velocity

(and density) at depth not captured in the original models are partially responsible for the

Ap pattern we observe.

The sedimentary layer generally exhibits positive Ap in all of the four profiles. A

notable exception occurs in the area directly above the volcanic island arc crust on profile

Page 83: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

76

67W. Here, the sedimentary cover is very thin, and it is possible that this negative density

anomaly results from the influence of the anomaly in the crystalline crust directly below

it.

The distribution of Ap in the igneous-metamorphic crust seems to divide it into

three main provinces: The Caribbean oceanic crust, the Leeward Antilles volcaninc island

arc and extended island arc, and the continental South American crust. The Caribbean

oceanic crust shows very large scatter in the individual solutions for all of the four

profiles, with the mean a solution in most cases showing a weakly negative Ap (Figure

2.10). Given the large standard deviations in the solutions and the relatively weak

constraint on the northern edge of the velocity models, this set of results is inconclusive.

The extinct Leeward Antilles arc shows a generally negative Ap. It is worth

noting that this area of the velocity models is well constrained and that the velocity

structure of this volcanic belt was shown to be consistent laterally along the island chain

[Magnani et al., 2009; Bezada et al, 2010], suggesting similar lithologies throughout the

arc. The negative density anomaly (Ap values of-0.10 to -0.17 Mg/m3) is consistent in all

of the profiles with the exception of 70 W where it becomes very weakly positive near the

center of the profile. On profile 67W the area covering the volcanic island arc has a

stronger anomaly than the extended volcanic island arc crust and the boundaries show

good agreement with the interpretation of the seismic velocity models (Figure 2.11). The

Vp-p relation that results from applying this negative Ap to the Z89 curve differs

significantly from empirical observations on magmatic arc rocks [Behn and Kelemen,

2006], producing smaller densities throughout the relevant Vp interval (Figure 2.12). A

thorough analysis and interpretation of these results in petrological terms is beyond the

Page 84: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

77

scope of this paper, instead, we will briefly explore the possible causes of this anomaly. If

we consider the density and P-wave velocity of some of the mineral components of

basaltic and gabbroic rocks (Figure 2.12), we can speculate that a reduction in density

could be produced by an above average proportion of quartz and feldspars. However, this

would also tend to reduce the seismic velocity, while the Leeward Antilles arc exhibits

relatively high velocities in the models [Magnani et ah, 2009; Bezada et al, 2010]. In

addition to a higher quartz and feldspar content, a shift to the magnesium end member in

the orthopyroxene solid solution series might produce the results we observe, since it

would decrease the proportion of the very dense ferrosilite and increase that of the

relatively light and seismically fast enstatite. Therefore, these results suggest that the

Leeward Antilles crustal rocks contain less iron and perhaps a greater percentage of

quartz and feldspars than the average island arc crust.

4

3.8

3.6

4T 3 4

Q .

3

2.8

2.6

T A

"

"

-

o FSP

MUS 0

5§^

OFER

O H O R ^ ^ ^ *

^ ' '

-

-

-

^ ^ O •

ENS

-

1

6.5 7.5 V (km/s)

p

Figure 2.12: Z89 curve (solid black line) and the same curve shifted in density by -0.15 Mg/m3 (dashed black line) to represent the velocities and densities of the island arc and extended island arc crust as determined by this study. Star represents a reference point on the Z89 curve with Vp=6.5 km/s. Velocities and densities of mineralogical components of basalt and gabbro from Bass (1995). ENS - Enstatite; FER - Ferrosilite; FSP - Feldspar (79% Orthoclase, 19% Albite); HOR - Hornblende; MUS - Muscovite; QZ - Quartz.

Page 85: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

78

In contrast to the Leeward Antilles chain, the South American crust exhibits a

generally positive Ap (0.05 to 0.10 Mg/m3, with the exception of profile 70W, where the

continental crust shows negative Ap values). An interesting result is that the boundary

between the positive Ap values of the South American crust and the negative values of

the volcanic island arc and extended volcanic island arc to the north occurs abruptly and,

for profiles 67W and 64W coincides with the location of the strike slip fault system that

represents the boundary between the Caribbean and South American plates. Based on

differences in Moho depth and velocity structure in the north and south of the strike-slip

fault system in these two profiles, the fault zone was interpreted to be the juxtaposition of

two distinct types of crust [Magnani et al, 2009; Clark et al, 2008]. This analysis of the

residual gravity supports this hypothesis as it shows different Vp-p signatures on the

opposite sides of the margin. On profile 65W, the boundary between the positive and

negative Ap zones occurs further inland, in spatial agreement with a deepening of the

Moho south of the strike slip boundary. Unlike profiles 67W and 64W, profile 65W did

not show significant variation of the velocity structure across the strike slip fault system

[M. J. Bezada et al, 2010]. Again, the analysis of the residual gravity is consistent with

the interpretation of the velocity model since the Ap signature seems to be continuous

across the strike-slip fault system and the step towards slightly positive values occurs -40

km inland. As a final exercise we can take the mean a solution for each profile and add it

to the corresponding seismically derived model to obtain absolute densities (figure 2.13).

In some cases this results in unrealistic absolute densities in parts of the models (e.g. the

upper crust in the center of profile 65 W) meaning that the density adjustments may be too

large. As the noisy synthetic tests showed, the amplitude of the density adjustments are

Page 86: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

79

not always recovered correctly which is why we have focused our analysis on the

polarity, location and lateral extent of the anomalies.

-150 -100 -50

-150 -100 -50

-200 -150 -100 -50

50 100 ISO 200 250 300 Distance (km)

50 100 150 200 250 300

-150 -100 -SO 0

N

50 100 150 200 250 300 350 400 Distance (km)

-250 -200 -150 -100 -50 50 100 150 200

-200 -150 -100 -50 0 50 100 150 200 250 300 Distance (km)

-250 -200 -150 -100 -50 0 50 100 150 200 Distance (km)

2.372 2.527 2.682

- 1 1 1 M i l l . . . 1.886 2.127 2.279 2.431 2.592 2.800 2.946 3.115

Density (Mg/mJ)

Figure 2.13: Absolute density models for four boundary-normal profiles obtained by applying the density adjustments from the mean a solutions to the corresponding seismically derived models. 64W (a), 65W (b), 67W (c) and 70W (d). The density models were derived from P-wave seismic velocity models using the Z89 curve. Observed gravity anomaly shown as black crosses, calculated anomaly as red line. RMS misfits are indicated in the figure and differ slightly from those shown in figure 2.10 due to differences in the parameterization.

2.5. Discussion and conclusions

Seismically derived density models should be regarded as initial approximations

to the subsurface density structure as they often produce significant levels of misfit with

Page 87: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

80

respect to gravity anomaly observations. The residual gravity (not explained by the

seismically-derived model) can be viewed as resulting from a distribution of density

anomalies or adjustments (Ap) in the subsurface. Here, we have presented a new method

for the analysis of residual gravity curves. The method uses the Genetic Algorithm

paradigm in combination with a linear inversion technique to find a simple distribution of

density adjustments that will minimize the gravity anomaly misfit. The multi-realization

nature of the algorithm allows it to provide a range of different solutions that fit the data

at a similar level, which helps to assess the degree of non-uniqueness of the solution at

different locations in the model. After multiple realizations have been completed, the

mean and standard deviation of all the individual solutions can be taken as an estimate of

the A/? distribution in the subsurface and the associated uncertainty.

Tests showed excellent results when the noise-free synthetic residual gravity

calculated from a known model was used as input to the algorithm. In those tests, the

lateral boundaries as well as the magnitude of all the prescribed anomalies were

recovered accurately. Additionally, the algorithm proved to be robust with respect to the

number of provinces per layer used, achieving better results when the number exceeded

that of the true model. When noise was added to the synthetic residual gravity to simulate

a scenario where the model parameterization was inadequate, artifacts were observed in

the mean solutions and there was not a major improvement associated with increasing the

number of provinces per layer beyond one more than in the true model. However, even in

this sub-optimal scenario, the main features of the true model were recovered.

Specifically, the boundaries were recovered more accurately than the magnitude of the

anomalies and the polarities of the anomalies were always recovered accurately.

Page 88: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

81

To test the algorithm on real data, we applied it to velocity models and gravity

data from the Caribbean-South American plate boundary. For all of the four profiles, a

significant amount of scatter was seen in the sets of individual solutions. However, some

consistent trends were observed in the mean solutions for each profile. Despite the

remaining ambiguity, having multiple profiles across the same tectonic units allows us to

make the following general observations: The sedimentary layer showed generally

positive Ap values, reflecting the fact that the Vp-p curve used to obtain the initial model

may underestimate the density of sedimentary rocks. The lithospheric mantle showed a

transition from higher Ap values in the south to lower values in the north occurring near

the Southern Caribbean Deformed Belt near the northern end of the profiles. This

observation is difficult to explain in terms of mantle depletion, suggesting that other

factors must play an important role. The crust of the island arc and extended island arc

parts of the profiles generally show negative Ap values suggesting that the Leeward

Antilles have lower iron content and possibly a higher proportion of quartz and feldspars

than average island arc crust. Continental South American crust showed generally

positive Ap values. The transition between this positive Ap regime and the negative Ap

regime to the north occurs abruptly and coincides with the strike slip plate margin for

profiles 64W and 67W, supporting the interpretation of the velocity models that places

different types of crust on either side of the fault zone [Magnani et al, 2009; Clark et al,

2008]. For profile 65W, the step to positive Ap values occurs inland from the plate

margin, also in accordance with the interpretation of the seismic velocity model [M. J.

Bezada et al., 2010].

Page 89: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

82

The application we have presented serves as an example of the usefulness of our

algorithm in the analysis of residual gravity data. Due to the reliance of the algorithm on

the layer boundary depths defined by the velocity model, the robustness of the results will

improve with increasing constraint on the velocity models. Given the inherent non-

uniqueness of gravity problems, we consider the ability of the algorithm to yield a range

of different solutions its most important feature. In the application we present here we

have used the mean of the set of individual realization solutions as an estimate of the

subsurface distribution of Ap. This approach proved to be successful in the synthetic tests

we carried out, but different alternatives are also valid. For example, an interpreter may

choose to deviate from the mean solution when warranted by additional geologic

information. The advantage of having a set of possible solutions is that they provide an

idea of the acceptable model space from which to proceed. We acknowledge that many

simplifications were made in order to make the algorithm work, but consider it a step in

the right direction, and an improvement over the reliance on forward modeling or single-

layer inversion schemes.

Page 90: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

83

Chapter 3

Subduction in the Southern Caribbean: Images from finite-frequency P-

wave tomography.

M. J. Bezada, A. Levander, B. Schmandt

Abstract

The eastern boundary of the Caribbean plate is marked by subduction of the Atlantic

under the Lesser Antilles. The southeastern plate boundary is characterized by a strike-

slip margin while different configurations of subduction of the southwest Caribbean

under South America have been proposed. We investigate the slab geometry in the upper

mantle using multiple-frequency, teleseismic P-wave tomography. Waveforms from P

and PKIKP phases from 285 (Mb > 5.0) events occurring at epicentral distances from 30°

to 90° and greater than 150° were bandpass filtered and cross-correlated to obtain up to

three sets of delay times for each event. The delay times were inverted using approximate

first Fresnel zone sensitivity kernels. Our results show the subducting Atlantic slab, as

well as a second slab in the west of the study area that we interpret as a subducting

fragment of the Caribbean plate. Both slabs have steep dips where imaged and can be

traced to depths greater than 600 km. These results are consistent with transition zone

boundary topography as determined by receiver function analysis. The Atlantic slab

extends continent-ward south of the plate bounding strike-slip margin. We interpret this

extension as continental margin lithospheric mantle that is detaching from beneath South

America and subducting along with the oceanic Atlantic slab. The steep subduction of the

Caribbean occurs ~500 km landward from the trench, implying an initial stage of shallow

Page 91: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

84

subduction as far to the east as the Lake Maracaibo - Merida Andes region, as has been

inferred from intermediate depth seismicity.

3.1. Introduction

The Caribbean is a relatively small plate that originated in the Pacific, and now

lies between the North and South American plates. It has a complicated tectonic history

that includes episodes of voluminous magmatism in the Cretaceous that generated the

Caribbean Large Igneous Province [Coffin and Eldholm, 1994], a subduction polarity

reversal [e.g. Burke, 1988; Pindell and Dewey, 1982], clockwise rotation after collision

with North America at -55-60 Ma [Pindell and Kennan, 2007] and highly oblique

collision with South America for the past ~55 Ma [Dewey and Pindell, 1986;

Rosencrantz et al, 1988].

Motion of the Caribbean plate relative to a fixed South America has a magnitude of ~20

mm/yr in the southern Caribbean. The orientation of relative motion is E-W offshore

eastern Venezuela while in the western Caribbean there is a small but significant

(~5mm/yr at San Andres island, Figure 3.1) southward component. As a result, NW-SE

compression between the two plates occurs offshore the Colombian coast [Calais and

Mann, 2009] (Figure 3.1). The nature of the Caribbean-South America plate boundary

varies spatially according to the angle between the boundary and relative motion: The

eastern plate boundary is defined by the Lesser Antilles subduction zone, the boundary

between the Southeastern Caribbean and South America is characterized by strike-slip

motion along the San Sebastian-El Pilar fault system, and in the southwest Caribbean

Page 92: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

85

-90° -85° -80° -75° -70° -65° -60°

Figure 3.1: Tectonic setting of our study area (top): Red arrows show GPS vectors from a compilation by Calais and P. Mann [2009], plate boundaries from the UTIG plates database. SA - San Andres island; LA - Leeward Antilles; LM - Lake Maracaibo; SS-EP - San Sebastian-El Pilar fault system; BF - Bocono fault system; BB - Barcelona Bay; OR - Orinoco River. Red asterisk shows location of Kick'em Jenny volcano. Blue rectangle shows the area of our focus and location of the bottom panel. Station distribution and seismicity (bottom): Symbols representing seismic stations and earthquake epicenters as described in the figure. BN - Bucaramanga nest; PC - Paria cluster; MA - Merida Andes; PR - Perija Range; SM - Santa Marta Massif; SF -Serrania de Falcon. Seismicity from the FUNVISIS catalog.

Page 93: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

86

different configurations of shallow subduction of the Caribbean plate under South

America have been proposed [e.g. van der Hilst and Mann, 1994; Kellogg and Bonini,

1982] (Figure 3.1)

The Atlantic (oceanic extension of the North and South American plates)

subducts under the Caribbean at the Lesser Antilles island arc. The southernmost

historically active volcano on the arc is the Kick'em Jenny, an underwater volcano

located 8 km offshore northern Grenada and ~240 km from the Venezuelan coastline

along the arc trend [Watlington et al, 2002] (Figure 3.1). The Benioff zone defined by

intermediate depth seismicity along the arc reaches a depth of -200 km, and its southern

terminus is the Paria cluster of seismicity (Figure 3.1). This cluster has been interpreted

as the site of lithospheric tearing [Clark et al, 2008], and as defining a steeply subducting

slab [Russo et al, 1993]. VanDecar et al [2003], using non-linear teleseismic

tomography, imaged a landward continuation of the Atlantic slab onshore South

America. Based on their images, they proposed a model for shear tearing of subducted

passive margin lithosphere from the stable South American continental lithosphere. The

landward continuation of the subducting slab, beyond the island arc had previously been

inferred based on the outcropping of Plio-Pleistocene intrusive rocks onshore Venezuela

[McMahon, 2001; Santamaria and Schubert, 1974] and the location of hot springs that

may be associated with cooling of a shallow intrusive body [Urbani, 1989].

In a relatively small region at the southeast corner of the Caribbean, there is a

transition from subduction to transform motion that has been interpreted as a Slab-

Transform Edge Propagator (STEP) fault [Govers and Wortel, 2005]. Offshore eastern

and central Venezuela, the plate boundary is clearly defined at the dextral San Sebastian-

Page 94: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

87

El Pilar strike-slip fault system, with -80% of the plate motion in eastern Venezuela

being accommodated in an 80 km wide shear zone centered on those faults [Omar J

Perez et al, 2001]. Further west, the plate boundary is more diffuse and is complicated

by the presence of the Southern Caribbean Deformed Belt that extends from offshore

western Colombia to offshore eastern Venezuela and the northeast trending Bocono

strike-slip fault system (Figure 3.1). Gravity and seismic reflection studies have

indicated large accumulations of sediments along the Southern Caribbean Deformed Belt,

and the sedimentary sequence shows significant compressive deformation and faulting

[Case et al, 1990; Magnani et al, 2009; Talwani et al, 1977]. The Southern Caribbean

Deformed Belt has been interpreted as the boundary between the Caribbean and South

American plates and the site of Caribbean subduction under South America [Kellogg and

Bonini, 1982]. Under northwestern Colombia, several authors have used intermediate

depth seismicity to outline a subducting fragment of the Caribbean plate. This subduction

is thought to have a shallow dip of 20° to 30° towards the ESE [Malave and Suarez,

1995; Ojeda and Havskov, 2001; Pennington, 1981] or SE [Kellogg and Bonini, 1982]

with the subduction angle possibly steepening at depths greater than 150 km [Pennington,

1981]. This shallow subduction of the Caribbean plate is thought to be the driving force

behind Laramide-style deformation and uplift in the overriding South American plate,

specifically in the Santa Marta Massif, Perija Ranges and Merida Andes[Kellogg and

Bonini, 1982] possibly forming an orogenic float [Audemard and Audemard, 2002]. A

different interpretation of Caribbean subduction was put forth by van der Hilst and Mann

[1994] on the basis of seismic tomography images. They proposed a shallower

subduction of 17° toward the SSE with the Caribbean subducting beneath the northern

Page 95: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

88

Columbia and northwestern Venezuela to a depth of 225 km. Thus, there are two views

of subduction of the Caribbean under northwestern South America; one view calls for

shallow subduction from the west, the other calls for shallow subduction from the north.

In this paper, we investigate the slab geometry in the upper mantle using finite-

frequency P-wave tomography. The data consist of teleseismic delays obtained mainly

from the BOLIVAR (Broadband Ocean-Land Investigation of Venezuela and the Antilles

arc Region [Levander et al, 2006]) passive seismic array and supplemented with data

collected more recently in the west of the region. The study area covers the southeast

Caribbean margin and northernmost South America from the Guayana Shield in the south

to the Leeward Antilles in the north and from eastern Venezuela to the Venezuela -

Colombia border in the west. Our results show the subducting Atlantic slab, as well as a

second slab in the west of the study area that we interpret as a subducting fragment of the

Caribbean plate. Both slabs have steep dips where imaged and can be traced to depths

greater than 600 km. These results are consistent with transition zone boundary

topography as determined by receiver function analysis.

3.2. Data

The data were collected by the BOLIVAR seismic array which included the 39

stations of the Venezuelan National Seismological Network as well as a temporary (~17

month) deployment of 35 IRIS-PASSCAL, 8 Rice broadband stations and data from a

year long deployment of 11 OB SIP ocean bottom seismographs (Figure 3.1). The array

covers the eastern part of Venezuela with typical station spacing of ~50-100 km while

coverage is sparser in the west of the study area. There, we rely on the stations of the

Page 96: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

89

Venezuelan permanent network and a later deployment of 8 PASSCAL and Rice

instruments along a north-south profile at ~70°W. The array covers an area of -900,000

km2 from south of the Orinoco river at latitude of ~6°N, along the northern Guayana

Shield, to the Caribbean Sea at latitudes of up to 14°N; and from eastern Venezuela at a

longitude of 61.5°W to the Perija range in western Venezuela at a longitude of 72°W

(Figure 3.1).

Events with Mb > 5.5 and epicentral distances between 30-90° and 150-180°

were initially examined for this study. To fill under-sampled backazimuths, smaller (5 <

Mb < 5.5) events were examined, and those with consistent waveforms across the array

were selected and included in the analysis. Delay times from P and PKIKP phases from

285 events were included in the final inversion. The event distribution provided good

azimuthal coverage and there is a notable abundance of near-antipodal events from the

Sumatra-Andaman region (Figure 3.2). Waveforms from the selected events were

bandpass filtered at three narrow frequency bands centered on 1.0, 0.5 and 0.3 Hz.

Traveltime delays were obtained by cross-correlation of the filtered waveforms so that up

to three sets of delays were obtained for each event, bringing the final number of delay

times used in the inversion to 7273 (2484, 2453 and 2436 delay times for the 1.0 Hz, 0.5

Hz and 0.3 Hz bands; respectively).

3.3. Method

The multiple-frequency tomography we use differs from purely ray-theoretical

tomography in that it approximately accounts for shifts in arrival times caused by

velocity perturbations located off the geometrical ray path. This approach allows more

Page 97: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

Figu

re 3

.2: L

ocat

ion

of e

vent

s us

ed fo

r th

is s

tudy

in p

olar

pro

ject

ion

cent

ered

on

the

arra

y (le

ft) a

nd o

n th

e ar

ray

antip

ode

(rig

ht).

The

ep

icen

tral d

ista

nce

from

the

arra

y is

labe

led

on th

e fig

ure.

Das

hed

lines

repr

esen

t pla

te b

ound

arie

s.

O

Page 98: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

91

thorough use of high-quality broadband seismograms and imposes physically based

smoothness criteria on the inversion thereby reducing the need for ad hoc regularization

of the inverse problem. In this study, we use the method of Schmandt and Humphreys

[2010] that is based on the single scattering "Banana-Doughnut" formulation of Dahlen

et al. [2000] but only considers the sensitivity within the first Fresnel zone and applies

ray-normal smoothing to address ray-location uncertainty.

The model is defined as an irregular rectangular mesh. Cell size is variable to

accommodate the loss of resolution with model depth as well as towards the edges of the

array. Horizontal mesh node spacing is smallest inside the array footprint, where it is

42km, and increases stepwise until it reaches 56 km 400 km outside the array. Cell height

is also variable from 24km for the shallowest layer of cells (36 to 60 km depth) to 60 km

for the deepest layer (685 to 745 km depth).

To avoid mapping the effect of crustal velocity structure into the mantle, we apply

a crustal correction to the delay times. Ray-theoretical correction times are calculated

using a crustal model derived from 2D active-source P-wave velocity models, 3D

receiver function studies and 3D gravity data. The model consists of linear velocity

gradients from the surface to the igneous-metamorphic basement, from the basement to

the Moho discontinuity and from the Moho to a depth of 50 km. Basement depths were

estimated using gravity anomaly data [Izarra et al, 2005] and controlled with active

source seismic profiles [Christeson et al, 2008; Clark et al, 2008; Guedez, 2007;

Magnani et al, 2009; Schmitz et al, 2002; Bezada et al, 2010] and published estimates

of basement depth based on potential field data [Feo-Codecido et al, 1984]. Moho depth

was estimated by interpolating the depths determined by active source seismic profiles

Page 99: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

92

and selected Moho depths obtained through a receiver function study [ Niu et al, 2007;

Bezada et al, 2006].

In addition to model slownesses, the inversion solves for station and event terms.

The purpose of the station terms is to absorb the effect of shallow velocity anomalies

improperly accounted for by the crustal model. In order to prevent it from also absorbing

the effect of our target mantle heterogeneities we introduce a damping parameter. The

event terms aim to remove the dependence of an event's mean arrival time on the set of

stations that recorded it. We regularize the inversion by including model norm damping

as well as smoothing operators. Model norm damping drives the inversion towards the

minimum structure solution, minimizing the occurrence of anomalies not strongly

required by the data. The smoothing operator we apply is depth dependent, so that it is

strictest in the plane normal to the mean ray path at each depth.

The cost function we minimize is then [Schmandt and Humphreys, 2010]:

» . ,112 , | | . ||2 , II i|2

Am-d| | +&1|JLm|| +fc2 m

Where A is the partial derivative matrix that relates the model parameters m

(velocity perturbations in each voxel, event and station terms) to the observed delay times

d, L is the matrix that defines the depth dependent smoothing operators, and k] and fo are

free parameters defining the relative weight of the smoothness and norm damping

constraints, respectively. A solution m that minimizes c is found using the iterative

LSQR method [Paige and Saunders, 1982].

The selection of the values of the free parameters is somewhat subjective, but was

guided by the well-known L-curve that shows the trade-off between model norm and

variance reduction with varying damping coefficients (ki). The addition of the

Page 100: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

93

smoothness constraint means that an L-curve can be obtained for each smoothness

parameter (kj).

Figure 3.3 shows three of these curves that behave as expected (more smoothing

results in slightly larger model norms and smaller variance reductions). We chose a

moderate value of ki and a value of fe near the turning point of the corresponding L-

curve. Different values of these free parameters did not significantly modify the shape or

amplitude of the important and well resolved anomalies in the model (Figure 3.4)

We use the AK135 [Kennett et al, 1995] ID velocity model as our reference

model and to calculate the ray paths and sensitivity kernels.

84 88 Variance Reduction (%)

Figure 3.3: Trade-off between variance reduction and model norm for fo values ranging from 2 to 24 at an interval of 2 and different ki values as indicated in the figure. The larger symbol corresponds to the values used in the inversion: £/=15 and k2=\2. See text for a description of the ki and fo parameters.

Page 101: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

Smoothing Increases

k, = 5 k =15 k, = 25

k, = 2

*,= 12

O

3 •D. 3

IQ 5" n 3 EU

fc = 2 4

Vp Perturbation (%)

- 5 - 4 - 3 - 2 - 1 0 1 2 3 4 5

Figure 3.4: Effect of free parameters on the model. Model slices at 95 km depth. The preferred model is shown in the center panel.

Page 102: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

Smoothing Increases

k, = 5 k,= 15 k = 25

k=l

k = 12

O 0) 3 -q 3 -

5" n

a> in 0)

*, = 24

Vp Perturbation (%)

-S -4 -3 -2 -1 0 1 2 3 4 5

Figure 3.4 (continued): Effect of free parameters on the model. Model slices at 375 km depth. The preferred model is shown in the center panel.

Page 103: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

Smoothing Increases

k, = 5 k,= \S k, = 25

k=2

k = 12

D 3

• g 3 ' m 5" n - n> CD

it, = 24

Hip Ha£

Vp Perturbation (%)

- 5 - 4 - 3 - 2 - 1 0 1 2 3 4 5

Figure 3.4 (continued): Effect of free parameters on the model. Model slices at 685 km depth. The preferred model is shown in the center panel.

Page 104: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

97

3.4. Model

Here we describe the location, shape and amplitude of the significant velocity

anomalies we observe, as well as their correlation with tectonic features. At lithospheric

depths, we observe laterally small (-100 km in diameter) and spatially consistent

anomalies. Deeper in the mantle we observe the two most striking features of the velocity

model: Large, steeply dipping high velocity anomalies (2.5 to 4%) that we interpret as

subducted Atlantic and Caribbean slabs.

In the shallowest 200 km of the model, we find a clear pattern of low velocity

anomalies under the Caribbean plate, except under the islands of the Leeward Antilles

which show very shallow (upper 100 km) neutral to positive (1%) velocity anomalies

(Figure 3.5). North-South oriented shallow low velocity anomalies (-1.5 to -2.5%) also

characterize the area directly east of Lake Maracaibo, east of the Merida Andes, and

beneath the uplifted Serrania de Falcon. This anomaly extends to -200 km depth. The

most prominent feature in western Venezuela is a high velocity anomaly (up to -2.5%) in

the very northwest of the study area, near the Perija range (Figure 3.5) that is

concentrated beneath the westernmost station in our array but spreads out beneath Lake

Maracaibo with depth.

A negative anomaly is observed in central Venezuela south of the Barcelona Bay

extending as far south as the Orinoco River. South of the Orinoco and east of 64°W the

Guayana Shield is characterized by a small (-1%) positive velocity anomaly that

decreases in lateral extent with depth and reaches depths of-245 km (Figure 3.5). The

western part of the shield (south of the Orinoco and longitudes 68° W to 64°W) shows no

significant velocity anomalies. Given the distribution of stations along the northern edge

Page 105: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

P-V

elo

city

an

om

aly

, D

ep

th =

60

P-V

elo

city

an

om

aly

, D

ep

th =

95

14

12

Latitude 6 4

P--V

eloc

ity a

no

ma

ly,

Dep

th =

130

0 &

\ ^

-Wv

j^^

V

-74

-72

-70

-68

-66

-64

-62

-60

Lo

ng

itud

e

14

12

•§1

0

3 8 6 4

P-V

elo

city

an

om

aly

, D

epth

= 1

65

0

—?

*\

V

V"""

'' /

'

-74

-72

-70

-68

-66

-64

-62

-60

Long

itude

P-V

elo

city

an

om

aly

, D

ep

th =

205

-74

-72

-70

-68

-66

-64

-62

-60

Lo

ng

itud

e

P-V

elo

city

an

om

aly

, D

epth

= 2

45

14

12

Latitude 6 4

jp-

Av

/'A

K •''.•

"""!

..

. J

j

"VV

J^Jp

i

0 0

fc °

R

"*

,T

/"^

^ }'

-74

-72

-70

-68

-66

-64

-62

-60

Lo

ng

itud

e -7

4 -7

2 -7

0 -6

8 -6

6 -6

4 -6

2 -6

0 Lo

ngitu

de

Vp

Per

turb

atio

n (%

)

£-1

.5

-0.5

0.

5

-74

-72

-70

-68

-66

-64

-62

-60

Lo

ng

itud

e

>1.

5

Figu

re 3

.5: V

eloc

ity a

nom

alie

s in

the

shal

low

est 2

45 k

m o

f the

mod

el. T

he c

olor

sca

le is

sat

urat

ed a

t ±1.

5% to

hig

hlig

ht th

e G

uaya

na

Shie

ld a

nom

aly

Page 106: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

99

of the shield, the anomaly pattern we recover suggests that the northern limit of the

cratonic root under the eastern part of the shield reaches at least as far north as the

Orinoco River, and possibly ~100 km further. In contrast, in the western part of the

shield, the cratonic root does not seem to extend as far north as the Orinoco River. The

pattern of shallow mantle anomalies we present, and particularly the lack of a high

velocity anomaly under the northern edge of the western Guayana Shield, is consistent

with a recent S-wave velocity model derived from Rayleigh wave tomography [Miller et

al, 2009]. The latter model lacks the low velocity anomalies directly east of Lake

Maracaibo, but suffers from poor resolution in that part of the study area. All of the

anomalies described above become insignificant at depths greater than -245 km.

The two main features of the model, and the ones we will focus on for the

remainder of the paper, are vertically continuous positive anomalies located in the east

and west of the study area. In the east, we image the subducting Atlantic slab while in the

west we image a steeply dipping slab that we interpret as a subducted fragment of the

southernmost Caribbean plate (Figure 3.6). The Atlantic slab is visible from shallow

depths at the eastern end of the study area and at 95 km depth follows the trend of the

Lesser Antilles Arc. The northern edge of the anomaly is 100 km SSW of the island of

Grenada and the southern end is located onshore, 110 km southwest of the El Pilar fault.

The Atlantic slab anomaly widens with depth from -95 km at 115 km depth to -150 km

at 460 km depth, possibly as a result of decreased resolution. At depths greater than 285

km, the anomaly has a distinct "L" shape, with one arm trending Northeast-Southwest

along the subducting Atlantic plate and the other trending East-West, parallel to the plate

boundary faults. The anomaly dips towards the west at an angle of-70°. The magnitude

Page 107: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

P-Velocity anomaly, Depth = 95 P-Velocity anomaly, Depth = 205 P-Velocity anomaly, Depth = 330

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

P-Velocity anomaly, Depth = 470 P-Velocity anomaly. Depth =630 P-Velocity anomaly, Depth = 745

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

100

200

J300

•£400 o. <u Q500

600

700

A ^

!

V-i

h

-

-

A —^__, ^

i

I ' - - - - - j

ATL |

*~." i . ~7~~T~—r—-. 1.

100

200

J 300

•£400 o. K Q500

600

700

B B'

u—— ~ ~ :

I f

j r

100

200

| 3 0 0

£400 a. aj Q500

600

700

1

i

! CAR

|

\ \. i

"~ — - j

-800 -600 -400 -200 0 200 400 600 800 Distance (km)

-800-600-400-200 0 200 400 600 800 Distance (km)

Vp Perturbation (%)

-800 -600 -400 -200 0 200 400 600 800 Distance (km)

100

300

400

500

600

700

\ — x^

-

-

-

^

\

1

- 5 - 4 - 3 - 2 - 1 0 1 2 3 4 5

- __ *^

' - - • ; _ • .-- -_,-., -«? -_- " _ _ _ ^ ^

~ \ \ — " ~Rr\ ~ ~ " '' ^ "\ \ A 4 • i

v M ' i 1 ' fCARjO / *

\ j ^^\ ATL 1 1 1 1 1 1 r~~ -—T^ .

-74 -72 -70 -68 -66 Longitude

-64 -62 -60

Figure 3.6: Horizontal slices through the model at depths indicated in the figure (top two rows), vertical slices through the model at location indicated in top right panel (third row), and 3D image of the subducted slabs (1.5% isoanomaly surface) viewed from the south. ATL - Atlantic slab; CAR - Caribbean Slab.

Page 108: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

101

of the anomaly reaches 3.2 to 4.2% at its core with the E-W arm of the anomaly

consistently showing amplitudes - 1 % smaller than the NE-SW arm.

The second important positive velocity perturbation occurs in the west of the

study area. It becomes well resolved at a depth of -200 km and appears connected to the

shallower anomaly surfacing beneath the Perija range. This feature dips steeply towards

the ESE and is clearly visible to a depth of over 600 km, though it weakens and vanishes

by a depth of 700 km, most likely due to lack of resolution as it is near one side of our

seismograph array. The anomaly has an approximately SSW-ENE trend, but its shape is

not very well defined given the poor station coverage in the area. This Caribbean slab

anomaly has smaller peak amplitude, compared to the Atlantic slab, of 2% to 2.7%.

3.5. Model Assessment

To gain insight into the resolution of our model, we performed a number of

synthetic tests in which we prescribed a velocity anomaly corresponding to a predefined

slab geometry. Synthetic delay times calculated from these models were used as input to

our inversion scheme and the recovered model was used to evaluate the success of the

inversion. The object of these tests is to assess the robustness of the geometry of the

features in our model, specifically their lateral extent and the maximum depth.

In our model, at a depth of 95 km, the northern edge of the Atlantic slab occurs

-100 km south of the island of Grenada, coinciding with the northern limit of our array.

Obviously, we would expect the slab to continue northward under the island chain. To

verify that this boundary results from the array aperture, we carried out a synthetic test

where the input model had the geometry of the Atlantic slab. The model contained a

Page 109: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

102

westward dipping (70°) +4% velocity anomaly following the trend of the island arc and

continuing landward under the continent. Synthetic delay times through the model were

calculated using our receiver geometry and, after adding random noise with a standard

deviation of 200 ms, these times were inverted using the same free parameters as in our

final inversion. The result shows that, at a depth of 95 km, the positive anomaly is only

recovered as far north as 11.75°N, consistent with our image (Figure 3.7). We performed

a similar test to investigate if horizontal smearing could be responsible for the landward

continuation of the Atlantic slab. In this case, the positive anomaly in the prescribed

model only extended as far south as the strike slip fault system (10.5°N). After adding

noise and inverting the synthetic delay times, we recovered the southern edge of the

anomaly at its prescribed location. No significant horizontal smearing of the anomaly was

observed (Figure 3.7).

The southwestern end of the study area is the site of one of the more interesting

features of the model, yet it is sparsely covered by stations. Given the sparse station

coverage, we ran a synthetic test in order to explore how well we could expect to recover

the geometry of a large anomaly in this part of the study area. Outside of our station

coverage, the input model for this test was based on the slab depths inferred from

seismicity by Malave and Suarez [1995] and was defined as follows: The 3% positive

anomaly dips towards the ESE and the dip increases stepwise from 20° at the trench

offshore western Colombia to 30° and finally 80° at depths greater than 245 km, under

the western edge of our array. The width of the synthetic subducting slab extends farther

to the northeast and southwest than the anomaly recovered in our model. We find that the

inversion successfully recovers the northeastern extension of the prescribed slab while its

Page 110: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

Synthetic Input Depth = 95 Synthetic Result, Depth = 95 Velocity Model, Depth = 95

14

12

•a 10

3 8

6

4

j p ' A > *

J ( y v

1 • *

\

0

• * • * • •

• * * h • • p

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

Synthetic Input, Depth = 285

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

Synthetic Result, Depth = 285

14

12

•8 10

Latit

00

6

4

0 • « a

^ P ' a * " . . ' . : • »

/Vw. V T * - ! ^ ^ ? i * * •• * •/.•-• y^

""\ • " i*""

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

Velocity Model, Depth = 285

14

12

•S 10

3 8

6

4

a . ' »

s?-*cr . . . : * • • ~t / A - ' " * i \ * • •» *?' '

\ • • • • • • • • • i _ 1 • • : • • • • . . . ~ > s

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

Synthetic Input Depth = 95

14

12

| 1 0

3 8

6

4

t

L* ' ::•'•••: f* \ • •r - '

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

Synthetic Result, Depth = 95

14

12

-8 10

3 8

6

4

0 • i

sy « v ..'.:•'

V ^ v ? T ^ ^ \ • • ! • • • • y^v

1 N.

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

Velocity Model, Depth = 95

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

Synthetic Input Depth = 285

14

12

-8 io

Latit

00

6

4

yy Q. v>

/ A' ••• 1 ' *

0

- . . • " * " • _ A * "

-TsiA^s '•" * *.".••• T3^

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

Synthetic Result, Depth = 285

14

12

•8 io La

tit

00

6

4

0 • i

/ V W / V > l A 3 M J

s . • • • . c

\ Y"xr"'/ -74 -72 -70 -68 -66 -64 -62 -60

Longitude

Velocity Model, Depth = 285

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

Pi

14

12

•8 10

3 8

6

4

j p - i v ,

f'i\':'" S . •

0 0 *

• • . . " • "

-*7vlA^S • * • * * «• ?

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

Vp Perturbation (%)

14

12

•8 10 3

3 8 6

4

yy ^v .

I A'-'*

A — - * _ •

0

• •••"* 8

• • _* • •* T>i-^X? • * • * • . • ?

• *5

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

-S -4 -3 -2 -1 0 1 2 3 4 5

Figure 3.7: synthetic tests of northward (top two rows) and southward (bottom two rows) extension of the Atlantic slab and comparison with the velocity model. The northern limit of the recovered anomaly is determined by the station distribution. Southward extension of the slab is recovered accurately, no significant horizontal smearing is observed.

Page 111: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

104

southwestern extension seems to be determined by station coverage; as the prescribed

anomaly was not adequately recovered beyond the southern limit of the anomaly in our

model (Figure 3.7).

We test the prospect of vertical smearing on our velocity anomalies by attempting

to recover synthetic anomalies meant to replicate the subducting slabs in our model but

truncated at a depth of 450 km. After adding noise and inverting the delay times we

obtain models that resolve the maximum depth of the anomaly to within ~50 km for both

anomalies, suggesting that significant vertical smearing of the anomalies is not likely

(Figure 3.9).

Finally, the conspicuous absence of a Caribbean slab shallowly dipping

southward beneath northwestern Venezuela prompted us to conduct tests to see if such a

feature could be resolved with our data set and method. In this case, the input model was

meant to replicate the geometry of the Caribbean slab as interpreted by van der Hilst and

Mann [1994]; dipping to the SSE at an angle of 17° and reaching a depth of 250 km, with

a velocity anomaly of +3%. The inversion of the synthetic delay times with added noise

does not adequately recover the prescribed anomaly. Instead, we see a pattern of small

positive anomalies in the area above the prescribed slab and small negative anomalies to

the east of the slab (Figure 3.10). The shallow dip of the slab may be the main reason

why our experiment geometry seems to be insensitive to such a feature.

It is worth noting that resolution tests such as the ones we have conducted are

inherently optimistic. The formulation used in the inversion process assumes the

subsurface consists of elastic, isotropic media, and that ray paths are accurately known.

Because the sensitivity kernels are used to produce the synthetic delay times, these

Page 112: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

Synthetic Input, Depth = 60 Synthetic Result, Depth = 60 Velocity Model, Depth = 285

14

12

3 10

Lat

it

00

6

4

0

' vp* aj"" . . * . . " • •

7 (\ ••• • ,•r— r« ^*c

\ • * f" ? r» \ Y ^ ^ y

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

Synthetic Input, Depth = 285

14

12

•S 10

Lat

it

CO

6

4

JO* i V ,

l'(I'1* \ '. '

-\*

0

•. . • • • • "

• • • • ?

' • • : * ' • • • : ^

" " J

-74 -72 -70 -68 - 6 6 -64 -62 - 6 0 Longitude

Synthetic Input, Depth = 375

14

12

•8 10

Lat

it

00

6

4

yy A_v>

MH-l i * "

0

• • • «

• * \

14

12

•S 10 3

S 8

6

4

» * ' / ; ; « . . * . ; . .

_n—' J \ [ - - J ^ ' » 'TN • • • "

AT.' ^ V ^ K I \_* ' :"!'•'•••: £ •

\ • * f* ? Vi \ K i '

-74 -72 -70 -68 -66 - 6 4 -62 -60 Longitude

Synthetic Result, Depth = 285

14

12

•S 10 3 a 3 8

6

4

/ > • i n

V • * 1 *

\

0

. . . : • » • • * * " '-T^iArMI .

• • * u

* *; ? V»

-74 -72 - 7 0 -68 - 6 6 -64 -62 -60 Longitude

Synthetic Result, Depth = 375

14

12

4jio .« S 8

6

4

0

> P ' ft_V . . * . . " • •

< 0 * :•. •••.'^ 1 • * "•• • *.".*•• 7^

14

12

•S 10 3

, t t

3 8

6

4

0 * o

/ Y i v , ' • . . ~!~~y^/J\*^' WS • • • » * "

AT'- ^-<L%"?ri3 4, \ J * •*• "* . • ? \ ^ * • • • • • • • 1 _ > • ' •••• •• ••• /^

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

Velocity Model, Depth = 285

14

12

•3 io

3 8

6

4

0

. ° & ^ P " eLv ' ..".."•• r A T * *^—V-ATCT .

4 V " • * • • • •*. 1 1 " * • • • • * • • J^^-\ • • * • C

\ * * c

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

Velocity Model, Depth = 375

14

12

•8 10 3 .e 5 8

6

4

JO* « X *

/0'* ^ • •

—-*-*•

0

• • « * " -^fe

a • • \_

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

-74 -72 -70 -68 - 6 6 -64 -62 -60 Longitude

-74 -72 -70 -68 - 6 6 -64 -62 -60 Longitude

Vp Perturbation (%)

Figure 3.8: Synthetic test of lateral extension of the Caribbean slab and comparison with the velocity model. The northeastward extension of the anomaly is well resolved but resolution is lost to the southwest of Lake Maracaibo.

Page 113: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

Synthetic Input, Depth = 285 Synthetic Result, Depth = 285 Velocity Model, Depth = 285

14

12

•S 10

atit

6

4

A

3

®^^W' S,!S;A'-- v.-.•:.. ^ K . • * ' • C

V . '{*

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

100

200

I 30° &400 O

500

600

700

I i

!

1 . . . . :

f -! I

i - - i -300 -200 -100 0 100 200 300

Distance (km)

-200 -100 0 100 200 Distance (km)

14

12

10

8

6

4

A

0

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

A' 100

200

J 300

& 400 O

500

600

700

| 1

1 1 1 i 1

1 i 1

1 1

i

1 -300 -200 -100 0 100 200 300

Distance (km)

100

200

I 300

& 400 o

500

600

700

B B

~

! !

L f

1 ! i

-200 -100 0 100 200 Distance (km)

Vp Perturbation (%)

14

12

-S io

atit

6

4

A ,

0

• * 6

\iv ;•; ••••. s 1 « ^ * r t •- • • • . "^^N

v — * " " * 5

-74 -72 -70 -68 -66 -64 -62 -60

Longitude

A A' 100

200

300

400

500

600

700

-300 -200 -100 0 100 200 300 Distance (km)

-200 -100 0 100 200 Distance (km)

-5 - 3 - 2 - 1 0 1 2 3 4 5

Figure 3.9: Synthetic test of vertical smearing of slab anomalies and comparison with velocity model. The maximum depth of the synthetic anomalies is recovered well, no significant vertical smearing of the anomalies is observed.

Page 114: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

107

Synthetic Input, Depth = 95 Synthetic Result, Depth = 95 Velocity Model. Depth = 95

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

Synthetic Input, Depth = 165

14

12

•S 10

3 8

6

4

,c 1 • 9 (.

JC? A** . . . . * • * -*—/^tt^isri"*^ * * •& * *

/ * / \ *3r»•' \^u&**-*****»*C^r

LL.c :'"'•'•••: j ; ? Vi \ V ^ / '

-72 -70 -68 -66 -64 -62 -60 Longitude

C 100

200

300

u 400^

500

600

7 0 0LL -300 -200 -100 0 100 200 300

Distance (km)

14

12

•8 10

Lat

it

00

6

4

/ ' A •

1 "•

i x * • \ *

s.c

, ,

0

. . . . • " !>^OT '::*•'•••: 7s

^

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

Synthetic Result, Depth = 165

14

12

•8 10 3 3 8

6

4

f ' 1 » 6

-O" A \ * • • • „ " • "

A/C-A. V1?^??^? V • I • • • * • • •*• 1 ^

\V .c :".'•'••:j?" ? Vi \ Y ^ ~ / '

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

C C 100

200

J 300

S" 400 Q

500

600

700

|

1 1

•1 1

1 i

I

i

1

1.

j i . 1

i

.1 '

1 • i •

-300 -200 -100 0 100 200 300 Distance (km)

Vp Perturbation (%)

14

12

-S 10

Latit

00

6

4

f 1 o fc

yy K •. . . . . * . »

/Yv-V VTV--T[!^<5

\ " ' c :*:••••: c~ ? Vi \ Y ^ ?

-74 -72 -70 - 6 8 - 6 6 - 6 4 -62 -60 Longitude

Velocity Model, Depth = 165

-74 -72 -70 -68 -66 -64 -62 -60 Longitude

C C 100

200

1 300

&400 a

500

600

700

i j

i • 1

1

i 1 j

I

( j

1 1. I 1

J' -300 -200 -100 0 100

Distance (km) 200 300

- 5 - 3

Figure 3.10: Synthetic test of data sensitivity to a shallow slab such as described by Van der Hilst and Mann [1994] and comparison with velocity model. The prescribed velocity anomaly is not recovered well in this case. The positive anomaly seen in the velocity model at depths greater than 300 km is the Caribbean slab intersecting the C-C plane from the west.

Page 115: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

108

assumptions are met in the tests. However, anisotropy has been documented in the upper

mantle in the Caribbean region [Growdon et at, 2009; Masy et at, 2009; Pinero-

Feliciangeli and Kendall, 2008; R. M. Russo et at, 1996], some degree of anelasticity is

likely and there is uncertainty in the ray paths stemming from hypocenter mislocations

and structure that is out of the volume we analyze. Therefore, these tests represent a best-

case scenario and place an upper bound on the expected resolution. As a result, the most

robust conclusions we can draw about these tests are those pertaining to which features

are irresolvable with our experiment geometry. With the above stated caveat, we

conclude that the velocity anomalies central to our analysis are well resolved both

vertically and laterally to the degree required by our interpretation. The loss of resolution

in certain areas of the model demonstrated in the synthetic tests was taken into account in

the interpretation process as discussed in the following section.

3.6. Discussion

The new images of mantle structure we present here warrant reinterpretation of

subduction processes in both the southeastern and southwestern Caribbean. In this section

we discuss our results in the context of previous studies and explore the implications of

the subduction geometries we define for regional tectonic models.

3.6.1. Absence of southward dipping slab

In the previous section we established that our experiment geometry is largely

insensitive to a very shallowly southward dipping Caribbean slab such as the one

described by van der Hilst and Mann [1994]. The signature of a shallowly dipping slab

Page 116: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

109

would be weak positive anomalies in the area above the slab and weak negative

anomalies elsewhere (Figure 3.10), whereas, our model shows low P velocity anomalies

at shallow depths under northwestern Venezuela, above the presumed location of the

shallow slab (Figure 3.10). While inconclusive in this regard, our study does not support

the existence of shallow subduction from the north. Although an early analysis of a small

subset of BOLIVAR active source data seemed to be consistent with the presence of a

shallowly subducted slab in northwestern Venezuela [Bezada et al, 2008], the full

analysis of these data [Guedez, 2007] as well as those of other boundary-normal profiles

[Clark et al, 2008; Magnani et al, 2009] showed a continuous Moho across the Southern

Caribbean Deformed Belt and beneath the Venezuelan archipelago, but no features that

would indicate a continuous Caribbean Plate subducting beneath South America from the

north. Similarly, a PdS receiver functions study found no evidence for such subduction

[Niu et al., 2007]. Lastly Rayleigh wave tomography shows low to intermediate upper

mantle shear velocities in this region, relative to further north in the Caribbean, and

further east beneath South America from the Moho to -130 km depth [Miller et al.,

2009]. We conclude that the evidence does not support a continuous slab subducting

southward beneath northwestern Venezuela.

3.6.2. Atlantic subduction

A very robust feature of this model is the landward continuation of the Atlantic

slab south of the island arc. This feature had been initially imaged by VanDecar et al.

[2003], and had been inferred since the 1970s due to the presence of igneous bodies

onshore eastern Venezuela. The Carupano Rhyolites -15 km inland from the eastern

Page 117: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

110

Venezuelan coast are Plio-Pleistocene igneous rocks with an age of ~5Ma whose

geochemical signature is consistent with bulk mixing of a slab derived source and

average continental crust [McMahon, 2001; Santamaria and Schubert, 1974]. Even more

recent magmatism is inferred from the presence of thermal springs near the El Pilar fault

system, south of where the Carupano Rhyolites outcrop. The ongoing cooling of a

shallow magmatic body is thought to be the heat source for the hot springs [Urbani,

1989]. Additional seismic evidence for the landward continuation of the slab comes from

a recent surface wave study. That S-wave velocity model shows a positive anomaly at 80-

125 km depth that correlates spatially with our image of the Atlantic slab, although it was

originally interpreted as part of the South American lithosphere [Miller et al, 2009].

Clark et al. [2008] proposed that the strike slip plate boundary produced a clean

tear through the entire lithosphere, a process that would be currently concentrated at the

Paria cluster of seismicity. This tear would separate subducting Atlantic lithosphere from

the buoyant South American lithosphere and mark the southern end of the subduction.

The distribution of intermediate depth seismicity was cited as strongly supporting this

hypothesis, as it is absent south of the strike-slip boundary. The body wave tomography

data showing the slab south of the strike slip fault in the tear region suggest a more

complicated mechanism for detaching the Atlantic lithosphere from the South American

lithosphere than a clean tear.

The Paria cluster is likely the result of water released by the subducting oceanic

plate, and high strain rates. Noting that traditional fault mechanics do not predict brittle

faulting below 50 km depth except under very high strain rates, in the last decade several

publications have pointed to dehydration embrittlement as a main cause for intermediate

Page 118: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

I l l

depth seismicity [e.g. Jung et al, 2004]. It is therefore possible that the absence of

intermediate depth seismicity south of the Paria cluster results from a lack of hydrous

minerals in the subducting lithosphere there and deformation occurring by mechanisms in

the ductile rather than brittle regime. Adding to the concentration of seismicity in the

Paria cluster could be a concentration of deformation and resulting elevated strain rates at

this point in the slab, where oceanic and continental margin lithosphere meet.

To explain the L shaped anomaly we can consider northward subduction of South

American passive margin lithosphere as suggested by [VanDecar et al, 2003]. However,

if the passive margin of South America had subducted towards the north we would expect

the E-W arm of the "L" shaped anomaly in our model to extend further west along most

of the Venezuelan coast. Furthermore, we must consider that the South American passive

margin is indeed part of the Atlantic slab and it cannot have subducted simultaneously

towards the north and towards the west under the Lesser Antilles, as this creates a

geometric problem.

We propose an alternative explanation where the strike slip fault system marks the

southern end of subduction for the Atlantic oceanic crust, but part of the northern South

American lithospheric mantle is torn off from beneath the plate and subducts along with

the Atlantic slab (Figure 3.11). Rifting formed a passive margin along northern

Venezuela during the opening of the South Atlantic and proto-Caribbean. This margin

has been tectonically active in the Cenozoic and therefore we presume it is not underlain

by a depleted, buoyant, cratonic mantle keel that would resist deformation, but rather by a

more mobile mantle capable of subduction. If this hypothesis were correct, the

lithospheric mantle subducting south of the strike slip fault system would be of a

Page 119: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

112

Figure 3.11: Cartoon to illustrate the proposed model for partial removal and subduction of South American continental margin lithosphere.

continental margin nature and it would be reasonable to assume that it did not contain

hydrous minerals. As outlined above, this would presumably prevent the occurrence of

intermediate depth seismic events and explain the absence of Benioff seismicity south of

the Paria cluster. If the mantle lithosphere indeed detached from the crust in northern

Venezuela, one might expect the upwelling of asthenospheric material to replace it. This

upwelling material could undergo decompression melting and basaltic magmatism would

be likely along northern Venezuela, but no such magmatism has been reported. We

suggest that the lithospheric mantle is only partially removed, and decompression melting

is not significant enough to produce intrusive bodies that reach the surface. We note that

[Miller et al., 2009] image a low shear velocity anomaly centered on Barcelona Bay at

depths greater than -70 km and interpret it as upwelling mantle material.

Based on this hypothesis, and the southwestward extension of the slab in our

model, we expect the northernmost -100-200 km of the Venezuelan mainland to be

Page 120: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

113

underlain by thinned lithosphere, which would leave the boundary susceptible to further

deformation and suggests isostatic rebound may play a role in the uplift of the coastal

ranges . An important corollary of our hypothesis is that the view of the San Sebastian -

El Pilar Fault system as a planar fault on a lithospheric scale [Burke, 1988; Clark et al.,

2008] needs to be revised to include more complicated plastic deformation and

downwelling of the lower lithosphere.

3.6.3. Caribbean Subduction

We begin the analysis of our image of Caribbean subduction by noting that the

steeply dipping Caribbean slab resolved in our model at depths greater than 200 km is

-500 km inboard from the trench (we assume the slab subducts under the Southern

Caribbean Deformed Belt offshore northwestern Colombia in an east-southeasterly

direction, roughly paralleling the GPS observations on San Andres island, Figure 3.1).

This would imply an initial, shallow, stage of subduction with a dip of -22° (Figure

3.13), consistent with the seismicity-derived estimates of Malave and Suarez [1995]

(25°), Pennington [1981] (20°) Ojeda and Havskov [2001] (27°), and the "northern slab"

of Zarifi et al. [2007] (25°). A stage of shallow subduction has been shown to produce

dehydration of the subducting slab at depths and temperature conditions unfavorable to

the production of dehydration-induced melting in the surrounding material [English et al,

2003] which would explain the lack of volcanism at the site of the steepening of the slab

near the Merida Andes. Shallow subduction of the Caribbean plate is widely thought to

play an important role in driving deformation and shortening in the Santa Marta Massif,

Perija ranges and Merida Andes on the overriding South American plate [Audemard and

Page 121: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

114

Audemard, 2002; Kellogg and Bonini, 1982]. Our model is consistent with that

hypothesis and additionally shows that initial shallow subduction is followed by an

abrupt steepening of the slab and that the steeply dipping slab reaches depths greater than

600 km.

Since only part of the Caribbean plate seems to be subducting under South

America, the question arises as to where the subducting section detaches or tears from the

stable section. Presumably, this would occur west of our study area, however there is no

clear indication from bathymetry, potential field data or seismicity of what the location of

this tear is. We speculate that an ancient fracture zone (obscured by the large volume

magmatism that generated the Caribbean Large Igneous Province) could provide a zone

of weakness that would facilitate such tearing, so that no significant seismicity is

produced.

An interesting problem associated with the Caribbean subduction we image is the

relationship between the down going slab and the Bucaramanga seismic nest (Figure

3.12). Seismicity studies have suggested that the seismic nest occurs within (or possibly

at the edge) of the subducting Caribbean slab [Malave and Suarez, 1995; Ojeda and

Havskov, 2001; Pennington, 1981; Zarifi et al, 2007]. Our resolution tests show that our

data do not effectively constrain the southwestward extension of the slab, and even

though we do not image the anomaly extending beyond the Colombia-Venezuela border

this is a consequence of our station distribution rather than the true slab geometry (Figure

3.8). In the future, efforts should be made to improve the image of the mantle structure in

the west of our study area and extend it westwards to include the shallow subduction

segment of the Caribbean slab as well as the area surrounding the Bucaramanga nest to

Page 122: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

Figu

re 3

.12:

Buc

aram

anga

Nes

t and

1.5

% is

ovel

ocity

con

tour

of t

he C

arib

bean

sla

b an

omal

y. M

ap v

iew

on

the

left,

3D

vie

w fr

om S

E

on th

e rig

ht. E

arth

quak

e hy

poce

nter

s w

ith d

epth

s gr

eate

r th

an 1

00 k

m a

re s

how

n. N

ote

that

the

Buc

aram

anga

Nes

t occ

urs

SW o

f the

C

arib

bean

sla

b an

omal

y be

yond

the

reso

lved

are

a of

our

mod

el.

Page 123: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

116

E c q > <u

2000

-2000 c—L

100

200

300 -

E

a. 400 Q

500

600 -

700

-300

1 "

i i i 1 | 1 I

i i

i • t

i i

!

- J ' 1 1

1 i

- j -

1 i i

i \

-200 -100 0 ^ 1 0 0 200 300

*&£ ?W^° d^sfp-**^ % ~ ° — o

^ o n B „ D ^«. N»

u XI N

>> \

\ \ i-\ \ \ \ _\ \ |-

* H * *

* " * "

* *

' » ' -\ I J _ _ _ J

i

i i i i i /

-300 -200 -100 0 100 Distance (km)

200 300

-5 -4 -2

Vp Perturbation (%) Figure 3.13: Cross section A-A' through the velocity model (location on Figure 3.9) with projected earthquake hypocenters from EHB bulletin [International Seismological Centre, 2009] (circles) and Malave and Suarez, [1995] (squares) (bottom) and corresponding elevation profile (top). SM - Santa Marta Massif; PR - Perija Ranege; LM - Lake Maracaibo; MA - Merida Andes.

Page 124: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

117

gain insight into the origin of the remarkable seismic activity in this area.

3.6.4. Correlation with transition zone topography

Due to the nature of the mineral phase transitions that generate the 410-km and

660-km seismic discontinuities, we expect the depth at which these transitions occur to be

affected by the subduction of cold material. Global and regional studies have consistently

shown a thickening of the transition zone caused by uplift of the 410-km and deepening

of the 660-km associated with subducting slabs [Vidale and Benz, 1992; Collier and

Helffrich, 1997; Flanagan and Shearer, 1998; Niu et al, 2005]. Since the slabs in our

model pass through the transition zone, we correlate our slab anomalies with the

transition zone discontinuities as imaged by PdS receiver functions [Huang et al, 2010].

The receiver function study used data from the BOLIVAR array and the IASP91 [Kennett

and Engdahl, 1991] velocity model to produce common conversion point images of 410-

km and 660-km topography. Here, we present a brief joint analysis of the two results.

Topography of the 410-km discontinuity is consistent with our image of the

Atlantic slab, as 20-35 km of uplift are observed for the NE-SW arm of the Atlantic slab

anomaly, and a smaller 7-15 km for the E-W arm (Figure 3.14). The continent-ward

extension of the slab (E-W arm) seems to produce a less significant uplift. As discussed

above, this part of the anomaly has a ~ 1 % smaller magnitude with respect to the oceanic

NE-SW oriented part of the slab. This anomaly amplitude difference implies a

temperature difference of ~ 170 K [Isaak, 1992]. We can compare this value to the

temperature difference estimated from the difference in 410-km discontinuity uplift

which gives us 100 to 150 k ( using a Calpeyron slope of 4Mpa for the phase transition

Page 125: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

118

[Katsura et ah, 2004]) and see that both estimates are roughly consistent. Considering

that the temperature at the base of the crust is higher under continents than under the

oceans by -300 K [McKenzie et ah, 2005] and the proposed nature of the E-W arm

(continental margin lithospheric mantle), a 100-170 K temperature difference between it

and the oceanic part of the slab seems plausible. We acknowledge there is a problem with

this explanation as the difference in the magnitudes of the two arms of the anomaly is

only seen at depths greater than -250 km, and if it is caused by a difference in the

original temperature it should be particularly clear at shallower depths. An additional

factor we must consider is the significant effect of water on transition zone

discontinuities [Higo et ah, 2001; Litasov, 2005; Wood, 1995]. The different response to

the part of the slab we interpret as being continental margin lithospheric mantle may be a

consequence of less water there than in the oceanic part of the slab. In the case of 660-km

discontinuity, there is a gap in the data in the area where we image the slab. However, in

the data around a slab the rim of a depression is apparent (Figure 3.14). This simple

analysis is very encouraging in that it shows consistency across two completely

independent data sets.

The response of the transition zone discontinuities to the subducting Caribbean is

even clearer, as a prominent uplift of the 410-km discontinuity of up to 44 km and an

equally important depression of the 660-km discontinuity (20-25 km) are observed with

good spatial agreement with the subducted slab (Figure 3.14). The greater response of the

discontinuities to the subducting Caribbean slab seems paradoxical given the smaller

amplitude of its velocity anomaly. However, in synthetic tests, the amplitude of the

Page 126: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

119

-68° -66° -64°

Uplift of 410 (km) -20 -10 1.0 20 30 40 •

-30 -20 -10 0 10 20 30 40 50 Depression of 660 (km)

-70° -68° -86° -64'

ATL

f-~

Figure 3.14: Subducted slabs and topography of transition zone discontinuities. 410-km (top), 660-km (middle) and 3D view from the northeast (bottom). ATL - Atlantic slab; CAR-Caribbean Slab.

Page 127: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

120

Caribbean slab was underestimated by our inversion to a greater degree than that of the

Atlantic slab (Figure 3.9), so it is likely that the real velocity anomaly produced by the

Caribbean subduction is greater than the one in our model and it may be greater than the

Atlantic slab anomaly.

3.7. Conclusions

We find the lithospheric mantle under the Caribbean plate to be relatively slow

except beneath the Leeward Antilles. Shallow slow velocity anomalies also occur east of

Lake Maracaibo and south of Barcelona Bay in central Venezuela. The Guayana Shield

shows a small positive velocity anomaly. The distribution of this anomaly suggests that

the cratonic root of the shield extends at least as far north as the Orinoco River in the east

of the shield, but does not reach as far north as the river in the west, although the exposed

crustal rocks are cratonic.

We find no evidence to support shallow southward subduction of the Caribbean

plate beneath northwestern Venezuela. However, we cannot rule out the existence of this

subduction given the lack of sensitivity of our data set and method to such a feature.

We image the steep subduction of the Atlantic slab in the east of our study area as

well as the subduction of a southernmost section of the Caribbean plate in the west. Both

of these subducting slabs go through the transition zone as confirmed by transition zone

topography. The Atlantic slab extends continent-ward south of the surface location of the

main strike slip fault system. We interpret this southward extension of the slab to be

continental margin mantle lithosphere subducting with the oceanic lithosphere of the

Atlantic. This hypothesis would imply that the southward extension of the slab is warmer

Page 128: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

121

than the main (oceanic) part of the slab, consistent with observations of a smaller velocity

anomaly and smaller associated topography of the 410-km discontinuity. Alternatively,

the difference in 410-km topography could also be caused by a lack of water in the

continent-ward extension of the slab. The Paria cluster of seismicity does not mark the

southern boundary of the subducting Atlantic slab. The absence of intermediate depth

seismicity south of the cluster may be caused by a lack of hydrated minerals in the slab

there or could be a result of concentration of strain and higher strain rates at the site of the

cluster. Partial removal of the lithosphere beneath the coastal areas of Venezuela may

help drive uplift of the coastal ranges and will leave the boundary susceptible to future

tectonic deformation.

The steep Caribbean subduction we image occurs -600 km landward of the

trench, which implies a stage of shallow subduction west of our study area consistent

with previous studies. The relationship between this Caribbean slab and the Bucaramanga

nest of seismicity remains unclear as the seismic nest lies beyond the resolvable area in

our model, but it is possible that the seismicity occurs within the slab as has been

suggested. The response of the transition zone discontinuities to the Caribbean

subduction is greater than that for the Atlantic subduction in spite of the former showing

a smaller velocity anomaly than the latter. It is likely that our velocity model

underestimates the amplitude of the Caribbean Slab anomaly.

Page 129: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

122

Appendix A: Details of GA implementation

In this appendix we describe how each new generation of solutions is created from the

previous generation. For clarity, we will make a distinction between the set of parent and

offspring solutions, since not all of the offspring solutions will pass on to the next

generation and become potential parents (this is analogous to the fact that not all

individuals in a biological population reach sexual maturity). We will refer to the

individuals in generation n as parents„ and the solutions produced by combining them as

offspring,,. The next generation of solutions parents„+i is created by replacing a portion of

parents,, with an equal number of individuals from offspringn as described in the

following paragraphs. Also for clarity, we will add a superscript to denote the fitness of

the individual, with parents^ being the fittest individual solution in generation n

(referred to in the text as the a solution), parents„ the second fittest and so on.

Recall that each solution is represented by a vector whose components are real

numbers indicating the lateral locations of province boundaries. Each boundary location

(vector component) is termed a gene. The value of each gene is limited to the range

defined by the northern and southern extremes of the model.

The application of three different operators is necessary to create generation n+1

from the solutions of known fitness in generation n. These are: mating, mutation and

replacement.

The mating operation is performed P times with P being the population size. It

begins by selecting two individuals from parentsn to mate. An elitism factor Es (with a

value ranging from 0 to 1) is introduced so that the (Es x P) fittest solutions in the

Page 130: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

123

parentsn population of solutions automatically become selected as one parent with their

mating partner being selected at random from the population. For the remaining ((1- Es) x

P) of the matings, both parents are selected at random from within the population. For

example, for P = 100 and Es = 0.3, 100 mating operations will take place in each

generation; for the first 30 mating operations parents,, through parents,, will we chosen

as one of the parents with the second parent in each case being selected at random from

parentsn. In the remaining 70 mating operations both parents will be selected randomly

from parentsn.

From the two parent solutions selected, the mating operator produces one

offspring solution by uniform crossover: The genes in the offspring solution take the

value of the corresponding gene in one of the parent solutions, with an equal probability

of choosing either parent for each gene.

Once the mating has been completed, the offspring individuals undergo mutation

by random replacement: a certain percentage of all the genes is replaced by a different

value picked at random from within the range specified. The percentage of genes

replaced is termed mutation rate mr. This completes the process of creating the offspringn

population of solutions. This offspring population is then evaluated to measure the fitness

of the individual solutions.

The evaluation operator has been described in the text. Here we note that the

linear inversion of the system of equations Gm=d (where G, m and d are as defined in

the text) is given by m=(GTG)"1GTd yielding the vector m that minimizes ||Gm-d|| in a

least squares sense.

Page 131: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

124

Finally, generation n+1 is created by replacing a number of individuals from

parentSn with an equal number of individuals from offspring,,. Here too we introduce an

elitism factor E, (with a value ranging from 0 to 1), so that fittest {E, x P) individuals of

generation n proceed unchanged to generation n+1, and the ((1- Er) x P) least fit

individuals of generation n are replaced by the ((1- Er) x P) fittest individuals of the

offspring population. For example, for a P = 100, and Er = 0.2:

parents„+i=parents„ ,parentsn ,...,parents,, ,offspring„ ,offspringn ,...,offspring„ .

This completes the process and forms the population parents„+i of known fitness

from which parents„+2 can be calculated by repeating the steps described above.

The process is repeated for a set number of generations maxg.

Clearly, an initial population of known fitness is required for the first generation.

The initial population parentso is produced by choosing randomly generated values

within the range define by the model's extremes.

There are a number of free parameters that can be adjusted such as the population

size P, the mutation rate mr, the elitism factors in the selection phase of the mating

operator Es and the replacement operator Er. The values we have used were selected

through trial and error using the noise-free synthetic dataset and are shown in Table 1.

High elitism factors tend to cause the algorithm to converge prematurely, whereas for

small values convergence slows down and the algorithm requires a greater number of

generations to achieve a given level of misfit. Variations in the mutation rate have a

similar effect; exceedingly small values may cause premature convergence while high

Page 132: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

125

values tend to delay convergence. With respect to population size, a greater value implies

a more expanded search of model space in each generation and helps the algorithm

converge to a given misfit level on fewer generations but increases the CPU time per

generation. Finally, the user specifies a number of realizations. Here again a compromise

must be made, taking into account CPU time. The number of realizations should be large

enough to create a statistically significant sample of the space of adequate solutions. We

have settled on 50 realizations for all the runs presented in this paper based on tests

performed with the noise-free synthetic data set, as it is difficult to estimate a priori what

the level of variation in the sets of solutions will be.

It is important to note that while it is difficult to estimate the optimum values of

the free parameters for a given data set, and to evaluate the effect of the interaction

between the different free parameters, the tests we conducted with synthetic data

produced comparable results over a range of different values for each of these

parameters. This allows us to conclude that although trial and error is necessary to

optimize the values of free parameters, the algorithm is robust with respect to these

values.

Page 133: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

126

Description

Number of realizations

Number of generations

Populations size

Elistism factor for selection

Mutation rate

Elitism factor for replacement

Parameter

maxr

maxg

P

Es

mr

Er

Value

50

150

45

0.3

20%

0.2

Table 1: Parameter values used in the algorithm for all tests and real data applications in the text. These values were selected by trial and error with the noise-free synthetic data set described in the text.

Page 134: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

REFERENCES

Audemard, F. E., and F. A. Audemard (2002), Structure of the Merida Andes, Venezuela; relations with the South America-Caribbean geodynamic interaction, Tectonophysics, 345(1-4), 299-327.

Barton, A. J., and R. S. White (1997), Crustal structure of Edoras Bank continental margin and mantle thermal anomalies beneath the North Atlantic, Journal of Geophysical Research, 102(B2), 3109-3129.

Barton, P. J. (1986), The relationship between seismic velocity and density in the continental crust; a useful constraint?, Geophysical Journal of the Royal Astronomical Society, #7(1 1), 195-208.

Bass, J. D. (1995), Elasticity of minerals, glasses, and melts, AGU Reference Shelf, 2, 45-63.

Beardsley, A. G., and H. G. Ave Lallemant (2007), Oblique collision and accretion of the Netherlands Leeward Antilles to South America, Tectonics, 26(2), 16.

Behn, M. D., and P. B. Kelemen (2006), Stability of arc lower crust; insights from the Talkeetna Arc section, south central Alaska, and the seismic structure of modern arcs, Journal of Geophysical Research, lll(Bll).

Bevington, P. R. (1969), Data Reduction and Error Analysis for The Physical Sciences, 1st ed., McGraw-Hill Book Company.

Bezada, M. J., M. B. Magnani, C. A. Zelt, M. Schmitz, and A. Levander (2010), The Caribbean - South American plate boundary at 65°W: Results from wide-angle seismic data, Journal of Geophysical Research, doi:10.1029/2009JB007070.

Bezada, M. J., F. Niu, T. K. Baldwin, G. Pavlis, F. Vernon, H. Rendon, C. A. Zelt, M. Schmitz, and A. Levander (2006), Crustal Thickness Variations Along the Southeastern Caribbean Plate Boundary From Teleseismic and Active Source Seismic Data, Eos. Trans. AGU Fall Meet. Suppl, 87(52), Abstract T43D-1674.

Bezada, M. J., M. Schmitz, M. I. Jacome, J. Rodriguez, F. Audemard, and C. Izarra (2008), Crustal structure in the Falcon Basin area, northwestern Venezuela, from seismic and gravimetric evidence, Journal ofGeodynamics, 45(4-5), 191-200.

Birch, A. F. (1961), The velocity of compressional waves in rocks to 10 kilobars; Part 2, Journal of Geophysical Research, 66(7), 2199-2224.

Boschetti, F., M. C. Dentith, and R. D. List (1997), Inversion of potential field data by genetic algorithms, Geophysical Prospecting, 45(3), 461-478.

Page 135: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

128

Brocher, T. M. (2005), Empirical relations between elastic wavespeeds and density in the Earth's crust, Bulletin of the Seismological Society of America, 95(6), 2081-2092.

Burke, K. (1976), Development of graben associated with the initial ruptures of the Atlantic Ocean, Tectonophysics, 36(1-3), 93-111.

Burke, K. (1988), Tectonic evolution of the Caribbean, Annual Review of Earth and Planetary Sciences, 16, 201-230.

Calais, E., and P. Mann (2009), A Combined GPS Velocity Field for the Caribbean Plate and its Margins., Eos. Trans. AGU, 90(52, Fall Meet. SuppL), Abstract G33B-0657.

Case, J. E., W. D. MacDonald, and P. J. Fox (1990), Caribbean crustal provinces. Seismic and gravity evidence, in The Caribbean Region, vol. H, pp. 15-36, Geological society of America, Boulder, Colorado.

Chen Chao, Xia Jianghai, Liu Jiangping, and Feng Guangding (2006), Nonlinear inversion of potential-field data using a hybrid-encoding genetic algorithm, Computers & Geosciences, 32(2), 230-239.

Christensen, N. I., and W. D. Mooney (1995), Seismic velocity structure and composition of the continental crust; a global view, Journal of Geophysical Research, 100(B6), 9761-9788.

Christeson, G. L., P. Mann, A. Escalona, and T. J. Aitken (2008), Crustal structure of the Caribbean-northeastern South America arc-continent collision zone, Journal of Geophysical Research, 773(B08104), doi:10.1029/2007JB005373.

Clark, S. A., C. A. Zelt, M. B. Magnani, and A. Levander (2008), Characterizing the Caribbean-South American plate boundary at 64 degrees W using wide-angle seismic data, Journal of Geophysical Research, 113(B7), @Citation B07401.

Clark, S. A., M. Sobiesiak, C. A. Zelt, M. B. Magnani, M. S. Miller, M. J. Bezada, and A. Levander (2008), Identification and tectonic implications of a tear in the South American plate at the southern end of the Lesser Antilles, Geochemistry Geophysics Geosystems, 9(11), doi:10.1029-2008GC002084.

Coffin, M. F., and O. Eldholm (1994), Large igneous provinces; crustal structure, dimensions, and external consequences, Reviews of Geophysics, 32(1), 1-36.

Collier, J. D., and G. R. Helffrich (1997), Topography of the "410" and "660" km seismic discontinuities in the Izu-Bonin subduction zone, Geophysical Research Letters, 24(12), 1535-1538.

Crawford, F. D., C. E. Szelewski, and G. D. Alvey (1985), Geology and exploration of

Page 136: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

129

the Takutu Graben of Guyana and Brazil, Journal of Petroleum Geology, 8(1), 5-36.

Dahlen, F. A., S. H. Hung, and G. Nolet (2000), Frechet kernels for finite-frequency traveltimes; I, Theory, Geophysical Journal International, 141(1), 157-174.

Dewey, J. F., and J. L. Pindell (1986), Neogene block tectonics of eastern Turkey and northern South America; continental applications of the finite difference method; discussion and reply [modified], Tectonics, 5(4), 703-705, doi:10.1029/TC005i004p00703.

Diebold, J. B., P. L. Stoffa, P. Buhl, and M. Truchan (1981), Venezuela Basin crustal structure, Journal of Geophysical Research, 86(B9), 7901-7923.

Duran, E. (2007), Inversion gravimetrica 3D en el Graben de Espino, B. S., Universidad Simon Bolivar.

English, J. M., S. T. Johnston, and K. Wang (2003), Thermal modelling of the Laramide Orogeny; testing the flat-slab subduction hypothesis, Earth and Planetary Science Letters, 214(3-4), 619-632.

Fairhead, J. D., and C. S. Okereke (1990), Crustal thinning and extension beneath the Benue Trough based on gravity studies, Journal of African Earth Sciences and the Middle East, 77(3-4), 329-335.

Feo-Codecido, G., F. D. Smith, N. Aboud, and E. de Di Giacomo (1984), Basement and Paleozoic rocks of the Venezuelan Llanos basins, Memoir - Geological Society of America, 162, 175-187.

Flanagan, M. P., and P. M. Shearer (1998), Global mapping of topography on transition zone velocity discontinuities by stacking SS precursors, Journal of Geophysical Research, 103(B2), 2673-2692.

Gardner, G. H. F., L. W. Gardner, and A. R. Gregory (1974), Formation velocity and density; the diagnostic basics for stratigraphic traps, Geophysics, 39(6), 770-780.

Godfrey, N. J., B. C. Beaudoin, S. L. Klemperer, A. R. Levander, J. H. Luetgert, A. S. Meltzer, W. D. Mooney, and A. M. Trehu (1997), Ophiolitic basement to the Great Valley forearc basin, California, from seismic and gravity data; implications for crustal growth at the North American continental margin, Geological Society of America Bulletin, 109(12), 1536-1562.

Goldberg, D. E. (1989), Genetic Algorithms in Search and Optimization, Addison-wesley.

Gorney, D. et al. (2007), Chronology of Cenozoic tectonic events in western Venezuela

Page 137: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

and the Leeward Antilles based on integration of offshore seismic reflection data and on-land geology, AAPG Bulletin, 91(5), 653-684.

Govers, R., and M. Wortel (2005), Lithosphere tearing at STEP faults: response to edges of subduction zones, Earth and Planetary Science Letters, 236(1-2), 505-523, doi:10.1016/j.epsl.2005.03.022.

Growdon, M. A., G. L. Pavlis, F. Niu, F. L. Vernon, and H. Rendon (2009), Constraints on mantle flow at the Caribbean-South American plate boundary inferred from shear wave splitting, Journal of Geophysical Research, 114(B02), doi:10.1029/2008JB005887.

Guedez, M. C. (2007), Crustal Structure Across the Caribbean-South American Plate Boundary at 70W- Results from Seismic Refraction and Reeflection Data, Rice University, Houston, Texas.

Hamilton, E. L. (1978), Sound velocity-density relations in sea-floor sediments and rocks, The Journal of the Acoustical Society of America, 63, 366.

Hess, H. H. (1966), Caribbean research project, 1965, and bathymetric chart, Memoir -Geological Society of America, 1-10.

Hess, H. H., and W. M. Ewing (1940), Continuation of a gravity survey of the Caribbean region and the correlation of gravity field with geological structure, Year Book -American Philosophical Society, 4, 236-238.

Higgs, R. (2009), Caribbean-South America oblique collision model revised, in The Origin and Evolution of the Caribbean Plate, vol. 328, pp. 611-655, Geological Society of London.

Higo, Y., T. Inoue, T. Irifune, and H. Yurimoto (2001), Effect of water on the spinel-postspinel transformation in Mg (sub 2) SiO (sub 4), Geophysical Research Letters, 28(\%), 3505-3508.

van der Hilst, R., and P. Mann (1994), Tectonic implications of tomographic images of subducted lithosphere beneath northwestern South America, Geology (Boulder), 22(5), 451-454.

Hirsch, K. K., K. Bauer, and M. Scheck-Wenderoth (2009), Deep structure of the western South African passive margin; results of a combined approach of seismic, gravity and isostatic investigations, Tectonophysics, 470(1-2), 57-70.

Hoernle, K., F. Hauff, and P. van den Bogaard (2004), 70 m.y. history (139-69 Ma) for the Caribbean large igneous province, Geology, 32(8), 697-700, doi:10.1130/G20574.1.

Page 138: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

131

Hole, J. A., and B. C. Zelt (1995), 3-D finite-difference reflection traveltimes, GeophysicalJournal International, 121(2), 427-434.

Horsefield, S. J., R. B. Whitmarsh, R. S. White, and J. Sibuet (1994), Crustal structure of the Goban Spur rifted continental margin, NE Atlantic, Geophysical Journal International, 119(\), 1-19.

Huang, J., E. Vanacore, F. Niu, and A. Levander (2010), Mantle transition zone beneath the Caribbean-South American plate boundary and its tectonic implications, Earth and Planetary Science Letters, 289(1-2), 105-111, doi:10.1016/j.epsl.2009.10.033.

Hung, E. J. (1997), Foredeep and Thrust Belt Interpretation of the Maturin Sub-Basin, Eastern Venezuela Basin, Rice University, Houston, Tex.

International Seismological Centre (2009), EHB Bulletin, [online] Available from: http://www.isc.ac.uk

Isaak, D. G. (1992), High-Temperature Elasticity of Iron-Bearing Olivines, Journal of Geophysical Research, 97(B2), PAGES 1871-1885.

Izarra, C , M. I. Jacome, M. Schmitz, and P. Mora (2005), Analyzing gravity anomalies over the Caribbean and northern Venezuela tectonic plate boundary, Barcelona, Spain.

Jaimes Carvajal, M. A. (2003), Paleogene to Recent Tectonic and Paleogeographic Evolutionof the Cariaco Basin, Venezuela,

Jung, H., H. W. Green II, and L. F. Dobrzhinetskaya (2004), Intermediate-depth earthquake faulting by dehydration embrittlement with negative volume change, Nature, 428(6982), 545-549, doi:10.1038/nature02412.

Katsura, T., H. Yamada, O. Nishikawa, M. Song, A. Kubo, T. Shinmei, S. Y. A. Y. Aizawa, T. Yoshino, M. J. W. A. E. Ito, and K. Funakoshi (2004), Olivine-wadsleyite transition in the system (Mg,Fe)2Si04, Journal of Geophysical Research, 109, B02209, doi: 10.1029/2003JB002438.

Kauahikaua, J., T. Hildenbrand, and M. Webring (2000), Deep magmatic structures of Hawaiian volcanoes, imaged by three-dimensional gravity models, Geology [Geology]. Vol. 28, (10), 883-886.

Kellogg, J. N., and W. E. Bonini (1982), Subduction of the Caribbean Plate and basement uplifts in the overriding South American Plate, Tectonics, 1(3), 251-276.

Kennett, B. L. N., and E. R. Engdahl (1991), Traveltimes for global earthquake location and phase identification, Geophysical Journal International, 105(2), 429-465.

Page 139: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

132

Kennett, B. L. N., E. R. Engdahl, and R. Buland (1995), Constraints on seismic velocities in the Earth from traveltimes, Geophysical Journal International, 122(1), 108-124,doi:10.1111/j.l365-246X.1995.tb03540.x.

King, T. S. (2006), Inverse gravity modeling for depth varying density structures through genetic algorithm, triangulated facet representation, and switching routines, University of California, Riverside, Riverside, Califronia.

Klingelhoefer, F., C. Labails, E. Cosquer, S. Rouzo, L. Geli, D. Aslanian, J. L. Olivet, M. Sahabi, H. Nouze, and P. Unternehr (2009), Crustal structure of the SW Moroccan margin from wide-angle and reflection seismic data (the Dakhla experiment); Part A, Wide-angle seismic models, Tectonophysics, 468(1-4), 63-82.

Korenaga, J., W. S. Holbrook, R. S. Detrick, and P. B. Kelemen (2001), Gravity anomalies and crustal structure at the Southeast Greenland margin, Journal of Geophysical Research, 106(B5), 8853-8870.

Lee, C. A. (2003), Compositional variation of density and seismic velocities in natural peridotites at STP conditions; implications for seismic imaging of compositional heterogeneities in the upper mantle, Journal of Geophysical Research, 108(B9), 20.

Lee, C. A., D. M. Morton, R. W. Kistler, and A. K. Baird (2007), Petrology and tectonics of Phanerozoic continent formation; from island arcs to accretion and continental arc magmatism, Earth and Planetary Science Letters, 263(3-4), 370-387.

Levander, A. et al. (2006), Evolution of the southern Caribbean Plate boundary, Eos, Transactions, American Geophysical Union, 87(9), 97, 100.

Litasov, K. D. (2005), Wet subduction versus cold subduction, Geophysical Research Letters, 32(13), doi:10.1029/2005GL022921.

Locke, B. D., and J. I. Garver (2005), Thermal evolution of the eastern Serrania del Interior foreland fold and thrust belt, northeastern Venezuela, based on apatite fission-track analyses, Special Paper - Geological Society of America, 394, 315-328.

Ludwig, W. J., J. E. Nafe, and C. L. Drake (1970), Seismic refraction.

Macedo, J., and S. Marshak (1999), Controls on the geometry of fold-thrust belt salients, Geological Society of America Bulletin, 111(12), 1808-1822.

Magnani, M. B., C. A. Zelt, A. Levander, and M. Schmitz (2009), Crustal structure of the South American-Caribbean plate boundary at 67°W from controlled source

Page 140: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

seismic data,, 114(B023\2), doi:10.1029/2008JB005817.

Malave, G., and G. Suarez (1995), Intermediate-depth seismicity in northern Colombia and western Venezuela and its relationship to Caribbean Plate subduction, Tectonics, 14(3), 617-628.

Mann, P. (1999), Caribbean sedimentary basins: Classification and tectonic setting from Jurassic to present, Caribbean basins. Sedimentary basins of the world: Amsterdam, Elsevier Science BV, 4, 3-31.

Mann, P., M. Kroehler, A. Escalona, B. Magnani, and G. Christeson (2007), Subduction along the south Caribbean deformed belt; age of initiation and backthrust origin, Eos, 88(52, Suppl.), ©Abstract Tl 1D-02.

Masy, J., F. Niu, and A. Levander (2009), Seismic anisotropy and mantle flow beneath western Venezuela, Eos. Trans. AGUFall Meet. Suppl, 90(52), Abstract T53A-1554.

McKenzie, D., J. Jackson, and K. Priestley (2005), Thermal structure of oceanic and continental lithosphere, Earth and Planetary Science Letters, 233(3-4), 337-349, doi:10.1016/j.epsl.2005.02.005.

McMahon, C. E. (2001), Evaluation of the effects of oblique collision between the Caribbean and South American plates using geochemistry from igneous and metamorphic bodies of northern Venezuela,

Miller, M., A. Levander, F. Niu, and A. Li (2009), Upper mantle structure beneath the Caribbean-South American plate boundary from surface wave tomography, Journal of Geophysical Research, 114, B01312, doi:10.1029/2007JB005507.

Montesinos, F. G., J. Arnoso, and R. Vieira (2005), Using a genetic algorithm for 3-D inversion of gravity data in Fuerteventura (Canary Islands), Geologische Rundschau = InternationalJournal of Earth Sciences (1999), 94(2), 301-316.

Moticska, P. (1985), Volcanismo Mesozoico en el Subsuelo de la Faja Petrolifera del Orinoco, Estado Guarico, Venezuela., in Memorias VI Congreso Geologico Venezolano, vol. 3, pp. 1929-1943, Caracas, Venezuela.

Niu, F., T. Bravo, G. Pavlis, F. Vernon, H. Rendon, M. Bezada, and A. R. Levander (2007), Receiver function study of the crustal structure of the southeastern Caribbean plate boundary and Venezuela, Journal of Geophysical Research, i/2(Bll),@B11308.

Niu, F., A. Levander, S. Ham, and M. Obayashi (2005), Mapping the subducting Pacific slab beneath southwest Japan with Hi-net receiver functions, Earth and Planetary Science Letters, 239(\-2), 9-17.

Page 141: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

134

Ojeda, and Havskov (2001), Crustal structure and local seismicity in Colombia, Journal of Seismology, 5(4), 575-593, doi:10.1023/A:1012053206408.

Ostos, M., F. Yoris, and H. G. Ave Lallemant (2005), Overview of the southeast Caribbean-South American plate boundary zone, Special Paper - Geological Society of America, 394, 53-89.

Paige, C. C , and M. A. Saunders (1982), LSQR: An Algorithm for Sparse Linear Equations and Sparse Least Squares, ACM Trans. Math. Softw., 8(1), 43-71, doi:10.1145/355984.355989.

Passalacqua, H., F. Fernandez, Y. Gou, and F. Roure (1995), Crustal architecture and strain partitioning in the eastern Venezuelan ranges, AAPG Memoir, 62, 667-679.

Pennington, W. D. (1981), Subduction of the eastern Panama Basin and seismotectonics of northwestern South America, Journal of Geophysical Research, 86(Q\ 1), 10753-10770.

Perez, O. J., R. Bilham, R. Bendick, J. R. Velandia, N. Hernandez, C. Moncayo, M. Hoyer, and M. Kozuch (2001), Velocity field across the southern Caribbean plate boundary and estimates of Caribbean/South-American plate motion using GPS geodesy 1994-2000, Geophysical Research Letters, 25(15), 2987-2990.

Pindell, J., and J. F. Dewey (1982), Permo-Triassic reconstruction of western Pangea and the evolution of the Gulf of Mexico/Caribbean region, Tectonics, 1(2), 179-211.

Pindell, J., and L. Kennan (2001), Kinematic Evolution of the Gulf of Mexico and Caribbean., in Proceedings ofGCSSEPM2001.

Pindell, J., and L. Kennan (2007), Cenozoic kinematics and dynamics of oblique collision between two convergent plate margins: The Caribbean-South America Collision in Eastern Venezuela, Trinidad and Barbados., in Proceedings ofGCSSEPM2007.

Pindell, J., L. Kennan, W. V. Maresch, K. Stanek, G. Draper, and R. Higgs (2005), Plate kinematics and crustal dynamics of circum-Caribbean arc-continent interactions; tectonic controls on basin development in proto-Caribbean margins, Special Paper - Geological Society of America, 394, 7-52.

Pindell, J. L. (1985), Alleghenian reconstruction and subsequent evolution of the Gulf of Mexico, Bahamas, and proto-Caribbean, Tectonics, 4(1), 1-39.

Pinero-Feliciangeli, L., and J. Kendall (2008), Sub-slab mantle flow parallel to the Caribbean plate boundaries: Inferences from SKS splitting, Tectonophysics, 462(1-4), 22-34, doi:10.1016/j.tecto.2008.01.022.

Page 142: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

135

Ravat, D., Z. Lu, and L. W. Braile (1999), Velocity-density relationships and modeling the lithospheric density variations of the Kenya Rift, Tectonophysics, 302(3-4), 225-240.

Rosencrantz, E., M. I. Ross, and J. G. Sclater (1988), Age and spreading history of the Cayman Trough as determined from depth, heat flow, and magnetic anomalies, Journal of Geophysical Research, 95(B3), 2141-2157.

Roy, L., R. Shaw, B. N. P. Agarwal, and Anonymous (2002), Inversion of gravity anomalies over sedimentary basins; applications of genetic algorithm and simulated annealing, SEG Annual Meeting Expanded Technical Program Abstracts with Biographies, 21, 747-750.

Russo, R. M., R. C. Speed, E. A. Okal, J. B. Shepherd, and K. C. Rowley (1993), Seismicity and tectonics of the southeastern Caribbean, Journal of Geophysical Research, 98(BH), 14,299-14,319.

Russo, R. M., P. G. Silver, M. Franke, W. B. Ambeh, and D. E. James (1996), Shear-wave splitting in northeast Venezuela, Trinidad, and the eastern Caribbean, Physics of The Earth and Planetary Interiors, 95(3-4), 251-275, doi:10.1016/0031-9201(95)03128-6.

Salazar, M. (2006), Evolution Estructural e Implicaciones Tectonicas del Graben de Espino, M. S., Universidad Simon Bolivar.

Santamaria, F., and C. Schubert (1974), Geochemistry and Geochronology of the Southern Caribbean-Northern Venezuela Plate Boundary, Geological Society of America Bulletin, 85(7), 1085-1098.

Schmandt, B., and E. Humphreys (2010), Seismic heterogeneity and small-scale convection in the southern California upper mantle, Geochemistry Geophysics Geosystems, doi: 10.1029/2010GC003042.

Schmitz, M., J. Avila, M. Bezada, E. Vieira, M. Yanez, A. Levander, C. A. Zelt, M. I. Jacome, and M. B. Magnani (2008), Crustal thickness variations in Venezuela from deep seismic observations, Tectonophysics, 459( 1-4), 14-26.

Schmitz, M , A. Martins, C. Izarra, M. I. Jacome, J. Sanchez, and V. Rocabado (2005), The major features of the crustal structure in north-eastern Venezuela from deep wide-angle seismic observations and gravity modelling, Tectonophysics, 399(1-4), 109-124.

Schmitz, M., D. Chalbaud, J. Castillo, and C. Izarra (2002), The crustal structure of the Guayana Shield, Venezuela, from seismic refraction and gravity data, Tectonophysics, 345(1-4), 103-118.

Page 143: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

Schubert, C. (1982), Ongin of Canaco Basin, southern Caribbean Sea, Marine Geology, 47(3-4), 345-360.

Silver, E. A., J. E. Case, and H. J. MacGillavry (1975), Geophysical study of the Venezuelan borderland, Geological Society of America Bulletin, 86(2), 213-226.

Sivanandam, S., and S. N. Deepa (2007), Introduction to Genetic Algorithms, 1st ed., Springer.

Sobolev, S. V., and A. Y. Babeyko (1994), Modeling of mineralogical composition, density and elastic wave velocities in anhydrous magmatic rocks, Surveys in Geophysics, 15(5), 515-544.

Sumanovac, F. et al. (2009), Crustal structure at the contact of the Dinarides and Pannonian Basin based on 2-D seismic and gravity interpretation of the Alp07 profile in the ALP 2002 experiment, Geophysical Journal International, 179(1), 615-633.

Talukdar, S., and E. Bolivar (1982), Petroleum geologyofTuy-Cariaco basin, eastern Venezuela continental shelf: a preliminary appraisal, Informe Tecnico, Instituto Tecnico Venezolano del Petroleo, Caracas, Venezuela.

Talwani, M., C. C. Windisch, P. L. Stoffa, P. Buhl, and R. E. Houtz (1977), Multichannel seismic study in the Venezuelan Basin and the Curacao Ridge, in Island arcs, deep sea trenches and back-arc basins, pp. 83-98, American Geophysical Union.

Talwani, M., J. L. Worzel, and M. G. Landisman (1959), Rapid gravity computations for two-dimensional bodies with application to the Mendocino submarine fracture zone [Pacific Ocean], Journal of Geophysical Research, 64(\), 49-59.

Urbani, F. (1989), Geothermal reconnaissance of northeastern Venezuela, Geothermics, 18(3), 403-427.

VanDecar, J. C , R. M. Russo, D. E. James, W. B. Ambeh, and M. Franke (2003), Aseismic continuation of the Lesser Antilles slab beneath continental South America, Journal of Geophysical Research, 108(Bl), 12.

Vidale, J. E. (1990), Finite-difference calculation of traveltimes in three dimensions, Geophysics, 55(5), 521-526.

Vidale, J. E., and H. M. Benz (1992), Upper-mantle seismic discontinuities and the thermal structure of subduction zones, Nature (London), 556(6371), 678-683.

Vose, M. D. (1999), The Simple Genetic Algorithm: Foundations and Theory, The MIT Press.

Page 144: Maximiliano J. Bezada A THESIS SUBMITTED Doctor of Philosophy · Adeyemi Arogunmati, Jeniffer Masy and especially Maria Guedez, without whom I may not have survived my first year

137

Watlington, R., W. Wilson, W. Johns, and C. Nelson (2002), Updated bathymetric survey of Kick-'em- Jenny submarine volcano, Marine Geophysical Researches, 23(3), 271-276, doi:10.1023/A:1023615529336.

Weber, J. C , T. H. Dixon, C. DeMets, W. B. Ambeh, P. Jansma, G. Mattioli, J. Saleh, G. Sella, R. Bilham, and O. Perez (2001), GPS estimate of relative motion between the Caribbean and South American plates, and geologic implications for Trinidad and Venezuela, Geology [Geology]. Vol. 29, (1), 75-78.

Wood, B. J. (1995), The effect of H (sub 2) O on the 410-kilometer seismic discontinuity, Science, 268(5207), 74-76.

Ysaccis, R. (1997), Tertiary Evolution of the Northeastern Venezuela Offshore, Rice University, Houston, Tex.

Zarifi, Z., J. Havskov, and A. Hanyga (2007), An insight into the Bucaramanga nest, Tectonophysics, 443(1-2), 93-105.

Zelt, C. A., and D. A. Forsyth (1994), Modeling wide-angle seismic data for crustal structure; southeastern Grenville Province, Journal of Geophysical Research, 99(B6), 11.

Zelt, C. A., and R. B. Smith (1992), Seismic traveltime inversion for 2-D crustal velocity structure, GeophysicalJournalInternational, 108(1), 16-34.

Zelt, C. A. (1998), Lateral velocity resolution from three-dimensional seismic refraction data, GeophysicalJournal International, 135(3), 1101-1112.

Zelt, C. A., and P. J. Barton (1998), Three-dimensional seismic refraction tomography; a comparison of two methods applied to data from the Faeroe Basin, Journal of Geophysical Research, 103(B4), 7187-7210.

Zelt, C. A. (1989), Seismic structure of the crust and upper mantle in the Peace River Arch region, Ph.D. Thesis, University of British Columbia, 223 pp., Vancouver, Canada.