You are on page 1of 383

Tectonic evolution, fault architecture, and paleo-fluid

circulation in transpressive systems - southern Haiti


Richard Wessels

To cite this version:


Richard Wessels. Tectonic evolution, fault architecture, and paleo-fluid circulation in transpressive sys-
tems - southern Haiti. Earth Sciences. Sorbonne Université, 2018. English. �NNT : 2018SORUS220�.
�tel-02484820v2�

HAL Id: tel-02484820


https://theses.hal.science/tel-02484820v2
Submitted on 20 Feb 2020

HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est


archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents
entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non,
lished or not. The documents may come from émanant des établissements d’enseignement et de
teaching and research institutions in France or recherche français ou étrangers, des laboratoires
abroad, or from public or private research centers. publics ou privés.
DOCTORAL THESIS OF THE SORBONNE UNIVERSITY

Domain:
Geosciences

L’Institut des Sciences de la Terre de Paris (Paris)

Presented by:
Richard WESSELS

To obtain the grade of:


DOCTOR of the Sorbonne University

Thesis subject:
Tectonic evolution, fault architecture, and paleo-fluid circulation
in transpressive systems – southern Haiti

Supervised by Nadine Ellouz-Zimmermann and Sylvie Leroy

Defended on the 27th of June 2018

Members of the Jury:


Prof. Olivier Lacombe Sorbonne Université Examiner
Prof. Sveva Corrado Universita degli Studi di Roma III Reviewer
Dr. Manuel Pubellier Ecole Normal Superieure Paris Reviewer
Prof. James Evans Utah State University Examiner
Prof. Rudy Swennen KU Leuven Examiner
Prof. Roberte Momplaisir Université d’Etat d’Haïti Invited
Prof. Dominique Boisson Université d’Etat d’Haïti Invited
Dr. Nadine Ellouz-Zimmermann IFP Energies nouvelles Thesis Director
Dr. Sylvie Leroy ISTeP, Sorbonne Université Co-Director
2
Abstract (English)
Haiti is located on the western part of the island of Hispaniola, shared with the Dominican Republic in the east.
Haiti is situated within the northern Caribbean plate boundary region where relative motion between the
Caribbean and North American plates is accommodated by a complex system of fault-bounded microplates and
tectonic blocks. Two seismogenic strike-slip faults related to this system are found in Haiti; the Enriquillo –
Plantain Garden Fault Zone (EGPFZ) onshore southern Haiti, and the Septentrional Fault Zone (SFZ) offshore
northern Haiti, with the southwest-verging, forward-propagating Haitian Fold-and-Thrust Belt situated in
between them.

The geology and geodynamic setting of Haiti became the focus of increased scientific interest following the
January 12th 2010 Mw 7.0 Leogâne earthquake, which struck southern Haiti close to its capital Port-au-Prince.
This study, which is a collaboration between Sorbonne Université, IFP Energies nouvelles (IFPEn), Université
d’Etat d’Haïti (UEH), URGéo, and Bureau des Mines et de l’Energie d’Haïti (BME), is dedicated to increase our
knowledge of the onshore geology of southern Haiti. There are three main objectives to this study: 1) identify
the number and timing of deformation phases on the Southern Peninsula of Haiti, their regional impact, and the
associated structural style of deformation and paleo-stress evolution, 2) constrain the deformation history of the
southernmost onshore part of the Haitian Fold-and-Thrust Belt (the Chaîne des Matheux), the structural style of
deformation and the associated paleo-stress evolution, and 3) characterize the interaction between fluids and
deformation by examining the paleo-fluid circulation related to deformation in both regions.

To fulfill these objectives this study integrates geological data and observations onshore Haiti from field
campaigns in 2015 and 2017. Stratigraphic and structural data are combined with satellite imagery and digital
elevation models to create four small-scale (~1:50.000) geological maps and associated cross sections. These are
used to better understand and constrain the style of deformation in the region. Samples of host rocks and veins
from fault zones and fractures are analyzed using a suite of analytical techniques, which include optical and
cathodoluminescence microscopy, fluid inclusion microthermometry and Raman spectroscopy on fluid
inclusions, X-ray diffraction, stable oxygen and carbon isotope geochemistry, and whole-rock geochemistry, all
of which are integrated to constrain the paleofluid circulation.

The results of this study indicate that 1) the Southern Peninsula evolved by basement-involved inversion,
thrusting, and strike-slip, while 2) the style of deformation in the Chaîne des Matheux is predominantly thin-
skinned controlled by shallow dipping decollement levels, although a component of thick-skinned basement-
involved deformation is probable. The polyphase deformation history of the Southern Peninsula is characterized
by three major tectonic events: 1) Deformation and uplift during the Maastrichtian and early Paleocene, 2) early
Miocene compression and uplift that mainly affected the southwestern part of the Southern Peninsula, and 3)
transpressive deformation from the late Miocene to recent. This last phase is characterized by a progressive
focus of strike-slip activity along the EPGFZ, which accommodated a maximum of 15 km of left-lateral
displacement since the late Messinian. The EPGFZ is characterized by predominantly strike-slip in the west and
progressively more oblique-slip faults in the east, and paleo-stresses indicate that the fault zone is mechanically
weak. During phases 1 and 2 paleo-fluid circulation was mainly intra-formational, while the activation of the
EPGFZ provided a pathway for fluids derived from mantle reservoirs. Thrust faults rooted on the EPGFZ and minor
strike-slip faults delineating pull-apart basins along its trace do not show this mantle component, but rather acted
as pathways for the downwards percolation of meteoric surface waters. The evolution of the Chaîne des Matheux
from the Miocene onwards is subdivided into four tectonic phases; 1) middle Miocene, probably Serravallian,
layer-parallel shortening, 2) Serravallian to early Tortonian strike-slip, 3) thin-skinned folding-and-thrusting from
the Tortonian to the onset of the Pleistocene, 4) Quaternary strike-slip activity along the southern front of the
Chaîne des Matheux. Faults related to deformation phases 1 to 3 acted as pathways for the downwards
percolation of meteoric fluids, while Quaternary strike-slip and/or deeply rooted basement thrusts possibly
created pathways for mantle-derived fluids to ascent to the surface.

3
4
Résumé (Français)
Haïti est située sur la partie occidentale de l'île d'Hispaniola, qu'il partage avec la République dominicaine à l'est.
Haïti est située en limite septentrionale des Caraïbes, où le mouvement relatif entre les plaques Caraïbes et
Amérique du Nord est accommodé par un système complexe de microplaques de failles et de blocs tectoniques.
Deux failles décrochantes sismogènes liées à ce système se trouvent en Haïti ; la zone de faille d'Enriquillo –
Plantain Garden (EGPFZ) sur la partie sud d'Haïti et la zone de faille Septentrionale (SFZ) au large de la partie nord
d'Haïti, tandis que la chaîne trans-haïtienne, composée de chevauchements d’unités tectoniques haïtiennes se
propage vers le sud-ouest.

La géologie et le contexte géodynamique d'Haïti font l'objet d'un regain d'intérêt scientifique à cause du séisme
très destructeur de Mw 7.0, survenu le 12 janvier 2010, qui s’est produit au sud d'Haïti à Leogâne, près de la
capitale Port‐ au‐ Prince. Cette étude, qui est une collaboration entre Sorbonne Université, IFP Energies nouvelles
(IFPEn), l’Université d’Etat d’Haïti (UEH), URGéo, et le Bureau des Mines et de l’Energie d’Haïti (BME), participe
à l'accroissement des connaissances géologiques de la zone méridionale d'Haïti. Elle comporte trois objectifs
principaux ; 1) identifier le nombre et la chronologie des phases de déformation de la péninsule sud d'Haïti, leur
impact régional, leur style structural, et l'évolution des paléo-contraintes, 2) contraindre l'histoire de la
déformation de la zone frontale de la chaîne trans-haïtienne (la chaîne des Matheux), et (3) caractériser
l'interaction entre les fluides et la déformation en examinant la circulation des paléo‐fluides associée à la
déformation dans les deux régions.

Pour atteindre ces objectifs, cette étude intègre des données géologiques et des observations à terre sur Haïti
acquises lors de deux campagnes de terrain en 2015 et 2017. Les données stratigraphiques et structurales sont
combinées avec l’étude des images satellitaires pour établir quatre cartes géologiques à petite échelle
(~1:50.000) et des coupes transversales, qui permettent de mieux comprendre et de contraindre la déformation
dans la région. Des échantillons de roches et de veines sont analysés à l'aide d'une série de techniques
analytiques, incluant la microscopie optique, la cathodoluminescence, la micro‐thermométrie des inclusions
fluides et la spectroscopie Raman sur ces inclusions, la diffraction des rayons X, la géochimie des isotopes stables,
et la géochimie sur roche totale. L'ensemble des analyses est intégré afin de documenter et comprendre la
circulation des paléo‐fluides.

Les résultats de cette étude indiquent que la Péninsule du Sud s’est développée sur une large zone d’inversion
bordée par des chevauchements impliquant le socle, tandis que la déformation dans la Chaîne des Matheux est
principalement à l’origine d'une tectonique contrôlée par des niveaux de décollement peu profonds, suivie
tardivement par des inversions de socle. L'histoire de déformation de la Péninsule du Sud est polyphasée et
caractérisée par trois phases tectoniques majeures; 1) Compression et soulèvement durant le Maastrichtien et
le Paléocène inférieur, 2) compression et soulèvement du Miocène inférieur qui affectaient principalement la
partie sud‐ ouest de la Péninsule du Sud, et 3) déformation transpressive du Miocène supérieur à l’actuel. Cette
dernière phase est caractérisée par une concentration progressive d'activité décrochante le long de l'EPGFZ, qui
enregistre un maximum de 15 km de déplacement sénestre depuis la fin du Messinien. L'EPGFZ se caractérise
principalement par un décrochement à l'ouest et évolue vers un système transpressif oblique à l'est. L’étude des
paléo-contraintes indique que la zone de faille est mécaniquement faible. Au cours des phases 1 et 2, la
circulation des paléo‐fluides était principalement intra‐formationelle, tandis que l'activation de l'EPGFZ
fournissait un chemin pour les fluides dérivés des réservoirs de la croûte ou du manteau. Les chevauchements
enracinés sur l'EPGFZ et les failles mineures qui délimitaient les bassins décrochants le long de son parcours ne
montrent pas de paléo‐fluides riches en éléments chimiques provenant du manteau, mais servent plutôt de
zones de passage pour la percolation de l’eau météorique ou intraformationnelle. L'évolution de la Chaîne des
Matheux à partir du Miocène est subdivisée en quatre phases tectoniques ; 1) Miocène moyen, probablement
Serravallien, du raccourcissement parallèle à la chaîne , 2) décrochement d’âge Serravallien à Tortonien inferieur
probable, 3) tectonique de couverture avec le développement de plis et chevauchements du Tortonien jusqu’au
début du Pléistocène, et 4) décrochement Quaternaire le long de la bordure sud de la Chaîne des Matheux et
inversion tardive probable de socle. Les failles liées aux phases 1 à 3 servent plutôt de passage à la percolation
de l’eau météorique, tandis que le décrochement quaternaire a probablement créé des chemins pour des fluides
hydrothermaux plus profonds (croûte, manteau).

5
6
Table of content
Abstract (English) ......................................................................................................................... 3

Résumé (Français) ........................................................................................................................ 5

Table of content ........................................................................................................................... 7

1 Introduction ............................................................................................................................ 11

2 Geological review of the Caribbean region ............................................................................... 15


2.1 Geodynamics of the Caribbean ................................................................................................... 15
2.2 Origin of the Caribbean plate ...................................................................................................... 17
2.3 Caribbean crustal structure ......................................................................................................... 20
2.4 Mesozoic evolution of the Caribbean realm ............................................................................... 25
2.5 Seismic tomography and paleomagnetism in the northern Caribbean ...................................... 28
2.6 Accepted chapter for Elsevier book publication “Transform plate boundaries and fracture
zones”, edited by João C. Duarte ...................................................................................................... 30

3 Geological review of Hispaniola ............................................................................................... 51


3.1 Morphotectonic zones ................................................................................................................ 51
3.2 Igneous and metamorphic domains in Hispaniola ...................................................................... 53
3.2.1 CLIP-related tholeiitic flood basalts...................................................................................... 55
3.2.2 Island arc units related to the Great Arc of the Caribbean .................................................. 57
3.2.3 Basalts with a mantle-plume component ............................................................................ 59
3.2.4 Metamorphism ..................................................................................................................... 60
3.2.5 Cenozoic volcanism .............................................................................................................. 63
3.3 Tectono-stratigraphy of western Hispaniola ............................................................................... 64
3.3.1 Late Cretaceous .................................................................................................................... 64
3.3.2 Paleocene ............................................................................................................................. 64
3.3.3 Eocene .................................................................................................................................. 65
3.3.4 Oligocene .............................................................................................................................. 67
3.3.5 Miocene ................................................................................................................................ 68
3.3.6 Pliocene – Recent ................................................................................................................. 70
3.3.7 Evolution of the Peralta Belt ................................................................................................ 71
3.4 Deformation phases .................................................................................................................... 72
3.4.1 Laramide phase .................................................................................................................... 72
3.4.2 Eocene deformation ............................................................................................................. 72
3.4.3 Oligocene - Miocene deformation ....................................................................................... 73
3.4.4 Miocene – Recent deformation ........................................................................................... 73

4 Geology of the Southern Peninsula .......................................................................................... 75


4.1 Introduction ................................................................................................................................. 75
4.2 Submitted paper to Tectonophysics ........................................................................................... 77
4.3 Latest Cretaceous – Paleocene palinspastic position ................................................................ 145

7
4.4 Development of orthogonal joints ............................................................................................ 146
4.5 Shortening of the Southern Peninsula ...................................................................................... 148
4.6 Displacement along the EPGFZ.................................................................................................. 151
4.6.1 Previous work on the EPGFZ .............................................................................................. 153
4.6.2 Approach ............................................................................................................................ 154
4.6.3 Offset of the Rivière Momance and Rivière Froide ............................................................ 156
4.6.4 Evolution of the Rivière Momance – Rivière Froide EPGFZ segment................................. 156
4.6.5 Offset of the Rivière de Cavaillon ....................................................................................... 158
4.6.6 Evolution of the Rivière de Cavaillon segment................................................................... 160
4.6.7 Discussion ........................................................................................................................... 160

5 Geology of the Chaîne des Matheux ........................................................................................163


5.1 Stratigraphy of the Chaîne des Matheux .................................................................................. 164
5.1.1 Pre-Eocene stratigraphy ..................................................................................................... 164
5.1.2 General stratigraphy of the eastern Chaîne des Matheux ................................................. 166
5.2 Haitian fold-and-thrust belt....................................................................................................... 168
5.3 Detailed stratigraphy of the eastern Chaîne des Matheux ....................................................... 170
5.3.1 Stratigraphy Trou Caïman – Mirebalais section ................................................................. 170
5.3.2 Stratigraphy Titanyen - Dubon section............................................................................... 173
5.4 Outcrop-scale style of deformation .......................................................................................... 176
5.5 Map-scale deformation structures ............................................................................................ 181
5.6 Shortening of the Chaîne des Matheux ..................................................................................... 182
5.7 Paleo-stress and relative timing ................................................................................................ 184
5.8 Absolute timing of paleo-stress states and deformation phases .............................................. 188

6 Paleo-fluid circulation .............................................................................................................191


Nomenclature.................................................................................................................................. 191
6.1 Structure, mechanisms, and fluid flow properties of fault zones ............................................. 193
6.1.1 Characteristics of fault zones, fault cores, and damage zones .......................................... 193
6.1.2 Lithological control on fault zone and fluid flow characteristics ....................................... 195
6.2 Vein formation, morphology, and deformation ........................................................................ 200
6.2.1 Mechanisms of vein formation .......................................................................................... 200
6.2.2 Vein morphology ................................................................................................................ 204
6.2.3 Mechanisms of vein deformation ...................................................................................... 206
6.3 Methodology ............................................................................................................................. 208
6.4 Results from the Chaîne des Matheux ...................................................................................... 210
6.4.1 Tectonic setting, vein types, and microscopic observations .............................................. 211
6.4.2 Stable oxygen and carbon isotopes .................................................................................... 216
6.4.3 Fluid inclusion microthermometry ..................................................................................... 218
6.4.4 Whole-rock geochemistry .................................................................................................. 220
6.4.5 Discussion of vein timing, geochemistry, and fluid sources ............................................... 222
6.5 Paleo-fluid circulation and evolution of the Chaîne des Matheux ............................................ 227
6.6 Results from the Southern Peninsula ........................................................................................ 229
6.6.1 Tectonic setting, vein types, and microscopic observations .............................................. 231

8
6.6.2 Stable oxygen and carbon isotopes .................................................................................... 238
6.6.3 Fluid inclusion microthermometry ..................................................................................... 242
6.6.4 Whole-rock geochemistry .................................................................................................. 244
6.6.5 Discussion of vein timing, geochemistry, and fluid sources ............................................... 248
6.7 Paleo-fluid circulation and evolution of the Southern Peninsula ............................................. 252

7 Discussion and regional implications .......................................................................................255


7.1 Shortening and shortening rates in southwestern Hispaniola .................................................. 255
7.1.1 Comparison of shortening with published cross sections .................................................. 256
7.1.2 Distribution of shortening .................................................................................................. 260
7.1.3 Implications for displacement along the EPGFZ ................................................................. 260
7.1.4 Average geological shortening rates and implications for SW Hispaniola ......................... 264
7.2 Miocene to present-day paleogeographic reconstruction of the Southern Peninsula in Haiti 268
7.2.1 Structural constraints ......................................................................................................... 268
7.2.2 Paleogeographic reconstruction maps ............................................................................... 273
7.3 Crustal-scale cross-section ........................................................................................................ 276
7.3.1 Crustal structure ................................................................................................................. 276
7.3.2 Cross section....................................................................................................................... 279
7.3.3 Fluid flow pathways............................................................................................................ 281
7.3.4 Geodynamic implications ................................................................................................... 283

8 Conclusions and perspectives ..................................................................................................285


8.1 Conclusions ................................................................................................................................ 285
8.1.1 Southern Peninsula ............................................................................................................ 285
8.1.2 Chaîne des Matheux ........................................................................................................... 286
8.2 Perspectives............................................................................................................................... 287

List of figures.............................................................................................................................289

References ................................................................................................................................299

Appendix 1 – Cross sections for shortening calculations (see Table 7 - 1 for references) ..............337

Appendix 2 – Extended stratigraphic field observations Southern Peninsula ...............................340


A2.1 Western domain (Tiburon) .............................................................................................. 340
A2.2 Central domain (L’Asile) .................................................................................................. 342
Interpretation .............................................................................................................................. 344
A2.3 Eastern domain (Marigot) ............................................................................................... 345

Appendix 3: Geological maps and cross sections.........................................................................349

Appendix 4: Representative bedding measurements ..................................................................357

Appendix 5: Whole-rock data ....................................................................................................360

Acknowledgements ...................................................................................................................380

9
10
Chapter 1: Introduction

1 Introduction

Haiti is located on the western part of the island of Hispaniola, which it shares with the Dominican
Republic in the east. It is situated in the northern Caribbean, with Cuba and Jamaica to the west and
Puerto Rico to the east. The island has a general hot and humid tropical climate, with unevenly
distributed temperatures and annual rainfall (Bourgueil et al., 1988). This partly results from the
rugged topography of Hispaniola, which is characterized by high mountainous regions and relatively
flat lowlands. The highest peak on Hispaniola is the Pico Duarte in the Dominican Republic, which at
3098m elevation is the highest peak of all the Caribbean islands. The highest peaks in Haiti are the Pic
la Selle (2680m), Pic Macaya (2347m), and Morne la Visite (2280m), located in the south of the country.
Surrounding the island are a number of basins and troughs, which reach depths in excess of -5000m.

Haiti is situated within the northern Caribbean plate boundary region (Figure 2 – 1). In this area,
relative motion between the Caribbean and North American plates (DeMets et al., 2010) is
accommodated by a complex system of fault-bounded microplates and tectonic blocks (Burke et al.,
1978; Mann et al., 1995, 2002; Benford et al., 2012a; Symithe et al., 2015; Calais et al., 2016). Two
seismogenic strike-slip faults related to this system are found in Haiti; the Enriquillo – Plantain Garden
Fault Zone (EGPFZ) onshore southern Haiti (Mann et al., 1984), and the Septentrional Fault Zone (SFZ)
offshore northern Haiti (Calais and Mercier de Lépinay, 1995), while the southwest-verging, forward-
propagating Haitian Fold-and-Thrust Belt is situated in between them (Pubellier et al., 2000).
Earthquakes related to these strike-slip faults are documented by historical accounts since around
1500 A.D. (i.e. de Saint-Méry, 1798; Scherer, 1912; McCann, 2006; ten Brink et al., 2011). A possible
correlation between earthquake activity and spring activity is documented for a source around
Mirebalais (La Roche), in south-central Haiti (de Saint-Méry, 1798). This spring, with mildly warm
sulfur-rich waters that originate from fractured limestones, was flowing between 1750 and 1751, but
stopped shortly after the October 1751 earthquake. The spring started flowing again in 1760, only to
stop following the 1770 earthquake, after which it started flowing again in 1777 (de Saint-Méry, 1798).

The oldest comprehensive account of the geology of Haiti comes from the pioneering work by
Woodring et al. (1924), who also published the first geological map of Haiti in 1923. The Haitian geology
is extensively studied by Jacques Butterlin throughout the 1950s, ‘60s, and ‘70s (i.e. Butterlin, 1953,
1954, 1960, 1972). The late ‘70s saw an increase in geological research in Haiti (i.e. Butterlin et al.,
1976; Maurrasse et al., 1977, 1979; Mercier de Lépinay et al., 1979), which was partly in response to
the realization that Cretaceous basalts in southern Haiti are analogues to the Cretaceous flood basalts
on the Caribbean plate (Donnelly et al., 1973; Edgar et al., 1973). This also resulted in a number of PhD
theses focused towards the geology of Haiti in the 1980s (i.e. Van den Berghe, 1983; Calmus, 1983;

11
Chapter 1: Introduction

Bien-Aime Momplaisir, 1986; Boisson, 1987; Desreumaux, 1987). These studies, together with the
work of Woodring and Butterlin, are partly summarized in an extensive report by Bourgueil et al.
(1988). With the increased understanding about the geology of Haiti, a series of four 1:250.000 scale
are published by the Bureau de Mines et de l’Énergie d’Haïti (BME) in the late 1980’s (Boisson and
Pubellier, 1987; Boisson and Bien-Aime Momplaisir, 1987; Boisson and Pubellier, 1988; Bien-Aime
Momplaisir et al., 1988). Comparatively few studies focused on the onshore geology throughout the
‘90s and 2000s (i.e. Pubellier et al., 1991; Mann et al., 1995; Amilcar, 1997; Pubellier et al., 2000), with
attention shifting towards the offshore domain (i.e. Mauffret and Leroy, 1999; Leroy et al., 2000).

This changed after the January 12th 2010 Mw 7.0 Leogâne earthquake, which struck close to the capital
Port-au-Prince (Bilham, 2010a). Based on the understanding of the regional geology, it was initially
assumed that this earthquake occurred on the EPGFZ, but this was quickly disproven (Bilham, 2010b).
Subsequent studies showed that the geometry of the fault system in southern Haiti is complex (Hayes
et al., 2010; Prentice et al., 2010), and that the earthquake resulted from oblique displacement on a
steeply dipping blind fault (Calais et al., 2010; Mercier de Lépinay et al., 2011; Douilly et al., 2013). Pre-
2010 GPS velocity models did not predict the stress build-up associated with this earthquake (DeMets
et al., 2000; Calais et al., 2010), which were subsequently updated to better match the main and
aftershock patterns (Benford et al., 2012a; Symithe et al., 2015; Calais et al., 2016). A seismic
acquisition campaign launched after the earthquake helped to better characterize the geometry of the
faults offshore Haiti (Leroy et al., 2015; Corbeau et al., 2016a; b), while onshore seismometers were
used to constrain the thickness of the crust (Corbeau et al., 2017). Onshore work by a team from IFPEn
in 2014 and 2015 further improved our knowledge about the stratigraphy of Haiti.

Fault zones exert a strong control on the crust’s mechanical properties and may provide economically
significant fluid flow pathways (Faulkner et al., 2010). The structure and fluid flow properties of fault
zones evolve over space and time during fault activity (Kim and Sanderson, 2005, 2010). The
characteristics of a fault zone are controlled by the depth at which it forms (Ishii et al., 2010), the type
of lithology (Balsamo et al., 2010; Delle Piane et al., 2017), fault displacement (Evans, 1990; Wibberley
et al., 2008; Savage and Brodsky, 2011), and tectonic setting (Kim and Sanderson, 2005; Faulkner et
al., 2010). The presence of elevated fluid pressures within a rock or fault zone changes its failure
properties, which in turn has a strong impact on the seismogenic behavior of a fault zone (Twiss and
Moores, 2007; Fossen, 2010). A strong interplay exists between the mechanical characteristics of a
fault zone, and its fluid flow properties and geochemistry (Bons et al., 2012). This makes studying the
deformation history of a fault zone and the associated geochemical evolution of fault-related fluids
and mineralizations an important topic in understanding the seismic risk of a fault zone, and its
temporal and spatial evolution.

12
Chapter 1: Introduction

This study, which is a collaboration between Sorbonne Université, IFP Energies nouvelles (IFPEn),
Université d’Etat d’Haïti (UEH), URGéo, and Bureau des Mines et de l’Energie d’Haïti (BME), is
dedicated to increase our knowledge of the onshore geology of southern Haiti. There are three main
objectives to this study; 1) identify the number and timing of deformation phases on the Southern
Peninsula of Haiti, their regional impact, and the associated structural style of deformation and paleo-
stress evolution, 2) constrain the deformation history of the southernmost onshore thrust sheet of the
Haitian Fold-and-Thrust Belt (the Chaîne des Matheux), the structural style of deformation and the
associated paleo-stress evolution, and 3) characterize the interaction between fluids and deformation
by examining the paleo-fluid circulation related to deformation in both regions.

To fulfill these objectives this study integrates geological data and observations onshore Haiti from
field campaigns in 2015 and 2017. Stratigraphic and structural data is combined with satellite imagery
and digital elevation models to create four small-scale (~1:50.000) geological maps and cross sections,
which aid to better understand and constrain the deformation history in the region. Samples of host
rocks and veins from fault zones and fractures are analyzed using optical and cathodoluminescence
microscopy, fluid inclusion microthermometry and Raman spectroscopy on fluid inclusions, x-ray
diffraction, stable oxygen and carbon isotope geochemistry, and whole-rock geochemistry, all of which
are integrated to constrain the paleofluid circulation.

The thesis is subdivided into 8 chapters, of which this introduction is the first.

Chapter 2 is a literature review focusing on the geology of the Caribbean region. It assesses the
geodynamic setting of the Caribbean, the origin of the Caribbean Plate, its crustal structure, the
geodynamic evolution, and the present-day geometry of the northern Caribbean. The last part of this
chapter addresses the setting and structural evolution of the strike-slip faults in the northern
Caribbean. It identifies discrepancies in the current understanding of the evolution of the northern
Caribbean plate boundary region and Haiti in particular, and proposes possible solutions. It is accepted
as a chapter titled “Strike-slip fault systems along the northern Caribbean plate boundary” for the
Elsevier book publication “Transform plate boundaries and fracture zones”, edited by J.C. Duarte.

Chapter 3 is a literature review dedicated to the geology of Hispaniola, and Haiti in particular. It
identifies the different igneous and metamorphic domains in Hispaniola, the tectono-stratigraphy of
the western part of the island, and the number of known deformation phases and their timing. This
chapter serves as the framework with which to compare the results of this study.

Chapter 4 focuses on the geology of the Southern Peninsula in Haiti, as evidenced by this study. Section
4.2 focusses on the structural and stratigraphic evolution of the Southern Peninsula from the Late
Cretaceous through the Cenozoic era. It addresses the number of deformation phases, the relationship

13
Chapter 1: Introduction

between sedimentation and deformation, the responsible paleo-stress patterns and structural
controls on deformation, timing and development of the EPGFZ, and the geometry of the EPGFZ. These
results are presented in a paper submitted to Tectonophysics, titled “Polyphase tectonic history of the
Southern Peninsula, Haiti: From folding-and-thrusting to transpressive strike-slip”. Based on the results
in this paper, the latest Cretaceous to Paleocene palinspastic position of the Southern Peninsula is
tentatively revised. The setting of different joint sets and shortening of the Southern Peninsula in
response to the different deformation events is addressed next. Lastly, the displacement along the
EPGFZ is constrained using geomorphological criteria.

Chapter 5 is dedicated to the geology of the Chaîne des Matheux and the evolution of the fold-and-
thrust system in southern Haiti. It first constrains the stratigraphy of the area of interest, followed by
the deformation styles observed in outcrops. The regional deformation is addressed next, which is
subsequently summarized and placed in an absolute time frame. The chapter finishes by assessing the
amount of shortening in the Chaîne des Matheux.

Chapter 6 aims at characterizing the paleofluid circulation in southern Haiti. The first two parts are a
literature review on the structure, mechanism, and fluid flow properties of fault zones, and the
formation mechanisms of veins, their morphology, and deformation. After the methodology section,
two parts focus on the structural and geochemical characteristics of veins in the Chaîne des Matheux
and Southern Peninsula. Both parts follow the same outline. They start with the identification of
different vein sets and their corresponding structural setting and relative timing. Microscopic
observations on these veins are shown next, followed by the results of stable oxygen and carbon
isotope geochemistry, fluid inclusion microthermometry, and whole-rock geochemistry. The absolute
timing of the veins and the geochemical trends are addressed next, together with the possible sources
of the fluids. The chapter finishes with a conceptual model of the evolution of paleofluid circulation in
the Chaîne des Matheux and Southern Peninsula, respectively.

Chapter 7 discusses the results of the preceding chapters and assesses their regional implications. It is
subdivided into three parts. The first focusses on the amount of shortening and geological shortening
rates in southwestern Hispaniola, which provides additional constraints on the amount of
displacement along the EPGFZ, and how this is accommodated. In the next part the deformation
history of the Southern Peninsula is integrated with the stratigraphic evolution to construct 8
paleogeographic reconstruction maps. The last part integrates the cross sections from this study into
the crustal-scale, and uses this to discuss possible fluid pathways and sources in southern Haiti, as well
as the implications for the regional geodynamic framework.

Chapter 8 lists the main conclusions from this thesis and provides perspectives for future work.

14
Chapter 2: Geological review of the Caribbean region

The Cocos plate is subducting towards the NNE under Central America with velocities of 68 – 82 mm
yr-1 at the Middle America Trench (DeMets, 2001). Relative motion between North America and a fixed
Caribbean plate of 19 – 21 mm yr-1 towards the WSW (DeMets et al., 2007) is accommodated by west-
verging subduction of North Atlantic lithosphere under the Caribbean plate at the Lesser Antilles
Trench (Figure 2 - 2d to f). This subducting slab is connected to the North American plate that bends
to subduct under Puerto Rico at the Puerto Rico Trench and eastern Hispaniola at the North Hispaniola
Deformed Belt (Figure 2 - 2a to c; van Benthem et al., 2014; Symithe et al., 2015), with a possible slab-
gap under the Virgin Islands (Meighan and Pulliam, 2013).

A Dominican Republic D St Kitts Barbuda


0
0 -4
MT SF NHDB LAT
0 -4 0

Depth (km)
Depth (km)

A A’ D D’
0 300 400 0 300 400

B 0
E Martinique
MT -4 0
PRT LAT -4
0 0
Depth (km)
Depth (km)

B B’ E E’
0 300 400 0 300 400
0
C MT Virgin F Grenada Barbados Acc. Prism
-4 0
Islands PRT LAT
0 0 -4
Depth (km)
Depth (km)

C C’ F F’
0 300 400 0 300 400 600

Distance (km) Distance (km)

Figure 2 - 2: Geometry of the northern and eastern Caribbean subduction systems from Symithe et al. (2015).
Black line; plate interface between the North American and Caribbean plates. Dashed black line; underplating at
the Muertos Trough. Grey lines represent bathymetry at significant vertical exaggeration compared to the
earthquake depths, blue line is sea level. For locations and abbreviations see Figure 2 - 1.

As Cuba is welded to the North American plate, the plate boundary with the Caribbean plate in this
region is effectively marked by left-lateral transform faults bounding the Mid-Cayman Spreading
Center (MCSC) (Figure 2 - 1; Holcombe et al., 1973; ten Brink et al., 2002). Total short-term opening
rates for the MCSC based on geodetic results are 6 – 11 mm yr-1 (DeMets and Wiggins-Grandison,
2007), while GPS modelling constrains total spreading rates to 12.6 ± 0.6 mm yr -1 (Benford et al.,
2012b). This motion is partitioned along E – W (N080°E) trending faults that border the Cayman Trough.
The northern Oriente Fault undergoes 14.2 – 14.5 mm yr-1 left-lateral strike slip (Benford et al., 2012a),
the southwestern Swan Island Fault experiences 19.1 - 20.5 mm yr-1 left-lateral strike-slip (DeMets et
al., 2010), while the Walton Fault is subject to 9.3 – 8.9 mm yr-1 left-lateral slip (Symithe et al., 2015).
WNW-directed 20 mm yr-1 relative plate motion between South America and a fixed Caribbean plate

16
Chapter 2: Geological review of the Caribbean region

(DeMets et al., 2010) is partly accommodated by; 1) transtensional right-lateral strike-slip south of the
Leeward Antilles (Figure 2 - 1), and 2) ESE-directed subduction of the Caribbean lithosphere under
Venezuela and Colombia at the Southern Caribbean deformed belt (Trenkamp et al., 2002). The
western and eastern Caribbean plate are moving relatively to one another at 1 – 3 mm yr-1, which is
postulated to be accommodated at the Lower Nicaraguan Rise or Beata Ridge (Mattioli et al., 2014).

2.2 Origin of the Caribbean plate


There are two opposing groups of models for the origin of the Caribbean plate. The first group, known
as the ‘autochthonous’ models, argues that the Caribbean plate was created in-situ between the North
and South American plates (Figure 2 - 3a and b). The second group, known as the ‘allochthonous’ or
Pacific-origin models, argues that the Caribbean plate finds its origin in the Pacific Ocean (Figure 2 - 3c
to h). The first group of models imply that the crust of the Caribbean plate cannot be older than the
onset of spreading between the Americas, that the origin for the flood basalts of the Caribbean Large
Igneous Province (CLIP) lies between the American plates, and that the Caribbean plate experienced
little relative motion with respect to North and South America since its creation (Meschede and Frisch,
1998; James, 2006a; b). Recent studies have provided strong evidence against such an origin for the
Caribbean plate. The main arguments are (1) the age (195 Ma) of oceanic crust onshore Puerto Rico
pre-dates separation of the Americas by ~30 Myr (Pindell and Kennan, 2001a) and oceanic spreading
in the Atlantic by ~25 Myr (Bird et al., 2007), (2) the faunal assembly of Upper Jurassic cherts on La
Désirade Island is characteristic for Pacific-derived radiolarites, where Atlantic-derived radiolarites
would be expected for an in-situ origin (Mattinson et al., 2008; Montgomery and Kerr, 2009), (3) the
geochemical similarities between Upper Cretaceous CLIP-basalts and those associated with the
present-day Galápagos hotspot, located to the west of the Americas in the Pacific Ocean, are evidence
that the Caribbean plate has migrated since this time (Sen et al., 1988; Geldmacher et al., 2003) and
(4) a minimum of 1100 km (van Benthem et al., 2013) of west-directed subduction of Atlantic
lithosphere under the Aves Ridge and Lesser Antilles Arc since at least Late Cretaceous (Bouysse, 1988;
Neill et al., 2011), combined with passive margins on the eastern North and South American plates,
points to relative eastward motion of the Caribbean plate and not to a plate that remained largely
stationary through time and space.

The Pacific origin group of models involves a Pacific-derived Caribbean plate dragged into the gap
between the separating Americas (Pindell and Dewey, 1982) by means of slab-rollback of proto-
Caribbean crust, which subducts westwards below the Pacific Farallon lithosphere (Pindell and Kennan,
2009). This subduction system is responsible for creating the Great Arc of the Caribbean (Burke, 1988).

17
Chapter 2: Geological review of the Caribbean region

A North America B North America


GoM
GoM
Farallon Atlantic
Atlantic
Proto-
Farallon Caribbean
Proto-
Caribbean
130 Ma Africa Africa

South America South America


100 Ma

C D
Hauterivian-Albian (~135-99.6 Ma) subduction Turonian-Campanian (93.5-70.6 Ma)
reversal/initiation subduction reversal
9 Ma
Proto-Caribbean
33 Ma
55 Ma
72 Ma Great Arc of
84 Ma the Caribbean
90 Ma
100 Ma

Caribbean
oceanic plateau Caribbean
oceanic plateau South
America
Farallon Collision and obducting
The Great oceanic plateau material
Reversal/initiation Arc of the Plate
of subduction
Caribbean
E Foothills
F
100° 90° suture 70° 60° 50° 40° 100° 90° 80° 70° 60° 50° 40°

East dipping subduction North America plate Guerrero North American 90 Ma


following arc accretion V 144 Ma suture
plate
boyste

Gulf of Mexico
s
rd m
er

30° 30° 30° 30°


V
rif

Farallón
t

plate V
Maya
c
Ar

20° V 20° 20° block


Maya Farallón
ro

block proto-Caribbean
plate proto-Caribbean
rre

V spreading
Ca

Chortis
Sea
ue

block
rib
-G

10° Nicaragua 10° 10° 10°


be

Mezcalera
an
an

plate South
ar
be

South American Galápagos American


rib

0° 0° 0° 0°
Galápagos CLIP
plate
ophiolites plate
(95 Ma) back-arc
Ca

Cuba (present day position) present day extent


110° 100° 90° 80° 70° 60° 50° 40° 110° 100° 90° 80° 70° 60° 50° 40°

G 141 Ma
H 130 Ma
NORTH
*

NORTH
AMERICA AMERICA
*

(fixed) Proto- Pac/NA (fixed)


*

?
Extinct spreading pat- Caribbean
*
*

Mexican tern in Gulf of Mexico seaway Mexican Extinct spreading pat-


*
*

backarc Bahamas
basin ? backarc tern in Gulf of Mexico
*

Bahamas
*

Pac/NA hot spot


Colombian hot spot basin
*

Marginal
* * * *

seaway
*

Chortís
*

SOUTH
Chortís SOUTH AMERICA
Arc/NA
*

AMERICA These HP/LT and


arc rocks will
Arc/NA
*

NA become those of NA Arc Pac


Siuna and Guate-
Arc
Pac mala, including *
Int Antio- (stretching) SA rocks older than * Antio- SA
er quia
Pac/Arc
tra -Ame quia 135 Ma Pac/Arc * 1000 km
nsf ric 1000 km These HP/LT and arc rocks will
orm an become those of the Antilles (Cuba,
*

Ecuadorian
*

Jamaica), younger than 135 Ma backarc basin


* *

Ecuadorian
These HP/LT and arc rocks will Arc/SA
*

backarc basin
become those of the Predicted position of HP-LT rocks
Arc/SA
*

Predicted position of HP-LTrocks Quebradagrande belt, including * * * Predicted position, magma arc rocks
* *

rocks older than 135 Ma


Pac/SA * * Predicted position, magmatic arc rocks Area of known/inferred oceanic crust
Area of known/inferred oceanic crust Pac/SA
*

Figure 2 - 3. A and B; in-situ model after Meschede and Frisch (1998). C and D; subduction polarity reversal models
after Pindell et al. (2006) (C) and Kerr et al. (2003) (D). E and F; continuous west-dipping subduction after Mann
et al. (2007a). G and H; continuous west-dipping subduction after Pindell et al. (2012).

18
Chapter 2: Geological review of the Caribbean region

The earliest Pacific origin models require a reversal of subduction zone polarity within the Great Arc
(Figure 2 - 3c and d). By their account, the Farallon plate was subducting eastwards below the Americas
during the Triassic (?) to middle Jurassic. This subduction continued during separation of the Americas,
which was accompanied by Proto-Caribbean seafloor spreading in the Jurassic. During the Cretaceous,
this eastward subduction was halted and flipped to westward subduction of Proto-Caribbean
lithosphere under the Farallon/Pacific plate (Burke, 1988; Sinton et al., 1998; Kerr et al., 1999; Lewis
et al., 2002; Pindell et al., 2005; Hastie et al., 2013). These models rely on abrupt and regional changes
from primitive to evolved island arc volcanism (Lebrón and Perfit, 1994) and changes in the
geochemistry of the volcanic rocks (Hastie et al., 2013). With increased sampling over time throughout
the Caribbean, the initial temporal gaps are slowly filled, and compilations of island-arc related volcanic
rocks (Kerr et al., 2003; Boschman et al., 2014) now lack an arrest in volcanism in which to place the
subduction polarity reversal.

An intermediate Pacific origin model also exists, which is not a subduction reversal model sensu stricto,
but essentially a double-arc collision model. It relies on a far-travelled west-dipping subduction system
located within the Farallon plate (Caribbean – Guerrero arc of Mann et al., 2007a), which collides with
and usurps the older east-dipping subduction system (Ratschbacher et al., 2009) in Albian – Aptian
times (Escuder-Viruete et al., 2013). The difficulty with this model is the lack of arc magmatism in the
Great Arc pre-dating the Hauterivian (Mitchell, 2003; Rojas-Agramonte et al., 2006; Hastie et al., 2009;
Stanek et al., 2009), which would be expected if east-directed subduction of the Farallon plate
continued during the separation of the Americas.

Another set of Pacific origin models exists that do not require a subduction zone polarity reversal.
These models argue that spreading between the Americas was accommodated by a transform
boundary that separated the Proto-Caribbean lithosphere in the east from the Farallon plate in the
west (Pindell and Kennan, 2009). West-dipping subduction initiated at this transform boundary at
around 135 Ma, thereby accounting for all of the volcanism observed in the Great Arc (Pindell et al.,
2012; Boschman et al., 2014). The lack of an observation is however not a very strong argument, but
until pre-Hauterivian arc-related magmatic rocks are discovered along the Great Arc, the continuous
west-dipping subduction system with initiation at around 135 Ma is the most likely model.

19
Chapter 2: Geological review of the Caribbean region

2.3 Caribbean crustal structure


The interior of the plate is characterized by Jurassic oceanic crust overlain by Upper Cretaceous plateau
basalts (Edgar et al., 1973), known as the Caribbean Large Igneous Province (CLIP) (Burke, 1988).

The oceanic domain of the Caribbean consists of two main types of crust. The first type is an
anomalously thick oceanic crust overlain by a smooth B” seismic reflector, which represents the
Caribbean Large Igneous Province (CLIP). A thin crust with rough B” seismic reflector is associated with
normal MORB-type crust, which is mainly encountered in deeper parts of the Caribbean Ocean
(Diebold et al., 1981; Donnelly, 1994; Mauffret and Leroy, 1997; Mauffret et al., 2001). Oceanic crust
in the Cayman Trough results from post-49 Ma ultra-slow spreading at the Mid-Cayman Spreading
Center (Leroy et al., 1996, 2000; Hayman et al., 2011). An E – W cross section through the Caribbean
is shown in Figure 2 - 4, while a map displaying the thickness of the crust in the Caribbean and the
location of the cross section is shown in Figure 2 - 5. The stratigraphy of the igneous rocks is shown in
Figure 2 - 6, while their geographic extend is shown in Figure 2 - 7 for N-MORB and CLIP and Figure 2 -
8 for the IAT and CA series. sin
se
se

Ba
Ri
Ri

an
an

an

sin
en

t
en
gu
gu

W E
el
sin
m

Ba

n
m

zu
ra

le
ra

ea
rp

Ba

ge
rp

an

til
ne
ca
ica

e
ca

Oc
dg
b-
ca

An
id
Ni

el
Ve
rN

Es

Su

aR
Es

zu

ic
Ri
er

er
rth
o
pe

nt
ne
ss

es
at
iti
w
dr

ss

la
Up

He

No
Lo

Ha

Be

Av
Ve

Le
Pe

At

Km
Gabbro Volcano

10
Continental Picrites and cumulates
Crust 7
20
Sub-B” Oceanic
Plateau basalts Oceanic Crust
Crust 30
0 100 500 km

Figure 2 - 4: From Mauffret and Leroy (1997). Schematic W – E cross section through the Caribbean domain. See
Figure 2 - 5 for location.

Oceanic crust in the interior Caribbean plate characterized by acoustic basement with a rough B”
reflector has a crustal thickness, current depth, and accumulated sediment thickness indicative for
Jurassic crustal age (Mauffret and Leroy, 1997). Based on magnetic anomalies, spreading in the
Farallon/Caribbean plate continued through the Early Cretaceous (Maus, 2010; Guevara et al., 2013),
although normal MORB Farallon-derived fragments of this age are not encountered onshore. There is
no evidence for spreading in the Caribbean plate following emplacement of the CLIP, apart from the
pull-apart Cayman Trough and Yucatan Basin, and the fore-arc Grenada Basin.

20
Chapter 2: Geological review of the Caribbean region

ll /

sin
(* we

Ba
r
es B

re m
rio

da
pl / P

go
an ate )

/G e
ca Tro inte

na
e yst

ba
rib sam 65

d /
an les
To
dg s
Ca ed P 1

n gh

Ri illes

l
ad ti
s
n
g D

id A n
pl

co
s
ed / O

an

nt
a
Ba
Ca ean

ol

Ri

in d
rA
I sl
ca
Dr 15

Tr ar
i
an

to
ta

se
in
b

ai

ew
es
ym
DP

er
sp
ba
m

rg

s
Ma

Av
Le

Le
Pu
Hi
Ja

Cu

Vi
DS

Yu
0 Pleist. Q
Pliocene IAT island-arc [extrusive]
Neogene

10 IAT island-arc [plutonic]


Mio. IAT SSZ ophiolites
20 CA island-arc [extrusive]
Oligo. CA island-arc [plutonic]
30 CA SSZ ophiolites
? Oceanic Plateau basalts
Paleogene

40 Oceanic crust [MORB]


Eoc.
50
?
HP mélange formation
60 Paleoc.
* ?

Maast. Caribeana subduction


70
Late Cretaceous

Camp. *
80
Sant. *
Con.
90 Tur. CLIP main phase
Cen.
100
Albian ?
110
Early Cretaceous

Aptian ? ?
120
Barr. ?
130 Haut ? Onset Great Arc subduction
Valan.
140
Berr.
Torth. ?
150
Kimm.
Jurassic

160 Oxford.
Call.
Bath. 195
170 Baj. Ma
Figure 2 - 6: Compilation of ages and settings for intrusive and extrusive rocks, modified from Boschman et al.
(2014). For <65Ma only island-arc related or MORB units shown. Dredge samples from Mauffret et al. (2001). HP
mélange formation from Boschman et al. (2014). Intrusives are from Kerr et al. (2003). Other references in text.

The western part of the Yucatan Basin is underlain by a rectangular NNE – SSW striking segment of
Paleocene – middle Eocene oceanic crust, which originated as a pull-apart basin from left-lateral
motion between Cuba and the Belize margin (Rosencrantz, 1996). It is bounded on both sides by
currently inactive left-lateral transform faults (Gordon et al., 1997; Cruz-Orosa et al., 2012) that extend
onshore Cuba (Rosencrantz, 1990). Oceanic crust formed as early as 49 Ma in the ultra-slow spreading
Cayman Trough (Leroy et al., 2000). Between late Oligocene and early Miocene the Mid Cayman
Spreading Center (MCSC) propagated southwards by ~30 km resulting in creation of the Swan Island
restraining bend and the Hendrix pull-apart basin, the latter in combination with activity on the Walton
fault (Leroy et al., 2000).

Abnormally thick (15 – 20 km) crust with oceanic-plateau affinities is observed both offshore and
onshore. In the offshore domain this crust is characterized by a smooth B” seismic acoustic basement
reflector (Diebold et al., 1981; Donnelly, 1994; Mauffret and Leroy, 1997). Deep sea drilling (DSDP and
ODP) penetrated smooth B” basement and encountered basalts and dolerites intercalated with
sediments of Turonian to Santonian (94 – 84 Ma) age (Sinton et al., 1998). These ages are associated

22
Chapter 2: Geological review of the Caribbean region

2.4 Mesozoic evolution of the Caribbean realm


Geological evidence in the Caribbean region points strongly towards a Pacific-origin for the Caribbean
plate (Kerr et al., 2003; Pindell et al., 2005; Pindell and Kennan, 2009). Without the geochemical
evidence for a subduction zone polarity reversal (Kerr et al., 2003; Boschman et al., 2014) and with the
arguments against a far-travelled Farallon-Guerrero Arc (Pindell and Kennan, 2009), the model
assuming an inverted Inter-American transform system (Pindell et al., 2012) resulting in continued
west-dipping subduction from 135 Ma onwards appears the most viable and is explored here, with a
focus on the northern Caribbean realm. The main events are highlighted, and for a full discussion the
reader is referred to Mann et al. (2007a), Pindell and Kennan (2009), Pindell et al. (2012) and Boschman
et al. (2014). The reconstructions of Figure 2 - 9a to d are from Pindell and Kennan (2009) and are
shown in the Indo-Atlantic hotspot reference frame.

-100° -90° -80° -70° -60° -50° -40° -100° -90° -80° -70° -60° -50° -40°
30° 30° 30° 30°

A GOM = Gulf of Mexico


YUC = Yucatan V
N. Am. Plate 125–120 Ma B G N. Am. Plate 100 Ma
CA = Caribbean
NA = North America V
SA = South America Proposed
V
CHO = Chortis Farallon- Guerrero V
20°
JAM = Jamaica 20° 20° Caribbean Arc 20°

HPR = Hispaniola - Puerto Rico subduction


V
SPBR = Southern Peninsula / Beata Ridge zone Inverted
V
AVE = Aves Ridge CHO rifts
MAR = Maracaibo block FA/CA Caribeana
ZIH GOM V B
V CA/NA
10° In all maps, area V 10° 10° 10°
G Granitoids to be subducted V Mesquito V
? Proto-Car.
between this and Proto-Car.
V Volcanic next younger map B YUC V CA/SA Plate
V CHO Plate
B HP/LT is shown grey V Caribeana eastern SLAB
CA/NA Jamaica? SPBR
GAP
JAM V CA/NA Galapagos
0° 0° 0° V 0°
FA/CA
Proto-Car. HS today Caribbean
? CUBA
Farallon Caribbean V Plate
Galapagos Plate MAR V
Plate Plate HPR Margarita
HS today NW V G
Carib. V HPLT core
AVE
NW-Carib. Arc? SPBR
-10°
subduction V
-10° -10° -10°
zone CA/SA V
Early deformation S. Am.
Early deforma- pre-CLIP
G CA/SA Plate
tion pre-CLIP V
S. Am. 1000 km
1000 km
V Plate
-20° -20° -20° -20°
-100° -90° -80° -70° -60° -50° -40° -100° -90° -80° -70° -60° -50° -40°

-100° -90° -80° -70° -60° -50° -40° -100° -90° -80° -70° -60° -50° -40°
30° 30° 30° 30°

C Guerrero
V N. Am.
Plate
84 Ma D Guerrero Arc
V
Carib. Arc
collision
N. Am.
Plate Ophiolites
71 Ma
Arc
V Onset emplaced
shortening FA/NA V
Nascent
CHO
Proto-
20° CHO
V 20° 20° Siuna Caribbean 20°

V
Caribeana Proto-Car.
V

Mesquito
10d inversion
zone
Plate V
V SPBR
V FA/CA CA/NA SLAB
Mesquito
V
GAP
Rough
10° CA/NA SPBR SLAB
GAP
G
basement
(no LIP)
10° 10° V
10b/c
Caribbean ABC
CA/SA
G ??
10°

Caribbean V
Plate G Arc established V Plate
during Late Cret.
CA/SA G V
? V
Galapagos FA/CA ABC Isl. V Galapagos
0° HS today V G 0° 0° 0°
HS today V
? G
Grey ellipse: Possible Grey ellipse: Possible V
V
Position of Galapagos V G Position of Galapagos
hot spot CA/SA hot spot S. Am.
-10° -10° -10°
FA/CA
Plate -10°
Onset of subduction,
V
S. Am.
10a
V
arc volcanism in Plate
interval 84-71 Ma V
1000 km 1000 km
FA/SA V
-20° -20° -20° -20°
-100° -90° -80° -70° -60° -50° -40° -100° -90° -80° -70° -60° -50° -40°

Figure 2 - 9: Palinspastic reconstruction of the Caribbean realm at 125 – 120, 100, 84 and 71 Ma, from (Pindell
and Kennan, 2009). Blue lines are location of cross sections in Figure 2 - 10a to d.

The spreading centers in the Colombian Marginal seaway and the proto-Caribbean seaway merged by
135 Ma, forming the proto-Caribbean plate (Pindell et al., 2012). This timing coincides with the onset
of west-dipping subduction of proto-Caribbean lithosphere under the Farallon plate (Rojas-Agramonte

25
Chapter 2: Geological review of the Caribbean region

et al., 2011; Boschman et al., 2014). This subduction resulted in island arc tholeiite (IAT) volcanism
which created the Great Arc of the Caribbean (Kerr et al., 2003; Hastie, 2009).

During the remainder of the Early Cretaceous the Caribbean part of the Farallon plate became
progressively engulfed in the widening gap between the Americas (Figure 2 - 9a and b; Figure 2 - 10a;
Pindell et al., 2005). Relative eastwards motion with respect to the Americas was accommodated by
right-lateral strike-slip along the Columbian margin and left-lateral strike-slip along the Siuna terrane,
which effectively delineated the Caribbean plate as a separate entity from the Farallon plate. Within
the newly formed Caribbean plate strike-slip occurred between the arc segments of Jamaica, Cuba,
Hispaniola and Puerto Rico (Pindell and Kennan, 2009). Relative motion of South America with respect
to North America was towards the southeast from 135 Ma until 100 Ma. This was followed by strong
southward divergence between 95 and 85 Ma (Somoza and Zaffarana, 2008).

Subduction of the Proto-Caribbean spreading ridge under the Farallon plate resulted in a progressively
widening slab window (Figure 2 - 9b to d; Pindell et al., 2006). This slab window possibly acted as a
gateway for the mantle plume that fed the flood basalts of the CLIP, which resulted in contamination
of the Aptian – Albian ophiolites in Cuba and Hispaniola with a plateau basalt geochemical signature
(Pindell and Kennan, 2009). Plateau basalt magmatism of the CLIP lavas (90 – 85 Ma) in Curaçao (Sinton
et al., 1998; Kerr et al., 2003) however do not show such a signature (Hastie and Kerr, 2010).

Counterclockwise rotation of Farallon – Caribbean motion from 100 Ma onwards resulted in an


inversion of the Costa Rica – Panama intra-oceanic transform fault to an east-dipping subduction zone.
Arc volcanism in Central America was thereby initiated between 84 – 71 Ma, which delineated the
western boundary of the Caribbean plate (Pindell and Kennan, 2009).

The Caribeana terrane on the Proto-Caribbean plate entered and blocked the subduction channel of
the Great Arc of the Caribbean during the Campanian (Figure 2 - 9c; Figure 2 - 10b; García-Casco et al.,
2008). The resulting southern trench jump in Cuba caused obduction of the Mayarí-Baracoa ophiolite
(Boschman et al., 2014). The subducted Caribeana terrane was later exhumed as HP – LT metamorphic
units on Cuba (Cangre, Piños, Escambray, Asunción) and Hispaniola (Samaná) (García-Casco et al.,
2008).

26
Chapter 2: Geological review of the Caribbean region

A S-SW pre-Late Campanian N-NE


Caribbean plate Proto-Caribbean (North American plate)
Central American trench
Antillean Antillean trench
Central American Cretaceous
Cretaceous Arc Arc Oceanic crust and
Caribeana Bahamian and Mayan borderlands
CLIP
plutonism
distal sedimentary sections
serpentinite-mélanges (Placentas, Rosario North, Quiñones
with HP blocks
(subduction channel)
B Late Campanian

Escambray (unit I) Bahamian / Mayan


arrest of volcanic arc activity Eastern Yucatan borderlands
CLIP proto-caribbean
future
detachment Medium- Escambray (unit II)
Subduction-accretion of Caribeana high grade Pinos
Arrest of arc magmatism Sierra Verde
Escambray (unit III) Samana
(gneises and eclogites) Escambray and Samana eclogites
C Maastrichtian (central Cuba - Yucatan basin segment) ophiolites
Cayman ridge - Sierra Maestra (central Cuba) Bahamian / Mayan
(arc/back arc fragments) borderlands
CLIP proto-caribbean
Low-angle detachment as Yucatan basin
begins to open
Escambray
Pinos
footwall core-complexes
First formation of syntectonic basins

D Maastrichtian - Eocene (Hispaniola - Puerto Rico segment)


Hispaniola - Puerto Rico
arc Hispaniola - Puerto Rico Bahamas
trench (collides with trench
in Eo-Oligocene times)
CLIP n
a
be
Continued subduction and accretion to forearc.
a rib Samana (marbles, phyllites)
Normal subduction. Synsubduction exhumation. -c Samana (eclogites)
Final exhumation during the Eocene - Oligocene to
p ro
(= collision of Bahamas banks)

Figure 2 - 10: Paleotectonic cross sections through the Caribbean, from (García-Casco et al., 2008). Location of
cross sections in Figure 2 - 9.

The western Cuban segment collided with Yucatan during the Campanian (Figure 2 - 9c). The
Guaniguanico terrane in northwestern Cuba closed and a left-lateral transform margin developed at
Belize (Pindell et al., 2005). This collision resulted in back-thrusting along the Pedro escarpment and
northward underthrusting beneath the Siuna, Nicaraguan Rise and Jamaica. Arc-volcanism continued
through the early Eocene (Pindell and Kennan, 2009). Tomographic imaging in this area does not reveal
a NW-dipping slab and therefore the amount of underthrusting must be limited (van Benthem et al.,
2013).

The Chortis and Siuna terranes became dislodged from the North American Maya block in the
Maastrichtian - Paleocene and began to move as an independent terrane towards the east along the
Motagua – Polochic fault system (Figure 2 - 9d; Harlow et al., 2004). Rollback of the Proto-Caribbean
slab resulted in a NW-directed motion of Cuba and termination of arc-related volcanism in western
Cuba (Pindell et al., 2005). A relatively flat slab characterized subduction under central Cuba in the
Maastrichtian, while subduction under Hispaniola – Puerto Rico was characterized by normal-angle
subduction accompanied by syn-subduction exhumation (Figure 2 - 10d; García-Casco et al., 2008).

27
Chapter 2: Geological review of the Caribbean region

2.5 Seismic tomography and paleomagnetism in the northern Caribbean


Slab remnants of the Cretaceous – Eocene Great Arc of the Caribbean (GAC) subduction are found at
around 1000 km depth in the lower mantle to the west of the Eocene – present Lesser Antilles (LA)
slab, following slab detachment during the Eocene (Figure 2 - 11; van Benthem et al., 2013). The
northern Lesser Antilles (nLA) slab is connected to the North American plate and is highly obliquely
subducting under Puerto Rico and eastern Hispaniola (Calais et al., 1992; van Benthem et al., 2013). In
Hispaniola this convergence is accommodated by oblique thrusting at the North Hispaniola Fault, while
strike-slip is the dominant process along the OFZ, SFZ, EPGF and diffuse Hispaniola boundary fault
zones (Benford et al., 2012a). To the west of the Cordillera Oriental earthquake depth diminishes
rapidly, marking the edge of the slab (Molnar and Sykes, 1969), where deformation induced by the
west-moving nLA is taken up aseismically (van Benthem et al., 2013; Symithe et al., 2015). Northward
thrusting of Caribbean lithosphere under eastern Hispaniola to depths of <100 km is accommodated
at the Muertos Trough (Granja Bruña et al., 2009; ten Brink et al., 2009; Llanes Estrada et al., 2012; van
Benthem et al., 2013).

Figure 2 - 11: Slabs in the northeastern Caribbean realm derived from seismic tomography, from van Benthem
et al. (2013). Abbreviations for the slabs: Mu = Muertos Trough, nLA = northern Lesser Antilles, sLA = southern
Lesser Antilles, sGAC = southern Great Arc of the Caribbean, nGAC = northern Great Arc of the Caribbean, SC =
Southern Caribbean.

28
Chapter 2: Geological review of the Caribbean region

Paleomagnetic studies on Cretaceous and Cenozoic units throughout the northern Caribbean reveal
counterclockwise (CCW) rotations following their deposition. CCW rotations are encountered on Cuba
(Renne et al., 1991; Chauvin et al., 1994; Alva-Valdivia et al., 2001; Tait et al., 2009), Jamaica (Watkins
and Cambray, 1970; Guja and Vincenz, 1978) and Puerto Rico (Reid and Plumley, 1991), which have
been related to strike-slip processes (Tait et al., 2009). CCW rotations are also observed in Hispaniola,
for the Cordillera Central (Vincenz and Dasgupta, 1978) and the Beloc region in the Massif de la Selle
(33° CCW since 60 Ma, Fossen and Channell, 1988). The only clockwise rotations are noted in the Massif
du Nord, close to Ennery (37° CW since 50 Ma, Fossen and Channell, 1988). The latter authors
calculated an Eocene paleolatitudinal separation between these terranes of 8° (or ~900 km). Taking
their current position into account, this indicates around 750 km of latitudinal convergence since
middle Eocene times. These values can be questioned based on the statistical significance of their data
(van der Boon, 2015, personal communication). Around 500 – 700 km of Eocene and Oligocene ~E –
W convergence between the Nicaragua Rise/southwest Hispaniola and northeast Hispaniola is
nonetheless required based on paleotectonic reconstructions (Sykes et al., 1982; Müller et al., 1999;
Boschman et al., 2014). Such a convergence can be accommodated by subduction under Hispaniola or
at the Nicaraguan Rise, or by strike-slip processes.

29
Chapter 2: Geological review of the Caribbean region

2.6 Accepted chapter for Elsevier book publication “Transform plate boundaries
and fracture zones”, edited by João C. Duarte

CHAPTER

Strike-slip fault systems along the northern Caribbean plate boundary

Richard J.F. Wessels1, 2

1. Sorbonne Université, CNRS-INSU, Institut des Sciences de la Terre Paris, ISTeP UMR 7193, F-

75005 Paris, France

2. IFP Energies nouvelles, Geosciences, 1&4 Avenue du Bois-Préau, 92852 Rueil-Malmaison,

France

e-mail address corresponding author: rjfwessels@gmail.com

NON-PRINT ITEMS

Key Words

Northern Caribbean; Hispaniola; strike-slip faults; seismic activity; earthquakes; plate boundary

30
Chapter 2: Geological review of the Caribbean region

CHAPTER STARTS HERE

Abstract

This chapter focusses on the strike-slip fault systems along the northern Caribbean plate boundary.
First, the present-day geodynamic setting of the northern Caribbean realm is described to comprehend
the partitioning of relative motion between the Caribbean and North American plates. Next, the
geological evolution of the region is examined to understand when and why the different faults
became active. The two main plate bounding strike-slip fault systems are addressed in detail. The focus
of the chapter then shifts towards historical and recent earthquakes on these strike-slip faults in
Hispaniola. Lastly, the age for initiation of the southern fault system and the way in which convergence
in western Hispaniola is accommodated are discussed.

1. Introduction

The northern Caribbean plate boundary is a complex zone where relative east-northeastward motion
between the Caribbean and North American plates is accommodated (figure 1a; DeMets et al., 2010).
The geological setting and geodynamic evolution of the western part of this region is strongly
controlled by two strike-slip fault systems (figure 1b; Burke et al., 1978). These two seismogenic fault
systems do not only bound a spreading center and oceanic lithosphere (Rosencrantz and Mann, 1991),
but also transect and displace lithosphere affected by arc volcanism and flood basalts (Mann and
Burke, 1984). The evolution of both fault systems through time and space is described below, to better
understand their formation, geometry and seismicity.

The aim of this chapter is to provide an overview of the northern Caribbean realm, with a specific focus
towards the strike-slip systems, which is concise in volume, but comprehensive in content. This chapter
offers a regional framework on which to apply knowledge from other strike-slip systems and could
provide analogues for other regions. Last but not least I hope to spark enthusiasm amongst the
scientific community about resolving the questions that still exist regarding the evolution of Hispaniola.

31
Chapter 2: Geological review of the Caribbean region

of plate fragments bounded by a spreading center, two strike-slip systems and two
subduction/underthrusting systems (figure 1b).

There is presently no consensus on how to label the plate fragments that occupy the northern
Caribbean plate boundary zone, which are interchangeably termed microplates, blocks or tectonic
slivers (Manaker et al., 2008; Benford et al., 2012a; Symithe et al., 2015; Calais et al., 2016). In the
remainder of this chapter the Gonâve fragment will be labelled a microplate, while the Hispaniola,
Puerto Rico – Virgin Islands and North Hispaniola fragments will be labeled blocks. The South Jamaica
block of Benford et al. (2012a) will not be addressed in detail, since its existence and geographical
extend are uncertain (Symithe et al., 2015; Calais et al., 2016). The outline and existence of the
different blocks is dependent on the way the GPS modeling is performed, as are the resulting relative
velocity vectors. The velocity vectors in figure 1b are taken from Symithe et al. (2015) and Calais et al.
(2016) .

The Cayman Trough is a 110 by 1400 km oceanic basin, separated in an eastern and western part by
the Mid-Cayman Spreading Center (MCSC) (figure 1b; Bowin, 1968). Seafloor spreading at the MCSC
occurs with total spreading rates of 12.6 ± 0.6 mm yr-1 (Benford et al., 2012b). The southern boundary
of the western Cayman Trough is the left-lateral Swan Island Fault Zone (SIFZ) (figure 1b; Mann et al.,
1991). The eastern part of the Cayman Trough is bounded by strike-slip faults (Molnar and Sykes,
1969); the northern Oriente Fault Zone (OFZ) and the southern Walton Fault Zone (WFZ) (figure 1b;
Leroy et al., 1996). The MCSC, WFZ and OFZ define the western boundaries of the Gonâve microplate.

The northern strike-slip system runs to the north of the Gonâve microplate and Hispaniola block. It
consists of the Oriente Fault Zone in the west and the Septentrional Fault Zone (SFZ) in the east (Mann
and Burke, 1984; Heubeck and Mann, 1991a). This northern strike-slip system separates Gonâve and
Hispaniola from the North Hispaniola block that itself is separated from the North American plate by
the thrust-related North Hispaniola Deformed Belt (NHDB) (figure 1b).

The southern strike-slip system runs south of the Gonâve microplate (figure 1b). It consists of two
parts, the Walton Fault Zone (Leroy et al., 2000) in the west and the Enriquillo – Plantain Garden Fault
Zone (EPGFZ) (Mann et al., 1984) in the east. These segments are connected via the Jamaican
restraining bend (Mann et al., 1985, 2007b). The border between the Gônave microplate and the
Hispaniola block is diffuse and could coincide with the Haitian fold-and-thrust belt (HFTB) (Benford et
al., 2012a), the Plateau Central – San Juan Basin in central Hispaniola (Calais et al., 2016), or be a broad
zone of deformation between Haiti and Hispaniola (Symithe et al., 2015).

The northern subduction system is the North Hispaniola Deformed Belt (NHDB) – Puerto Rico Trench
(PRT) system, which accommodates south-dipping subduction of the North American plate under the

33
Chapter 2: Geological review of the Caribbean region

Caribbean plate (figure 1b). In the east, the highly oblique subducting slab is attached to the North
American plate (van Benthem et al., 2013). In the west the North American slab is detached (van
Benthem et al., 2014) and convergence with the Bahamas platform is accommodated at the North
Hispaniola Deformed Belt (NHDB) (Dillon et al., 1992). The location of the slab edge roughly coincides
with the eastern end of the Bahamas Platform (van Benthem et al., 2014), at the boundary between
the North Hispaniola – PRVI blocks from Calais et al. (2016).

Hispaniola and Puerto Rico are back-thrusted over the Caribbean plate at the Muertos Trough (MT),
which is considered to result as a response to subduction of the North American plate (figure 1b; Mann
et al., 2002; ten Brink et al., 2009). Seismic tomography across the MT shows north-dipping Caribbean
lithosphere to depths of around 100 km (van Benthem et al., 2013), which is supported by regional
seismicity (Granja Bruña et al., 2010; Symithe et al., 2015).

From a plate tectonic point of view the Oriente – Septentrional and Swan Island fault systems, together
with the MCSC, are sufficient to accommodate relative motion between the North American and
Caribbean plates. If this is the case then the other boundaries of the Cayman Trough, being the western
Cayman Ridge and the WFZ – EPGFZ could be expected to act as passive, non-seismogenic fracture
zones. The northwestern boundary to the Cayman Trough, close to the Cayman Ridge, is indeed a
passive, non-seismogenic fracture zone. The southeastern boundary to the Cayman Trough is however
not a passive system. This is evidenced by active left-lateral displacement and associated seismicity
along the WFZ – EPGFZ segments.

2.2 Evolution of the northern Caribbean realm

Two different models have been used to explain the evolution of the Caribbean. The first one is the
‘in-situ’ model, which assumes an autochthonous, stationary origin for the Caribbean plate (Meschede
and Frisch, 1998; James, 2006a; b). The second, more commonly accepted model, is the ‘Pacific’ origin
model (e.g. Mann, 2007 and Pindell and Kennan, 2009).

According to these ‘Pacific’ origin models, the Caribbean plate originated as part of the Farallon plate
in the Pacific Ocean (Pindell and Dewey, 1982). Proto-Caribbean lithosphere formed in the Mesozoic
during the separation of the North and South America plates (Pindell and Kennan, 2001a). This
lithosphere subducted westwards under the Farallon plate in Cretaceous times and formed the
volcanic Great Arc of the Caribbean. Rollback of the proto-Caribbean slab dragged the Caribbean
portion of the Farallon plate between the diverging North and South American plates. This eastwards-

34
Chapter 2: Geological review of the Caribbean region

directed motion was accommodated by strike-slip faulting along the northern and southern plate
boundaries (Pindell et al., 2012; Boschman et al., 2014).

Several variations to this ‘Pacific’ origin model exist. Some invoke a flip of initial eastward to westward
dipping subduction at the Farallon – proto-Caribbean boundary (Burke, 1988; Sinton et al., 1998; Kerr
et al., 1999; Lewis et al., 2002; Pindell et al., 2005; Hastie et al., 2013), while others postulate a double-
arc collision model where a far-travelled west-dipping subduction system collides with an older east-
dipping subduction system (Mann, 2007; Ratschbacher et al., 2009; Escuder-Viruete et al., 2013). All
of the ‘Pacific’ origin models have a west-dipping subduction system established by the end of the
Cretaceous. A detailed discussion of these different systems is beyond the scope of this chapter, but
the reader may refer to the articles by Pindell and Kennan (2009), Mann (2007), Boschman et al. (2014),
and references therein.

2.2.1 Cretaceous
Two main volcanic events occurred during the Cretaceous in the northern Caribbean. The first event is
the construction of the Great Arc of the Caribbean (Burke, 1988) due to subduction at the Farallon –
proto-Caribbean plate boundary. Volcanism evolved from island arc tholeiite (IAT) to calc-alkaline (CA)
(Kerr et al., 2003; Hastie, 2009). The associated intrusive and extrusive rocks are found on the Greater
and Lesser Antillean islands (Mann and Burke, 1984). The second event is the extrusion of plume-
related flood basalts onto the Caribbean plate, forming the Caribbean Large Igneous Province (CLIP)
(Edgar et al., 1973). The main extrusive phase occurred around 90 Ma (Kerr et al., 2003). The crust
under the CLIP is anomalously thick up to 15 – 20 km (Burke et al., 1978; Mauffret and Leroy, 1997).
The CLIP can be found throughout much of the interior of the present-day Caribbean plate and on
some of the Greater Antilles islands (Mauffret and Leroy, 1997; Kerr et al., 2003).

At the end of the Cretaceous a left-lateral transform margin developed at Belize as a result of the
collision between the western Cuban segment of the Great Arc and eastern Yucatan (figure 2a/c;
Pindell et al., 2005, 2006). During this time the Caribbean plate was moving towards the NE relative to
North America (figure 2c; Pindell and Kennan, 2009).

35
Chapter 2: Geological review of the Caribbean region

2.2.2 Cenozoic
Opening of the Yucatan pull-apart basin resulted from tearing of the proto-Caribbean slab along the
Belize margin in the Paleocene (figure 2a/d; Rosencrantz, 1990). Subduction of the proto-Caribbean
ceased as a result of the collision between the Great Arc of the Caribbean and the Bahamas platform.
This diachronous collision, from west to east, started in the Paleocene (Gordon et al., 1997) and ended
arc volcanism. During the middle to late Eocene subduction and collision had terminated in central
Cuba (van Hinsbergen et al., 2009) and arc-related volcanism ceased (García-Casco et al., 2008).

Pull-apart tectonics in the Yucatan Basin terminated due to the collision of the Great Arc with the
Bahamas. This collision resulted in a change of relative motion from NNE (Maastrichtian) to E (middle
Eocene) of the Caribbean plate with respect to the North American plate (Pindell et al., 2005). This
change in relative motion initiated a stepwise jump of the northern Caribbean plate boundary (figure
2b). This boundary first coincided with the eastern Yucatan transform margin. It then jumps by
subsequent initiation of the Pinar (Paleocene), La Trocha (Early Eocene) and Cauto (middle Eocene)
faults on Cuba, to eventually the Oriente Fault (early Oligocene) south of Cuba (Leroy et al., 2000;
Rojas-Agramonte et al., 2006; Cruz-Orosa et al., 2012). This process transferred Cuba to the North
American plate. Activity on the Cauto and subsequently the Oriente Fault was accompanied by pull-
apart extension in the Cayman Trough followed by seafloor spreading from 49 Ma onwards (figure 2e;
Leroy et al., 2000). The effects of rifting during the Paleocene to early Eocene are evident from
transtensional half-grabens in Jamaica (Mann and Burke, 1990; Draper, 2008) and tilted fault blocks
on the Gônave microplate (figure 3b; Leroy et al., 1996; Corbeau et al., 2016a).

Collision between Hispaniola and the Bahamas started during Eocene – Oligocene times (figure 2f;
Escuder-Viruete et al., 2011). Convergence between North and South America accelerated during the
Oligocene (Somoza, 2007). Between 26 and 20 Ma the Mid-Cayman Spreading Center grew southward
by about 30 km. This formed the Swan Island restraining bend, the Hendrix pull-apart basin, and
initiated the Walton Fault (Leroy et al., 2000). The Southern Haiti – Beata Ridge block possibly collided
with the northern island-arc terrane in the early Miocene (Mauffret and Leroy, 1999),
contemporaneous with the creation of a fold-and-thrust belt in central Haiti (figure 2g; Pubellier et al.,
2000). Hispaniola starts to separate from Cuba along the Septentrional – Oriente fault system and
Puerto Rico rifts away from Hispaniola, a process that continues throughout the Miocene and Pliocene
(figure 2h; Pindell and Kennan, 2009). A restraining bend developed in Jamaica in the middle to late
Miocene (Abbott et al., 2013; James-Williamson et al., 2014), linking the Walton Fault in the west to
the Enriquillo-Plantain Garden Fault Zone in the east (Mann et al., 2007b). During the late (?) Miocene
the Muertos Trough retro-arc system became active (Mann et al., 2002; ten Brink et al., 2009).

37
Chapter 2: Geological review of the Caribbean region

3. Fault zones and seismicity

3.1 The Septentrional – Oriente fault system

The Septentrional – Oriente fault system is the northern left-lateral strike-slip system that borders the
northern margin of Hispaniola. The western part of the ENE-trending Oriente Fault Zone (OFZ) is
bounding the Cayman Trough in the north (Rosencrantz et al., 1988). South of Cuba the OFZ runs
through the Santiago Deformed Belt (SDB), where the Gonâve microplate underthrusts Cuba (Calais et
al., 1998; Moreno et al., 2002). The eastern part of this fault system is the Septentrional Fault Zone
(SFZ), consisting of the continuation of the strike-slip fault to the east of Cuba. East of the SDB the trace
of the SFZ is bending clockwise to an ESE trend north of Haiti (Leroy et al., 2015). The SFZ continues
onshore in the Dominican Republic (Edgar, 1991; de Zoeten and Mann, 1991), extends offshore
eastwards (Mann et al., 1998; van Gestel et al., 1998) and terminates at the Mona Passage (Mondziel
et al., 2010).

The OFZ became active in the early Oligocene when motion along the Cauto Fault was terminated
(Leroy et al., 2000; Rojas-Agramonte et al., 2006). The SFZ between Cuba and Haiti developed from the
Burdigalian (20 Ma) onwards, as evidenced by offshore seismic data and onshore correlatable
stratigraphy and unconformities between Cuba and northern Hispaniola (Pindell and Barrett, 1990;
Calais and Mercier de Lépinay, 1995).

The SFZ in Hispaniola consists of a western and an eastern segment. The onshore eastern segment dips
roughly 50° north with a vertical offset of 5 – 7 km (Erikson et al., 1998). This is in contrast to the
offshore part of the Septentrional – Oriente fault system where the main faults are sub-vertical and
positive flower structures are well developed (figure 3a; Calais and Mercier de Lépinay, 1995). The
onshore eastern segment of the SFZ is activated in the early to middle Miocene (~20 – 15 Ma) (de
Zoeten and Mann, 1991; Erikson et al., 1998; de Zoeten and Mann, 1999). From the late Miocene until
the early Pliocene (~11 – 4 Ma) the onshore SFZ is characterized by major strike-slip motion (Erikson
et al., 1998). Renewed compression reactivates the onshore eastern segment of the SFZ during the
early Pliocene (~4 – 2.5 Ma) (Edgar, 1991; Erikson et al., 1998; de Zoeten and Mann, 1999). Throughout
its evolution, the left-lateral strike-slip related offset along the SFZ is estimated at 85 km (Erikson et
al., 1998). The western offshore segment of the SFZ became active during the Pliocene (Erikson et al.,
1998). The onset of motion along this active segment is estimated at 1.8 Ma (Leroy et al., 2015).

38
Chapter 2: Geological review of the Caribbean region

Onshore Haiti the EPGFZ is a N80°E trending lineament (Mann and Burke, 1984) that continues east
(Mann et al., 1995) and possibly connects with the MT (Mauffret and Leroy, 1997). Pull-apart basins
are found along its onshore trace (Mann et al., 1983; Calmus, 1983). Ongoing relative subsidence of
these basins is evidenced by active alluvial fans along the basin borders (Mann et al., 1983; Bourgueil
et al., 1988). Present-day strike-slip activity is further supported by the left-lateral offset of fluvial
channels along the EPGFZ trace (Prentice et al., 2010). Lateral offset along the EPGFZ is estimated
between 25 and 50 km based on lithological markers (Calmus, 1983; Van den Berghe, 1983a), although
these values should be considered with some caution (Mann et al., 1995).

3.3 The Haitian Fold-and-Thrust Belt

The Haitian fold-and-thrust belt (HFTB) developed in western Hispaniola and consists of a series of
stacked NW – SE trending southwest-verging thrust sheets (Pubellier et al., 2000). The offshore
continuation of the HFTB is visible on seismic profiles (Mann et al., 1995). The HFTB is bounded by the
SFZ in the north and the EPGFZ in the south. Progressive activation of the thrust faults is recorded by
shallowing of depositional facies in piggyback basins. The HFTB consists of two thrust sheets onshore.
The thrust bounding the northeastern thrust sheet (Montagnes Noires – Sierra de Neiba) was activated
during the early Miocene (Pubellier et al., 2000). The thrust bounding the southwestern thrust sheet
(Chaîne des Matheux – southern Sierra de Neiba) developed in the middle-late Miocene (Pubellier et
al., 2000). The southern flank of the southwestern thrust sheet was affected by strike-slip tectonics
from the Pliocene onwards (Vila et al., 1986a; Pubellier et al., 1991).

3.4 Seismicity

Seismicity changes along the main plate-bounding faults from the MCSC in the west to the Virgin
Islands in the east (figure 4a). Focal mechanisms along the OFZ close to the MCSC indicate strike-slip
motion, which corresponds well with GPS velocity vectors in this region (Benford et al., 2012a; Symithe
et al., 2015). The OFZ experiences oblique compressional deformation south of Cuba at the SDB
(Moreno et al., 2002). Seismicity along the fault segment east of the SDB and offshore Haiti appears
mainly strike-slip, although the constraint from focal mechanisms in this area is poor (Ali et al., 2008).
The SFZ accommodates most of the left-lateral strike-slip component of the North America – Caribbean
motion in this region (Dolan and Wald, 1998). Convergence between the Bahamas platform and
Hispaniola accommodated by the NHDB causes dip-slip earthquakes (Ali et al., 2008). Earthquakes in
the western NHDB are limited to crustal levels (< 50 km) since there is no longer a slab attached to this

40
Chapter 2: Geological review of the Caribbean region

3.5 The 2010 Leogâne earthquake

The seismogenic nature of southwestern Hispaniola was demonstrated by the Mw 7.0 Leogâne
earthquake that struck Haiti close to its capital Port-au-Prince on January 12th, 2010 (Hayes et al., 2010;
Bilham, 2010b). This event caused widespread destruction and an estimated mortality of 160.000 in
the Port-au-Prince area alone (Kolbe et al., 2010). The transpressive earthquake most likely did not
rupture the main, south-dipping EPGFZ fault segment in this area (Prentice et al., 2010), but probably
occurred on the NNW-dipping blind Leogâne Fault (Mercier de Lépinay et al., 2011; Douilly et al., 2013).
Oblique displacement on this fault was partitioned into one third reverse and two-thirds strike-slip
(Calais et al., 2010). Aftershocks with a reverse motion (Symithe et al., 2013) occurred along the SSW-
dipping Trois Baies Fault of Bien-Aime Momplaisir (1986).

Older GPS velocity models predicted parallel to slightly oblique displacement along the EPGFZ (Dixon
et al., 1998; DeMets et al., 2000; Calais et al., 2010). Newer GPS velocity models (Benford et al., 2012a;
Symithe et al., 2015; Calais et al., 2016) that separate the Gonâve microplate from the Hispaniola block
arrive at velocity vectors that are more consistent with the deformation recorded by the 2010 main-
and aftershock patterns.

3.6 Historical earthquakes

Historical accounts of earthquakes in Hispaniola go back as far as 500 years (ten Brink et al., 2011;
Bakun et al., 2012). Associating these earthquakes with known fault zones such as the EPGFZ and SFZ
is difficult and best done by combining the historical record with paleoseismic data (Prentice et al.,
2013; ten Brink et al., 2013).

In southern Hispaniola around 6 earthquakes with Mw > 6.5 have been documented post-1500 that
occurred within the vicinity of the EPGFZ. These are the 1701, October 1751, November 1751, 1770,
1830 and 2010 earthquakes (McCann, 2006). The October 1751 event probably occurred offshore
southern Hispaniola on the Muertos system (McCann, 2006), although it could be related to the
onshore connection between the EPGFZ and the MT (Bakun et al., 2012). The other 5 events likely
occurred proximal to the EPGFZ, although there is some disagreement on their location and magnitude
(Dolan and Wald, 1998; Ali et al., 2008; Bakun et al., 2012). The locations and rupture lengths presented
in figure 4a (from Ali et al. (2008)) should therefore be treated with caution. The 1751 or 1770 event
is possibly responsible for left-lateral surface offset found along the trace of the EPGFZ south of Port-
au-Prince (Prentice et al., 2010). Other historical earthquakes in southern Hispaniola occurred in the
vicinity of the Muertos Trough (McCann, 2006; ten Brink et al., 2011).

42
Chapter 2: Geological review of the Caribbean region

In northern Hispaniola the two historical earthquakes of 1842 and 1887 with Mw > 7.5 are associated
with the offshore part of the SFZ (Scherer, 1912; Mann et al., 1998). Paleoseismological studies found
that the youngest surface rupture along the SFZ east of Santiago in the Dominican Republic dates back
to 1230 A.D. (Prentice et al., 1993; Mann et al., 1998; Prentice et al., 2003). This makes it unlikely that
the Mw > 6.5 earthquakes that occurred in northern Hispaniola in 1562 and 1783 ruptured the SFZ
east of Santiago (Prentice et al., 2013). Other Mw > 6.5 earthquakes in northern Hispaniola, such as
the 1897 event (McCann, 2006), are associated with subduction of the North American plate under
Hispaniola at the North Hispaniola Deformed Belt (ten Brink et al., 2011). The more recent 1946, 1953
and 2003 earthquakes in northern Hispaniola are also associated with thrusting under northern
Hispaniola (Dolan and Wald, 1998).

Earthquakes in Hispaniola with magnitudes larger than 6.5 are temporally and geographically clustered
(Dolan and Wald, 1998). Between the early 1600s to late 1700s and from the 1960s to recent,
earthquake activity was focused in southern Hispaniola. Prior to the 1600s and from the mid-1800s to
mid-1900s, earthquake activity was concentrated in northern Hispaniola (figure 4b, and references
therein). The anti-correlation between earthquake activity in northern versus southern Hispaniola is
possibly explained by a small reduction of Coulomb stress build-up on one fault system during
earthquake activity on the other system. This reduces the likelihood that earthquakes occur
simultaneously on both the northern and the southern fault systems (Ali et al., 2008).

4. Discussion

The following discussion focusses on two questions that are: 1) what is the timing of initiation of the
EPGFZ, and 2) how was Eocene – Recent convergence between the Gonâve microplate and Hispaniola
block accommodated?

4.1 Timing of initiation of the EPGFZ

The first constraints about the timing of activity along the EPGFZ comes from the inferred regional
geodynamic evolution. Between 26 and 20 Ma the Mid-Cayman Spreading Center grew southwards,
creating the Hendrix pull-apart basin and activating the Walton Fault (Leroy et al., 2000). The
Wagwater Graben in Jamaica (Draper, 1987) is inverted as part of the Jamaican restraining bend (Mann
et al., 2007b). This inversion started in the (mid?)-late Miocene (Abbott et al., 2013; James-Williamson
et al., 2014), indicating that the EPGFZ in eastern Jamaica was active by at least the middle – late
Miocene (Mann et al., 2007b). Deformation related to the EPGFZ is well visible on seismic sections in
the Jamaica Passage (Corbeau et al., 2016a).

43
Chapter 2: Geological review of the Caribbean region

The second constraint comes from southern Haiti, although the activation history of the EPGFZ in this
region is more difficult to determine. West of Camp Perrin lies a rectangular basin that, based on
satellite images, appears situated between two E – W trending branches of the EPGFZ. If this basin
developed as a pull-apart basin in response to activity on the EPGFZ as proposed by Bizon et al. (1985),
its infill should record the onset of activation. The sediments in this continental basin consist of
conglomerates overlain by lacustrine sands and clays with interbedded lignite layers of early
(Burdigalian) to at least middle (Langhian) Miocene age (Calmus, 1983; Bizon et al., 1985). About 10
km east of this basin, along-strike but south of the EPGFZ, middle Miocene marine marls are found,
which point to a significantly deeper depositional environment (Calmus, 1983; Bizon et al., 1985).

Bourgueil et al. (1988) argue that an early Miocene age of pull-apart type initiation for the basin west
of Camp Perrin is unlikely since the early to middle Miocene in southern Haiti was characterized by
relatively deep marine facies. A subsequently subsiding pull-apart basin would then be the site of even
deeper marine facies, not continental material. These authors propose that the fauna dated by Calmus
(1983) is reworked and redeposited, and are in favor of a Pliocene to Pleistocene age for the onset of
strike-slip activity in southwestern Haiti. Alternatively, the continental deposits west of Camp Perrin
could be part of an older basin that is unrelated to strike-slip activity on the EPGFZ. This older basin
then subsequently became incorporated and overprinted by the presently active Clonard pull-apart
basin of Mann et al. (1983, 1995).

A third constraint comes from the GPS velocity models and offset based on geological markers along
the EPGFZ.

The total amount of lateral offset along the EPGFZ is estimated between 25 and 50 km (Calmus, 1983;
Van den Berghe, 1983a). Fault parallel GPS velocities for the EPGFZ in southern Haiti are roughly 7 mm
yr-1 left-lateral (Benford et al., 2012a). Assuming that these GPS velocities are representative for short
geologic time scales, a rate of 7 mm yr-1 amounts to 7 km offset per Myr. Taking the offset based on
geological markers this yields 3.6 Ma (25 km) to 7.2 Ma (50 km) as a minimum time for the onset of
motion on the EPGFZ in southern Haiti. Although these numbers should be treated with caution (Mann
et al., 1995), they correspond well to the onset of activity on the EPGFZ in Jamaica (Abbott et al., 2013;
James-Williamson et al., 2014) and to the late Miocene age proposed for the onset of activity along
the EPGFZ by Leroy et al. (2000) .

Based on the above it is most likely that activity on the EPGFZ started in the middle to late Miocene in
southeastern Jamaica. The fault then possibly migrated eastwards with activity in southern Haiti
commencing during late Miocene times. This timing focusses on the onset of activity along the present-
day active trace of the EPGFZ. It cannot be ruled out that deformation commenced earlier, in a more

44
Chapter 2: Geological review of the Caribbean region

diffuse manner and over a broader zone covering on- and/or offshore southern Haiti. Several E – W
trending, discontinuous traces are observed in southwestern Haiti (Calmus, 1983; Amilcar, 1997). It is
possible that these fault strands accommodated part of the initial deformation, which afterwards
became focused on one single fault strand, the EPGFZ. A better constraint on the timing of initiation
of the EPGFZ in southern Haiti could be provided by improved biostratigraphic dating of deposits in
the pull-apart basins and by a detailed assessment of the amount of left-lateral offset along the EPGFZ.

4.2 Convergence between the Gonâve microplate and Hispaniola block

Oceanic spreading in the Cayman Trough since 49 Ma has separated its margins by around 950 km
(Leroy et al., 1996). Restoring the margins of the Cayman Trough to their position prior to oceanic
spreading in the early Eocene translates southern Haiti and the Beata Ridge to a position SE of the
Yucatan margin (figure 2d; Leroy et al., 2000). During this time, the Hispaniola block with its island-arc
units was located south of Cuba (Pindell and Barrett, 1990; Calais and Mercier de Lépinay, 1995), while
central Haiti occupied an intermediate position between the two (Mann, 2007; Pindell and Kennan,
2009). Using the palinspastic reconstruction at 46 Ma (figure 2e) from Pindell and Kennan (2009)
results in a separation of around 600 km between the eastern margin of the Cayman Trough and
western Hispaniola. As the eastern margin of the Cayman Trough is adjacent to western Hispaniola in
the present-day, this indicates that the amount of early Eocene separation between the two has been
accommodated since that time.

Prior to addressing possible scenarios of how this convergence is accommodated, it is important to


verify if the 600 km separation stated above is a correct number. This would require a detailed
evaluation of the palinspastic position of central Hispaniola and the eastern margin of the Cayman
Trough at 49 Ma. For the post-49 Ma period this evaluation should try to (i) identify and quantify
shortening within the Cayman Trough, if present, (ii) verify if the Yucatan peninsula and Bahamas
platform are part of the same rigid North American plate, did not experience relative motion with
respect to each other and can be used as palinspastic markers, (iii) identify the palinspastic position of
the eastern Cayman margin with respect to the Yucatan peninsula, (iv) identify the palinspastic position
of the island-arc terranes in Hispaniola with respect to southern Cuba, (v) restore the amount and
direction of shortening between Cuba and the Bahamas platform, and (vi) identify the timing and
amount of offset along the strike-slip faults between the Cayman Trough and the Yucatan Peninsula.

There are several scenarios that propose how this post-49 Ma convergence could have been
accommodated. A first option could be to subduct Hispaniola and the Caribbean plate northwest-
wards under the Cayman Trough and/or Nicaragua Rise (Sigurdsson et al., 1997b). No slab that could

45
Chapter 2: Geological review of the Caribbean region

be associated with such a subduction system has been found (van Benthem et al., 2013), nor does such
a setting appear compatible with the regional evolution and oceanic spreading in the Cayman Trough
(Leroy et al., 2000). A second option is to subduct the eastern margin of the Cayman Trough/Caribbean
plate under central Hispaniola (Leroy et al., 2000). Island-arc volcanism terminated in southern Cuba
and central Hispaniola at respectively 45 and 40 Ma (compilation of data in Boschman et al., 2014).
This timing has been used to propose a short-lived north-dipping subduction system located southwest
of Cuba and Hispaniola (Pindell et al., 2006). Such a system would result in a north-dipping slab under
Hispaniola or southern Cuba. Published seismic tomographic lines (e.g. Ten Brink et al., 2013; van
Benthem et al., 2013) unfortunately do not exactly cover this region. The Paleocene arc-related
volcanism could also be related to the final phase of southwest-directed subduction of proto-
Caribbean lithosphere (Rojas-Agramonte et al., 2006). A third way to account for the convergence is
using strike-slip tectonics as the dominant process. Strike-slip activity in central Hispaniola during (late)
Eocene to early Miocene (Witschard and Dolan, 1990; Draper, 1999; Hernaiz Huerta and Pérez-Estaún,
2002) could have accommodated between 350 and 700 km of left-lateral displacement (Pindell and
Barrett, 1990). Additional strike-slip activity in northeast central Hispaniola from the early Oligocene
onwards (Dolan et al., 1991; Coleman and Winslow, 1999) could have accommodated 120 km of left-
lateral displacement (Draper, 1999). These numbers are of the same order of magnitude as the total
amount of convergence needed. The structures described above presently trend roughly NW – SE. If
these structures had the same orientation during fault activity, left-lateral displacement alone is not
sufficient to accommodate the roughly E – W directed convergence. NW – SE trending left-lateral faults
accommodating E – W shortening would displace and translate blocks towards the southeast. A
complex amalgamation of left-lateral and right-lateral faults would be needed to allow blocks to escape
eastwards and to keep the general convergence between central Hispaniola and the Cayman Trough
margin operating in an E – W direction. If the strike of the faults in central Hispaniola was E - W during
their activity, left-lateral displacement alone could be sufficient to accommodate E – W convergence.
This would in turn require rotation of the blocks in the interior of Hispaniola since that time. A detailed
investigation of paleomagnetic data within the area could reveal if this is a possibility or not.

Part of the convergence between the eastern Cayman margin and Hispaniola has been accommodated
by folding and thrusting in the HFTB since the early Miocene (Pubellier et al., 2000), although no
estimate of the total amount of shortening is available. During the early Miocene Hispaniola became
detached from southern Cuba and moved 400 km eastwards along the Oriente – Septentrional fault
system to its present-day position (Calais and Mercier de Lépinay, 1995). Because the eastern Cayman
margin and Hispaniola are both bounded to the north by the Oriente – Septentrional fault system the

46
Chapter 2: Geological review of the Caribbean region

eastward motion of Hispaniola does not change the total amount of shortening required between its
western margin and the eastern margin of the Cayman Trough.

The exact amount of shortening between the eastern margin of the Cayman Trough and western
Hispaniola and the mechanism to accommodate this remains a question. A detailed kinematic plate
tectonic reconstruction of the Yucatan – Cayman – Cuba – Hispaniola region could help reduce the
uncertainties and limit the possible mechanisms.

5. Summary

1. Relative ENE-ward motion between the Caribbean and North American plates is

accommodated in the northern Caribbean by a complex amalgamation of tectonic blocks and

a microplate bounded by a spreading center, two strike-slip systems and two

subduction/underthrusting systems.

2. Strike-slip activity in the northern Caribbean has been ongoing since the Late Cretaceous as a

results of oblique convergence between the Caribbean and North American plates. At present,

strike-slip activity in this region is partitioned between a northern and a southern strike-slip

system.

3. The northern strike-slip system consists of the Oriente Fault Zone (OFZ) in the west and the

Septentrional Fault Zone (SFZ) in the east. The Oriente segment north of the Cayman Trough

became active in the early Oligocene. In the early Miocene the segment south of Cuba became

active which allowed eastward migration of the Hispaniola block. Convergence between

Hispaniola and the Bahamas Platform is accommodated by underthrusting at the North

Hispaniola Deformed Belt, while continued relative eastward motion is accommodated by the

left-lateral SFZ. The SFZ onshore Hispaniola became active between 20 – 15 Ma and

subsequently enjoyed several pulses of transpressional deformation. The active trace of the

SFZ offshore Haiti only became active around 1.8 Ma.

4. The southern strike-slip system consists of the Walton Fault Zone (WFZ) west of Jamaica and

the Enriquillo-Plantain Garden Fault Zone (EPGFZ) east of Jamaica and onshore Hispaniola. The

47
Chapter 2: Geological review of the Caribbean region

WFZ became active between 26 – 20 Ma as a result of southward propagation of the Mid-

Cayman Spreading Center. A restraining bend developed onshore Jamaica as a result of

interaction between the WFZ and the EPGFZ in the middle to late Miocene. Timing of initiation

of the EPGFZ onshore Hispaniola is still subject to debate. This could have started as early as

the early Miocene or as late as the latest Miocene.

5. The strike-slip fault systems in the northern Caribbean are seismically active, as illustrated by

historical and recent earthquakes. Earthquake activity is temporally and geographically

clustered in northern and southern Hispaniola, showing an anti-correlation in earthquake

activity between the two regions. The most recent Mw > 7.0 seismic event, the 2010 Léogâne

earthquake, resulted in widespread destruction in southern Haiti. This transpressive

earthquake ruptured a previously unmapped buried fault segment and likely did not involve

motion or stess release on the EPGFZ.

6. The eastern margin of the Cayman Trough and central Hispaniola have converged between the

initiation of spreading at 49 Ma and the present-day. The exact amount of convergence is

unknown, but is likely in the order of 600 km. How this convergence was accommodated, being

by subduction, strike-slip, or a combination of both, still remains unclear.

48
Chapter 2: Geological review of the Caribbean region

Acknowledgements

I want to thank Anouk Beniest for her continuous effort, support and suggestions that have greatly
helped shape and improve this chapter. I am grateful to Nicolas Bellahsen and Claudio Rosenberg for
their revisions of the manuscript. I want to extend my gratitude to James Pindell for allowing me to
use the redrafted palinspastic reconstructions. Lastly I would like to thank João Duarte for granting me
the opportunity to publish this chapter.

49
Chapter 2: Geological review of the Caribbean region

Figure captions

Figure 1. a) Geodynamic setting of the Caribbean. GPS velocities (red arrows) w.r.t a fixed Caribbean plate are from (DeMets
et al., 2010). b) Geodynamic setting of the northern Caribbean. Dashed lines delineate the Gonâve microplate and Hispaniola,
Puerto Rico – Virgin Islands and North Hispaniola blocks. Blue dashed lines and corresponding GPS velocity vectors are from
Calais et al. (2016), while black dashed lines and corresponding GPS velocity vectors are from Symithe et al. (2015). Velocity
vectors indicate motion of south(west) boundary w.r.t. north(east) boundary. Faults modified after Leroy (1995). Caribbean
– North American plate motion is from DeMets et al. (2010). Abbreviations: LA = Lesser Antilles; SCDB – South Caribbean
Deformed Belt SIFZ = Swan Island Fault Zone; MCSC = Mid-Cayman Spreading Center; OFZ = Oriente Fault Zone; SDB = Santiago
Deformed Belt; SFZ = Septentrional Fault Zone; NHDB = North Hispaniola Deformed Belt; MP = Mona Passage; PRT = Puerto
Rico Trench; MT = Muertos Trough; HFTB = Haitian Fold-and-thrust belt; EPGFZ = Enriquillo-Plantain Garden Fault Zone; HB =
Hendrix pull-apart Basin; WFZ = Walton Fault Zone; D.R. = Dominican Republic; P.R. = Puerto Rico; V.I. = Virgin Islands; PRVI =
Puerto Rico – Virgin Islands block.

Figure 2. a) Collision of the Great Arc of the Caribbean with the Yucatan Peninsula, modified from Stanek et al. (2009). b)
Timing of activation of faults between Yucatan and the Cayman Trough, modified from Leroy et al. (2000). c – h) Palinspastic
reconstruction of the northern Caribbean from 71 to 10 Ma, modified from Pindell and Kennan (2009). Modifications from
Pindell and Kennan (2009) in figure 2e; 1) Cauto Fault assumed active and left-lateral at this time; and 2) no underthrusting
of the LNR under the UNR assumed to occur post-49 Ma. Abbreviations: CA = Caribbean; NA = North American; UNR = Upper
Nicaragua Rise; LNR = Lower Nicaragua Rise; JAM = Jamaica; YUC = Yucatan Peninsula; SHBR = Southern Haiti - Beata Ridge;
HFTB = Haitian Fold-and-thrust Belt; HIS = Hispaniola; PR = Puerto Rico.

Figure 3. a) Seismic line CT2-25C crossing the SFZ in the Windward Passage Deep, from Calais and Mercier de Lépinay (1995).
B = Oligocene – Aquitanian; A’ = Burdigalian – Langhian and onlapping sequence B; A = Serravallian – Pliocene and onlapping
sequence B and A’. b) Seismic line H12-036 crossing the EPGFZ in the Jamaica Passage between Jamaica and Haiti, from
Corbeau et al. (2016a). Location of seismic lines in figure 1b. SB = seismic basement; U1 = syn-rift sequence, Paleocene to
Ypresian; U2 = post-rift sequence, Ypresian to early Miocene; U3 = post-rift sequence, early to middle Miocene; U4 = syn-
tectonic (EPGFZ activity) sequence, late Miocene to recent.

Figure 4. a) Selection of historical and recent earthquakes. Focal mechanisms in grey are historical (Mw > 7.0, pre-1900)
earthquakes, focal mechanisms in black are recorded earthquakes (Mw > 5.0, post-1900). References; 1 = Van Dusen and
Doser (2000), 2 = Wiggins-Grandison and Atakan (2005), 3 = Ali et al. (2008), 4 = International Seismological Centre (2014)
(Mw > 5.0, 1976 – 2014). b) Earthquakes of Mw > 6.5 in northern Hispaniola between 1500 and 2016. Earthquakes in green
from northern Hispaniola and possibly linked to the (offshore) SFZ and subduction at the NHDB. Earthquakes in red from
southern Hispaniola and possibly linked to the EPGFZ or underthrusting at the Muertos Trough. References; 1 = McCann
(2006), 2 = ten Brink et al. (2011), with magnitude on intensity scale (Mi), 3 = International Seismological Centre (2014) (NEIS),
4 = International Seismological Centre (2014) (GCMT). Note magnitude difference for event 1751b between McCann (2006)
and ten Brink et al. (2011). Abbreviations: SIFZ = Swan Island Fault Zone; MCSC = Mid-Cayman Spreading Center; OFZ = Oriente
Fault Zone; SDB = Santiago Deformed Belt; SFZ = Septentrional Fault Zone; NHDB = North Hispaniola Deformed Belt; MP =
Mona Passage; PRT = Puerto Rico Trench; MT = Muertos Trough; HFTB = Haitian Fold-and-thrust belt; EPGFZ = Enriquillo-
Plantain Garden Fault Zone; HB = Hendrix pull-apart Basin; WFZ = Walton Fault Zone.

50
Chapter 3: Geological review of Hispaniola

3 Geological review of Hispaniola

3.1 Morphotectonic zones


The island of Hispaniola is located along the transpressive plate boundary between the Caribbean and
North American plates. It consists of Haiti in the west and the Dominican Republic in the east. The
island can be divided into two major terranes based on the type of basement; (1) the southwestern
one-third of the island consists mainly of oceanic plateau basement with CLIP affinities, while (2) the
northeastern two-thirds consists of island-arc and oceanic terranes related to the Great Arc of the
Caribbean (Mann et al., 1991b). The surface geology of the island displays a series of mountain ranges
and Neogene basins, interpreted as tectonic terranes (Case et al., 1984; Mann et al., 1991b) or
morphotectonic zones (Lewis and Draper, 1990).

The morphotectonic zones of Hispaniola are listed below, from north to south, and the corresponding
structures are illustrated in Figure 3 - 1.

(1) The Cordillera Septentrional, bounded to the south by the Septentrional Fault Zone (SFZ), and
dissected by the Camú (CFZ) and Rio Grande (RGFZ) fault zones. Igneous and metamorphic
basement outcrops in the Puerto Plata, Rio San Juan and Samaná complexes (Escuder-Viruete
et al., 2011b; c), while basement in the Pedro Garcia complex and western region is non-
metamorphic (de Zoeten and Mann, 1991; Escuder-Viruete et al., 2013).
(2) The Cibao Basin, a NNW – ESE trending Oligocene – Pliocene basin bounded to the north by
the SFZ and to the south by the Cordillera Central (Edgar, 1991).
(3) Northern Plain, bounded to the north by the SFZ and Hispaniola Fault Zone (HFZ), and bounded
to the south by the Cordillera Central – Massif du Nord and Bonao – Guácara Fault Zone (BGFZ).
(4) The Massif du Nord, representing the westernmost extension of the Cordillera Central,
exposing Cretaceous island-arc and latest Cretaceous – Cenozoic sedimentary rocks (Boisson
and Vila, 1982).
(5) The Cordillera Central, which comprises, from NE - SW; (1) the eastern Cordillera Central,
bounded to the NE by the Hatillo Thrust (HT) and to the south by the HFZ, (2) the Bonao block,
limited by the HFZ and Hato Viejo Fault Zone (HVFZ), (3) the Jarabacoa block, bounded by the
HVFZ and the La Meseta Shear Zone (LMSZ), and (4) the Jicomé block, bounded by the BGFZ
and the San José – Restauración Fault Zone (SJRFZ). Exposed are mainly Cretaceous igneous
island-arc and mantle plume-influenced units, low-grade metamorphic rocks, and associated
sedimentary sequences (Escuder-Viruete et al., 2002, 2007b, 2010, 2011a; Lewis et al., 2002).

51
Chapter 3: Geological review of Hispaniola

(6) Cordillera Oriental – Seibo, located east of the Cordillera Central, mainly exposing island-arc
related material and Cenozoic sediments (Lebrón and Perfit, 1994; Escuder-Viruete et al.,
2006a).
(7) The Peralta Belt, a 320 km long NW – SE trending fold and thrust belt exposing Cretaceous to
Recent sediments (Witschard and Dolan, 1990). This belt is separated from the Cordillera
Central by the NE-dipping SJRFZ, and from the basins to the south by the NE-dipping Los Pozos
– San Juan Fault Zone (LPSJFZ) (Mann et al., 1991c).
(8) The Plateau Central – San Juan Basin – Azua, Artibonite Valley, and Cul-de-Sac – Enriquillo Plain
are all fault bounded Neogene thrust-related basins (Pubellier et al., 2000).
(9) The NW Peninsula – Terre Neuve – Montagnes Noires – northern Sierra de Neiba mountain
chain exposes mainly Cenozoic sedimentary and basaltic sequences, with minor volcanic-arc
derived units (Butterlin, 1960). This chain comprises a dismembered thrust sheet with a
sigmoidal trending fold axis, bounded to the SW by the Montagnes Noires Fault Zone (MNFZ).
(10)The Chaîne des Matheux – southern Sierra de Neiba chain exposes mainly Cenozoic sediments,
and possibly Cretaceous CLIP-related basement (Vila et al., 1986a). It is bounded to the SW by
the Matheux Fault Zone (MFZ).
(11)The Southern Peninsula comprises a 350 km long terrane of Cretaceous CLIP plateau basalts
and intercalated sediments, overlain by Cenozoic sediments (Woodring et al., 1924), exposed
in the Massif de la Hotte, Massif de la Selle and Sierra de Bahoruco (Butterlin et al., 1976). This
morphotectonic zone is transected by the E – W trending, left-lateral Enriquillo-Plantain
Garden Fault Zone (EPGFZ) (Calmus, 1983; Van den Berghe, 1983a; Mann et al., 1984).

3.2 Igneous and metamorphic domains in Hispaniola


A synthetic stratigraphic chart of the igneous and metamorphic units is shown in Figure 3 - 2, while
their geographic distribution is presented in Figure 3 - 3. The geodynamic setting of these igneous units
is found in Figure 3 - 4 and the setting, timing and P-T paths of the HP-LT metamorphics is shown in
Figure 3 - 5. A short summary is provided below, whereas a detailed geochemical and geodynamic
description of the igneous and metamorphic domains is given in the subsections of section 3.2. Ages
constrained by radiometric dating are indicated by numbers, with the dating technique and dated
minerals listed when available. Ages constrained by biostratigraphy are indicated by the name of the
series (i.e. Maastrichtian).

53
Chapter 3: Geological review of Hispaniola

Cenozoic non-arc related HP/LT


CLIP Quaternary Great Arc of the Caribbean Subduction
Southern Peninsula Cordillera Central Cordillera Septentrional

lt
u

be
Ea
d’

ta
a
u x/

zu

al
ba /
u
co

on es d heu

s
er ne Tro

er

al
/A
tte

ra
Ne ire
ru
le

Ba l /

nt
-P

ck

lle
t

uv /
l

rd
ho
Ho

de No

sin
an tra
Se

gn a

ie
Ne la

lo
ta s M

l di
s

No

k
i

an
ia
Or
Ba

k
e
u
re
ui /

a
Ju n

ab
oc
la
la

ra s

ra or
oc

rc
at
rre ins
riq ac

n Ce

ie

Ju
on e
llo

ra
du
de

bl
de

nt C

Ga
Pl
de

co

bl
M ne d
En e-S

iv

á
Te Pen
Si tag

ille
Sa u

n
Ce ern
é

an
sR

to
ba
sif

sif

o
ea
sif

Sa
o
rra

om
l-d

na

rd

dr
er

m

as

as
as

st
oi
at

ra
NW

o
e

Cu

Ch

Co
Bo

Pu

Sa
Jic
Ma

Pe
Ea
Tr
M

Ri
Pl

Ja
M

Si
0 Pleist. Q
Pliocene
Neogene

10 Mio. 0 kbar 25
0 °C 500
20
RM
Oligo. Perodin / El Aguacate
30
Formation Loma de SBS
Cabrera
Batho.
Paleogene

40
Eoc. Baradères Formation El Bao
Batho.
50 0 kbar 25

?
U. Tireo Fm. PB
60 Paleoc.
PPDB Siete Peralvillo j-bl
Cabeza Sur Fm. o-bl
70 Maast.
Late Cretaceous

Fm.
Camp.
80 La Mine Fm. Guayabas PBM

? ? ? ? ? ?
Sant.
Con. Fm.
90 Tur.

100
Cen. Rio
Verde
?
Terrier Rouge Fm. JCM
Albian Fm.
110 Perches Fm. ? ?
?
Early Cretaceous

Aptian ? ? ? L. Tireo
120 IAT / CA island-arc [extrusive] Fm.
Dumisseau Amina - Los
Barr.
130 Haut Formation
IAT island-arc / Tonalite [plutonic] Morne
Cabril ? Duarte ? Maimón Ranchos ? ? ?
Cenozoic CA Complex Loma Schists Fm.
Valan.
140 MA / CA Quaternary Caribe
Berr. Peridotite
IAT / CA SSZ ophiolites El Aguacate
150 Torth. Chert
Kimm. (E-MORB or OIB)
Jurassic

160 Oxford. Proto-Caribbean / Farallon (N-MORB) Loma La


Call. Limestone Chert Turbidite Monja Assemblage
Bath. Schist Volcaniclastics
170 Baj.
Fault Unconformity

Figure 3 - 2: Stratigraphy of the main igneous units in Hispaniola. The morphotectonic zones are found in Figure 3 - 1. HP-LT metamorphic paths after Escuder-Viruete et al.
(2011b; c). Solid circles are constrained P-T-t conditions, open circles are estimated P-T-t conditions by these authors. Abbreviations for the metamorphic units; PBM = Punta
Balandra Mélange, PB = Punta Balandra sediments, SBS = Santa Bárbara Schists, RM = Rincón Marbles.

54
Chapter 3: Geological review of Hispaniola

The exposed igneous basement on the island of Hispaniola consists of: (1) intrusive, extrusive, and
ophiolitic island arc-related mafic to felsic units associated with the Cretaceous to Eocene Great Arc of
the Caribbean (Bellon et al., 1985; Kesler et al., 1991a; Escuder-Viruete et al., 2007b; a, 2008, 2014),
and (2) Cretaceous tholeiitic basalts related to the Caribbean Large Igneous Province (Maurrasse et al.,
1979a; Sen et al., 1988; Sinton et al., 1998) and other Cretaceous basalts characterized by a plume-
influenced geochemistry (Escuder-Viruete et al., 2007c, 2011a). Non-arc related Cenozoic volcanism is
mainly restricted to the SW side of the Cordillera Central (Kesler, 1971; Wadge and Wooden, 1982;
Calmus, 1983; Bourgueil et al., 1988; Hernaiz Huerta et al., 2004; Kamenov et al., 2011a; b). HP – LT
metamorphism is limited to the Cordillera Septentrional in the Dominican Republic (Abbott et al., 2007;
Krebs et al., 2008; Escuder-Viruete et al., 2011d, c; b; Krebs et al., 2011; Escuder-Viruete et al., 2014),
while low-grade metamorphic facies are found throughout the Cordillera Central and in NW-Haiti
(Kesler et al., 1977a; Bourgueil et al., 1988; Kesler et al., 1991a; Draper et al., 1996; Escuder-Viruete et
al., 2002, 2007a; c).

3.2.1 CLIP-related tholeiitic flood basalts


Mantle-plume related flood basalts of the Caribbean Large Igneous Province are found in southwestern
Haiti (Figure 3 - 3). On the Southern Peninsula these Cretaceous tholeiitic flood basalts (Woodring et
al., 1924) are overlain by the Senonian Macaya Formation (Butterlin, 1960) and are subdivided into the
lower and upper Dumisseau Formation based on age and geochemistry (Maurrasse et al., 1979b). The
lower Dumisseau Formation yields 95 – 102 Ma (K-Ar; Bien-Aime Momplaisir, 1986) and 92 - 89 Ma
(Ar-Ar in whole-rock; Sinton et al., 1998) radiometric ages, while biostratigraphy constrains their age
from Aptian to Santonian (Reeside, 1944; Maurrasse et al., 1979a; Van den Berghe, 1983a; Bien-Aime
Momplaisir, 1986; Desreumaux, 1987). These basalts share similarities with basalts cored in the
Venezuelan Basin and on the Lower Nicaraguan Rise (Sen et al., 1988), whose N-MORB geochemistry
indicates an on-ridge or near-ridge hotspot origin (Sen et al., 1988). The upper Dumisseau E-MORB
basalts yield 75.0 Ma radiometric age (K-Ar; Sayeed et al., 1978) and Campanian to Maastrichtian
biostratigraphic ages (Maurrasse et al., 1979a; Desreumaux, 1987). These E-MORB basalts show
geochemical similarities with basalts cored on the southern Beata Ridge (Edgar et al., 1973; Donnelly
et al., 1973) and dredged along the northern Beata Ridge, the latter having 80 – 75 Ma radiometric
ages (Ar-Ar in whole rock and plagioclase; Révillon et al., 2000). Isotopic compositions of the
Dumisseau basalts overlap with those of the eastern Pacific Galápagos and Easter Island hotspots (Sen
et al., 1988). The northern extent of the CLIP-related tholeiitic basalts in Haiti is poorly constrained,
with some authors arguing for a continuation under the Chaîne des Matheux (Boisson and Pubellier,
1987; Vila et al., 1988).

55
Chapter 3: Geological review of Hispaniola

3.2.2 Island arc units related to the Great Arc of the Caribbean
Island arc units outcrop in the Massif du Nord and NW Peninsula in Haiti, and in the Cordillera
Septentrional, Cordillera Oriental and Cordillera Central in the Dominican Republic (Figure 3 - 2/Figure
3 - 3/Figure 3 - 4).

The Massif du Nord in Haiti hosts scattered outcrops of pre-Aptian ultramafic supra-subduction zone
(SSZ) (back-arc, fore-arc, or intra-arc) ophiolitic complexes (Morne Cabril Series; Nicolini, 1977) that
can be correlated with the Loma Caribe Peridotite in the Bonao block in the Dominican Republic. These
Haitian SSZ complexes are overlain by the more coherent IAT Terrier Rouge and 82.1 Ma CA La Mine
formations (K-Ar; Bellon et al., 1985), which are correlated with the lower and upper Tireo Group in
the Dominican Republic, respectively (Lewis et al., 1991). The La Mine Series is also recognized in the
NW Peninsula and Terre Neuve (Lewis et al., 1991). The SSZ, Terrier Rouge and La Mine series in Haiti
are intruded by granodiorites that are dated in the Massif du Nord as 103.5 Ma (Rb-Sr, Limbé batholith;
Cheilletz et al., 1978) and 82.1 to 62.5 Ma (K-Ar, unnamed intrusions; Bellon et al., 1985), while in the
Terre Neuve and NW Peninsula ages of 67.3 Ma (Rb-Sr; Cheilletz et al., 1978), 64.8 Ma (Rb-Sr; Cheilletz,
1976) and 66.2 Ma (K-Ar in biotite; Kesler and Fleck, 1967) are found. The Cordillera Oriental hosts the
118 – 111 Ma (U-Pb in zircon; Kesler et al., 2005; Escuder-Viruete et al., 2006a) Los Ranchos Formation
(LRF). These IAT volcanics (Escuder-Viruete et al., 2006a) extruded in the central part of a volcanic arc
(Kesler et al., 2005) and were subsequently intruded by 106 – 109 Ma batholiths (Ar-Ar in hornblende;
Escuder-Viruete et al., 2006a). The LRF is separated from the Cenomanian high-K CA Las Guayabas
Formation (Mann et al., 1991b; Lebrón and Perfit, 1994) by the unconformably overlying Albian Hatillo
limestones (Myczynski and Iturralde-Vinent, 2005).

The Cordillera Central hosts island-arc related units in four fault-bounded domains. The Eastern
Cordillera Central exposes (1) the Early Cretaceous (Horan, 1995; Draper and Lewis, 2008) Amina-
Maimón schists and metavolcanics, whose IAT mafic protolith (Escuder-Viruete et al., 2002) formed in
a suprasubduction-setting (Draper and Lewis, 1991; Horan, 1995; Andreu et al., 2015), with correlative
outcrops on Tortue Island (Lewis et al., 2002), and (2) The Campanian Peralvillo Sur Formation (Lewis
et al., 2002), consisting of SSZ IAT back-arc basalts (Escuder-Viruete et al., 2008; Nelson et al., 2011).
The Bonao block exposes (1) the Early Cretaceous (125 Ma, U-Pb in zircons; Escuder-Viruete et al.,
2010) Loma Caribe Peridotite, whose lithospheric mantle (Lewis and Jimenéz, 1991; Lewis et al., 2003)
and lower crustal (Escuder-Viruete et al., 2010) sequence is of SSZ geochemistry (Lewis et al., 2005,
2006; Proenza et al., 2007), which is similar to the eastern Cuban IAT ophiolites (Proenza et al., 1999;
Marchesi et al., 2006), and (2) the 120 – 110 Ma (Ar-Ar in hornblende; Escuder-Viruete et al., 2010)
transitional N-MORB to IAT SSZ Rio Verde Complex (Escuder-Viruete and Pérez-Estaún, 2008), which is
geochemically similar to intrusions in the Loma Caribe Peridotite (Escuder-Viruete et al., 2010) and the

57
Chapter 3: Geological review of Hispaniola

A Valanginian - Aptian:
Initial stage of subduction far
allo
np
Oceanic crust lat
e
+ +
SSZ IAT ophiolites inverted + SSZ IAT ophiolites
+ + +
transform
Island-arc crust fault
Back-arc crust Igneous units;
pro - Morne Cabril (SSZ-IAT)
Arc like magmatism to-
car
ibb
- Loma Caribe (SSZ-IAT)
OIB E-MORB
magmatism
en
pla
- Amina-Maimón (SSZ-IAT)
te - Puerto Plata (SSZ-IAT)
Forearc basin
Back-arc basin basalts - Los Ranchos (IAT)
(BABB)-type
B magmatism slab
far ba rollback
+ + Plutonism allo ck-
n/c arc
ari sp
bb rea
Limestones ean din
pla g
Possible plume te +
+
source Aptian - Turonian: +
+
+ + + BON + IAT
Asthenosphere: IAT dominated
melt paths
mantle
Lithosphere:
normal shear pro
zones and faults
to-
car Igneous units;
ibb - Terrier Rouge (IAT)
en
left-lateral strike- pla
slip shears and te - Lower Tireo (IAT)
faults - Las Guayabas (CA)

C Turonian - early Campanian: Forearc basin


slab
CA dominated rollback
car
ibb
ean
pla
te
Onset of extension and rifting
within the Caribbean island-arc + + + +
+
BON+IAT

Igneous units;
- La Mine (CA)
- Upper Tireo (CA) HP-LT complexes
- Peralvillo Sur (IAT) pro
to- of the accretionary
car wedge
ibb
en
pla
te

D SW Cordillera Central NE Middle Campanian - Maastrichtian:


Jicomé Jarabacoa Bonao Tonalite intrusion and plume-related
Pelona-Pico Duarte active arc
thinned back-arc
OIB E-MORB arc basement Siete Cabezas spreading
volcanism
remnant arc OIB E-MORB Cordillera Oriental and
platform eastern Puerto Rico slab
rollback
car
ibb
ean
pla
te
+ + +
+ +

mantle source mantle below


source of the extended arc
magmatism back-arc spreading system extension axis
separated from the
derived components volcanic front

the downgoing slab


Figure 3 - 4: Evolution of geodynamic setting for the Cretaceous igneous units in Hispaniola. Modified from
Escuder-Viruete et al. (2006a, 2007b, 2011a).

58
Chapter 3: Geological review of Hispaniola

Peralvillo Sur Formation (Lewis and Draper, 1995). The Jicomé block hosts the extrusive Albian -
Turonian (Lewis et al., 1991; Montgomery and Pessagno, 1999; Gómez et al., 2000) IAT Lower Tireo
Formation (Escuder-Viruete et al., 2007b) and the 93 – 83 Ma and 70 – 67 Ma CA Upper Tireo
Formation (U-Pb in zircon and Ar-Ar in hornblende; Escuder-Viruete et al., 2007b).

The Jicomé block is intruded in the west by the Loma Cabrera batholith in three distinct magmatic
pulses (102 – 97, 87 – 83 and 68 – 49 Ma, U-Pb in zircon and Ar-Ar in hornblende; Kesler et al., 1991b;
Joubert et al., 2004; Escuder-Viruete et al., 2007c, 2008) with predominantly CA-affinity (Kesler et al.,
1977b), in the central part by the IAT to CA Macutico batholith (93 – 84 Ma, Ar-Ar in hornblende;
Escuder-Viruete et al., 2006b), and in the eastern part by an unnamed foliated tonalitic batholith (63
– 56 Ma, K-Ar in hornblende and biotite; Kesler et al., 1991b). The Jarabacoa block is also intruded by
the Loma Cabrera batholith (Joubert et al., 2004) and by the 70 – 41 Ma tonalitic El Bao batholith (K-
Ar in hornblende and biotite; Kesler et al., 1991b).

During the initial stage of subduction (Figure 3 - 4a) in the Valanginian to Aptian, SSZ boninitic and IAT
ophiolitic units (Morne Cabril, Loma Caribe and Amina-Maimón) formed in the overriding Farallon
plate above the incipient subduction zone (Escuder-Viruete et al., 2006a, 2014). When subduction and
slab-rollback became well established during the Aptian to Turonian (Figure 3 - 4b), island-arc related
igneous units intruded and extruded on the overriding plate (Escuder-Viruete et al., 2007b, 2011a).
Magmatism was predominantly of IAT geochemistry (Lower Tireo Group, Terrier Rouge Formation),
although during the Cenomanian concomitant high-K CA magmatism occurred (Guayabas Formation)
(Figure 3 - 2). From the Turonian onwards slab-rollback resulted in intra-arc rifting and extension with
a significant sinistral strike-slip component (Figure 3 - 4c; Escuder-Viruete et al., 2007b, 2011a, 2014).
CA magmatism dominated (La Mine Series and Upper Tireo Group), with concomitant minor IAT
magmatism (Peralvillo Sur Formation) (Figure 3 - 2). The change from IAT to CA magmatism is gradual
over the Great Arc in Hispaniola, with both types occurring simultaneously. There is no abrupt, discrete
change in the type of magmatism that can be used as evidence to support a flip in subduction polarity
at the Great Arc, as was previously proposed by Burke (1988), Sinton et al. (1998), Kerr et al. (1999),
Lewis et al. (2002), Pindell et al. (2005), and Hastie et al. (2013).

3.2.3 Basalts with a mantle-plume component


The Cordillera Central hosts four formations comprised of OIB-type E-MORB or mantle plume-
influenced basalts, exposed in three domains separated by major fault zones (Figure 3 - 2/Figure 3 -
3/Figure 3 - 4). The Bonao block exposes the N- to E-MORB Loma la Monja Assemblage (LMA) that
originated at a spreading system (Lapierre et al., 1997, 1999), possibly under the influence of a mantle-

59
Chapter 3: Geological review of Hispaniola

plume source (Escuder-Viruete et al., 2007a). Faunal assembly of the associated Oxfordian to Tithonian
El Aguacate Chert points to a Pacific origin (Montgomery et al., 1994a) on the Farallon plate. The
Jarabacoa block exposes OIB-type E-MORB basalts (Sinton et al., 1998; Lewis et al., 2002) of the
Campanian – Maastrichtian (Montgomery and Pessagno, 1999) or 69 – 68 Ma (Ar-Ar in whole-rock and
plagioclase; Sinton et al., 1998) Siete Cabezas Formation, and the plume-influenced basalts of the
Duarte Complex (Draper and Lewis, 1991; Lapierre et al., 1997; Kerr et al., 2003; Escuder-Viruete et al.,
2007c). Oldest ages for the Duarte complex are 115 ± 20 Ma (Sm-Nd in whole-rock; Escuder-Viruete et
al., 2007c), which is intruded by the 90 Ma (U-Pb; Hernaiz Huerta et al., 2000a) to 86 Ma (Ar-Ar in
hornblende; Lapierre et al., 1999) Arroyo Caña batholith. The Duarte Complex basalts have
geochemical similarities with the Dumisseau Formation and basalts cored in the offshore CLIP (Lapierre
et al., 1997, 2000; Escuder-Viruete et al., 2007c). The Pelona-Pico Duarte Basalts (PPDB) are
unconformably overlying the Tireo Formation in the Jicomé block (Escuder-Viruete et al., 2008), whose
OIB-type E-MORB basalts (Kerr et al., 2002) have ages ranging from 79.4 to 68.4 Ma (Ar-Ar in whole-
rock; Escuder-Viruete et al., 2011a).

The origin of the Duarte Complex remains enigmatic. A possible explanation for the mantle-plume
signature of this Aptian-Albian complex can be found in the hypothetical subduction of the Proto-
Caribbean spreading ridge under the island-arc (Escuder-Viruete et al., 2007b; Pindell and Kennan,
2009). The resulting slab-window then provides a pathway for a Caribbean plume-like mantle source
to contaminate the Duarte basalts. A similar setting is possible for the Campanian – Maastrichtian IOB-
type E-MORB basalts of the Siete Cabezas Formation and Pelona-Pico Duarte Basalts (Escuder-Viruete
et al., 2007b). Another possible scenario is by allowing lateral flow of a deeper enriched Caribbean
plume source to produce off-ridge OIB-like magmatism (Figure 3 - 4d; Escuder-Viruete et al., 2011a).

3.2.4 Metamorphism
The Rio San Juan and Samaná complexes in the Cordillera Septentrional consist of HP – LT up to
eclogite-facies serpentinitic mélanges related to subduction at the Great Arc of the Caribbean
(Escuder-Viruete et al., 2011b; c, 2013). These mélanges contain crustal slices from both the overriding
(Farallon/Caribbean) and down-going (Proto-Caribbean) plates, parts of which were accreted in the
subduction channel (Krebs et al., 2008, 2011) and subsequently exhumed (Escuder-Viruete et al.,
2011c; b). These complexes share similarities with the Sierra del Convento mélange in eastern Cuba
(Hernaiz Huerta et al., 2012). The Cordillera Central region in the Dominican Republic is generally
characterized by low-grade, up to greenschist facies metamorphism, and amphibolite facies conditions
restricted to shear zones formed during emplacement of the batholiths. Metamorphism in Haiti is
limited to: (1) calcareous schists on Tortue Island (Woodring et al., 1924) of Mesozoic age (Butterlin,

60
Chapter 3: Geological review of Hispaniola

1960), correlated to the Amino-Maimón Schists (Bowin, 1966; Kesler et al., 1977a), and (2) gneiss, mica
schists, amphibolites and chlorite of unknown age in the Massif du Nord (Bourgueil et al., 1988).

The serpentinite mélange of the Rio San Juan complex in the Cordillera Septentrional contains 136.4
Ma (U-Pb in zircon; Escuder-Viruete et al., 2011d) and 139.7 – 137.8 Ma (U-Pb in zircon; Krebs et al.,
2008) eclogite-facies N-MORB mafic and ultramafic blocks derived from the down-going Proto-
Caribbean plate (Escuder-Viruete et al., 2011d), as well as high-P SSZ IAT mafic blocks from the
overriding Farallon/Caribbean plate (Escuder-Viruete et al., 2011d). It also exposes ultra-HP garnet-
spinel dikes that formed under magmatic conditions of 1550°C and 40 kbar at approximately 125 km
depth in the mantle above the subducting slab, which became subsequently incorporated into the
subduction channel (Figure 3 - 5a; Abbott et al., 2006, 2007). Within the Jagua Clara mélange (Figure 3
- 5a) of the Rio San Juan Complex, the eclogites record peak metamorphism (23 kbar/750 °C) at 103.6
Ma (Lu-Hf in whole-rock) and cooling below 400 °C at 73.4 Ma (Ar-Ar in phengite; Krebs et al., 2008).
Omphacite blueschist (o-bl) records peak metamorphism (17 kbar/520 °C) at 80.3 Ma (Rb-Sr in whole-
rock) and cooling below 400 °C at 73.8 Ma (Ar-Ar in phengite; Krebs et al., 2008). Jadeite blueschist (j-
bl) records peak metamorphism (18 kbar/380 °C) at 62.1 Ma (Rb-Sr in whole-rock; Krebs et al., 2008).

The internal structure of the Samaná Complex consists of an imbricate stack of nappes made up of
high-P pelagic and platform metasediments derived from the overriding plate (Escuder-Viruete et al.,
2011b). Mafic eclogites within the high-pressure Punta Balandra mélange (PBM, Figure 3 - 5a) record
peak metamorphism (23 kbar/620 °C, Escuder-Viruete and Pérez-Estaún, 2006) at 86 – 78 Ma (Sm-Nd
in whole-rock; Escuder-Viruete et al., 2011c, and references therein) and cooling below 400 °C at 35.7
Ma (Ar-Ar in phengite; Escuder-Viruete and Pérez-Estaún, 2004). The Punta Balandra (PB, Figure 3 -
5b) metasediments document the onset of arc-continent collision and continental subduction,
estimated to start at around 65 Ma, reaching peak metamorphism (19 kbar/450 °C) in the early Eocene
(Escuder-Viruete et al., 2011c). The Punta Balandra mélange and metasediments became fully
exhumed to the surface at around 30 Ma (Figure 3 - 5c; Escuder-Viruete et al., 2011b; c). The Santa
Bárbara Schists (SBS) and Rincón Marbles (RM) in the Samaná Complex (Figure 3 - 5b) started to
subduct at around 48 and 35 Ma, respectively. Following peak metamorphism at 37 Ma (14 kbar/450
°C) and 28 Ma (10 kbar/350 °C), respectively, both units were exhumed to the surface in the early
Miocene (Figure 3 - 5c; Escuder-Viruete et al., 2011c). The low-grade Puerto Plata Complex exposes
126 Ma (U-Pb in zircon; Hernaiz Huerta et al., 2012) SSZ ophiolitic fragments (Mann et al., 1991b) of
primitive IAT affinity (Escuder-Viruete et al., 2014). The Puerto Plata complex is relatively undeformed
and non-metamorphic, apart from greenschist-facies shear zones (Hernaiz Huerta et al., 2012) that
record cooling below <450 °C at 90–82 Ma (Ar-Ar in hornblende; Escuder-Viruete et al., 2014) and to
T<150°C at 35 .8Ma (Ar-Ar in plagioclase; Escuder-Viruete et al., 2014).

61
Chapter 3: Geological review of Hispaniola

In the Cordillera Central, syn-obduction metamorphism in the Rio Verde and Amina-Maimón schists
records emplacement of the Loma Caribe Peridotite (Kesler et al., 1991a; Draper et al., 1996) and
closure of the back-arc basin of the Great Arc (Escuder-Viruete et al., 2002). Amphibolite facies syn-
kinematic metamorphism In ductile shear zones in the Loma la Monja and Duarte complex records
emplacement of tonalite plutons (Escuder-Viruete et al., 2007c; a). Regional greenschist
metamorphism possibly records thickening of oceanic crust due to the accumulation of plume-
influenced basalts of the Siete Cabezas and Pelona-Pico Duarte formations (Escuder-Viruete et al.,
2002).

3.2.5 Cenozoic volcanism


Paleocene volcanism of the Baraderes Formation in the northwest Southern Peninsula (Calmus, 1983;
Bourgueil et al., 1988) is interpreted to result from extensional activity (Calmus and Vila, 1988). The
Perodin Formation of Butterlin (1960) consists mainly of transitional to alkaline basalts (D’Meza, 1989)
intercalated with limestones. Ages of these basalts are either 57.1 Ma (NW Peninsula; Bellon et al.,
1985) or late middle Eocene to early late Eocene (Montagnes Noires; Desreumaux, 1987). Similar
alkaline basalts are found in the western extremity of the Chaîne des Matheux (Boisson and Pubellier,
1987), and the middle Eocene El Aguacate Formation in the Sierra de Neiba (Biju-Duval et al., 1982;
Hernaiz Huerta et al., 2007). The structural setting for the Perodin and El Aguacate volcanics is not well
constraint. Bourgueil et al. (1988) propose that the Perodin Formation is extruded in a back-arc
environment, Pubellier et al. (2000) argue that these volcanics are related to fissural activity related to
rifting in the Cayman Trough, while Hernaiz Huerta et al. (2004) propose that the temporally equivalent
El Aguacate Formation results from intra-plate volcanism related to a mantle plume.

Miocene calc-alkaline (CA) basalts in the southwestern part of the Chaîne des Matheux are intercalated
with calcareous or detrital sediments of Tortonian age (Boisson, 1987). Quaternary volcanics SW of the
Cordillera Central are subdivided into two groups based on geochemistry. The first group consists of
mafic alkaline (MA) basalts that are exposed in three volcanic centers (Kamenov et al., 2011a).
Nepheline MA basalts of 2.3 – 1.6 Ma (K-Ar; Wadge and Wooden, 1982; Pubellier et al., 1991) in the
La Vigie and Thomazeau centers in Haiti, and 1.7 – 0.5 Ma limburgite and olivine basalts in the San Juan
center (Wadge and Wooden, 1982; Kamenov et al., 2011a). These basalts are sourced by melts derived
from Proterozoic subcontinental lithospheric (Kamenov et al., 2011a; b). The second group are 2.8 –
0.3 Ma CA basalts (K-Ar; Electroconsult, 1983; Pubellier et al., 1991) that erupted on the SW flank of
the Cordillera Central in Hispaniola (Kamenov et al., 2011a; b). Geochemistry of the CA lavas indicates
mixing of (1) the MA source and (2) a mantle source modified by subduction of proto-Caribbean
oceanic crust (Kamenov et al., 2011a; b).

63
Chapter 3: Geological review of Hispaniola

3.3 Tectono-stratigraphy of western Hispaniola

This section covers the general stratigraphic trends and formations that are found in western
Hispaniola. A generalized tectonostratigraphic chart is presented in Figure 3 - 6 and a simplified
geological map in Figure 3 - 7. Corresponding morphotectonic terranes and fault zones are found in
Figure 3 - 1. A detailed account of the stratigraphy of the Southern Peninsula is discussed in Chapter 4
and for the Chaîne des Matheux in Chapter 5.

3.3.1 Late Cretaceous


In the Massif du Nord (Haiti) and the Cordillera Central (Dominican Republic), island arc-related rocks
are unconformably overlain by strongly folded middle Campanian – Maastrichtian flysch-like rocks of
the Trois Rivières Formation (TRF) (Boisson and Vila, 1982; Heubeck et al., 1991). In the western
Cordillera Central the TRF is conformably overlain by upper Maastrichtian limestones of the Bois de
Laurence Formation (Figure 3 - 7; Escuder-Viruete et al., 2007b). In the Terre Neuve region (Haiti) the
TRF is dated Albian – Cenomanian by Ayala (1959), although the biostratigraphy is imprecise (Bourgueil
et al., 1988). The TRF is part of the Peralta fold-and-thrust belt, whose evolution is characterized by SE-
wards along-strike migration of the depocenter resulting from progressive closure and uplift of the
basin from NW to SE over time (Heubeck et al., 1991).

On the Southern Peninsula in Haiti, the Cretaceous sedimentary cover corresponds to the deformed
pelagic limestones of the Macaya Formation (Butterlin, 1954, 1960; Calmus and Vila, 1988). Deposition
occurred from Coniacian (Maurrasse, 1982; Calmus, 1983) to middle Maastrichtian (Maurrasse et al.,
1979b; Van den Berghe, 1983a; Bien-Aime Momplaisir, 1986; Desreumaux, 1987). Macaya-like
limestones are also found in the western extremity of the Chaîne des Matheux (Vila et al.. 1988).

3.3.2 Paleocene
The Paleocene is mainly outcropping on the Southern Peninsula. Its existence elsewhere is only
confirmed in the Artibonite-I well offshore western Haiti (Figure 3 - 7; Bourgueil et al., 1988), and the
Padre Las Casas area of the eastern Peralta Belt in the Dominican Republic (Figure 3 - 7; Heubeck et
al., 1991). The Artibonite-1 (Crux) well penetrated Paleocene conglomerates sourced from the island-
arc terranes in the east (Bourgueil et al., 1988), which have similarities with post-Paleocene
conglomerates in the Terre Neuve area (Butterlin, 1960; Cheilletz and Lewis, 1976). The Padre Las
Casas area exposes Paleocene – lower Eocene turbiditic deposit (Heubeck et al., 1991).

64
Chapter 3: Geological review of Hispaniola

The Paleocene on the Southern Peninsula locally records a change in depositional environment from
Maastrichtian bathyal pelagic limestones of the Macaya Formation to lower Paleocene (volcani)clastic
shallow water to turbiditic slope deposits of the Rivière Glace (Calmus, 1983), Marigot (Butterlin, 1954;
Desreumaux, 1987), and Beloc Formations (Maurrasse et al., 1979b; Maurrasse, 1980). In the western
Massif de la Selle the transition from Maastrichtian to Paleocene is marked by a ravinement surface
(Desreumaux, 1985a; Bourgueil et al., 1988). The Paleocene on the Southern Peninsula generally
records a regressive trend from Maastrichtian pelagic slope limestones (Biju-Duval et al., 1980;
Andreieff and Desreumaux, 1984) to platform carbonates in the late Paleocene (Calmus, 1983) or early
Eocene (Bourgueil et al., 1988).

3.3.3 Eocene
Massive platform carbonates dominate most of Haiti during the early Eocene, with the exception of
the southern part of the Southern Peninsula (Butterlin, 1960; Vila and Feinberg, 1982). This part is
characterized by deeper pelagic chalky and cherty limestones that become dominant over the entire
Southern Peninsula during the middle and late Eocene (Biju-Duval et al., 1980; Van den Berghe, 1983b;
Bourgueil et al., 1988). Lower Eocene deep marine bathyal clastic limestones are locally found along
the Plateau Central (Bermudez, 1949) and NW Massif du Nord (Desreumaux, 1987) in Haiti, and the
Cordillera Septentrional (de Zoeten and Mann, 1999) in the Dominican Republic. These deposits display
similarities with the Paleocene Marigot Formation on the Southern Peninsula (Bourgueil et al., 1988).
Turbiditic sandstones with island-arc material derived from the Tireo Formation observed in the lower
Eocene stratigraphy of the Peralta Belt (Dolan et al., 1991; Hernaiz Huerta and Pérez-Estaún, 2002).
Lower Eocene conglomerates are drilled in the offshore Artibonite-1 (Crux) well in western Haiti
(Bourgueil et al., 1988).

The middle – upper Eocene in central and northern Haiti consists of relatively homogeneous facies
with a general deepening trend. Middle Eocene massive foraminiferal platform carbonates grade both
laterally and stratigraphically upwards into deeper upper Eocene chalky and cherty limestones. This
trend is documented in the NW Peninsula and Terre Neuve (Bourgueil et al., 1988), Ennery Valley and
Trois Rivières Basin (Maurrasse, 1982), and the Plateau Central – San Juan Basin and Chaîne des
Matheux (Bourgueil et al., 1988). Instabilities within the platform are recorded by middle to late
Eocene conglomeratic Mass Transport Deposits (MTD’s) in the Northern Plain (Vila et al., 1987), NW
Peninsula (Pubellier et al., 1991), Ennery Valley (Maurrasse, 1982) and Artibonite-1 (Crux), with local
uplift recorded by a shallowing upwards trend in the Cul-de-Sac-1 (Crux) well (Bourgueil et al., 1988).

65
Chapter 3: Geological review of Hispaniola

A diachronism in the timing of change from platform to pelagic slope carbonates is observed in Haiti,
which migrates from south to north. The southernmost part of the Southern Peninsula remains entirely
pelagic throughout the Eocene. The northern part of the Southern Peninsula experiences a change
from platform to pelagic facies at the early to middle Eocene boundary, while the same change occurs
in central and northern Haiti at the middle to late Eocene boundary.

3.3.4 Oligocene
Prior to the 1980’s, stratigraphic studies mainly relied on large benthic foraminifera to date Oligocene
sediments. Using these fossils to distinguish between the Oligocene and Miocene proved to be very
difficult, if not impossible, in the Caribbean realm (Butterlin, 1960). After the 1980’s, the use of high
resolution planktonic microfauna permitted the distinction between the Oligocene and the Miocene
(Andreieff, 1981, 1983, 1985; Andreieff et al., 1987; Butterlin, 1981).

In the northern Dominican Republic, the lower Oligocene is strongly folded in the Cordillera
Septentrional (Dolan et al., 1991; de Zoeten and Mann, 1999) but undeformed in the Cibao Basin
(Dolan et al., 1991; Edgar, 1991). In these regions it consists of a basal conglomerate overlain by
turbidites, both of which are sourced from the Duarte and Tireo Formations in the Cordillera Central
(Biju-Duval et al., 1982). The unconformably overlying upper Oligocene is composed of clastic and
calcareous turbidites (Boisseau, 1987; Saunders et al., 1986). The Hispaniola Fault Zone (HFZ) to the
southwest of the Cibao basin became active in the early Oligocene (Bowin, 1975), as evidenced by the
infill of pull-apart basins along its trace (Coleman and Winslow, 1999).

The San Juan Basin in the Dominican Republic is also characterized by Oligocene turbiditic facies, which
are unconformably overlying the basement (Biju-Duval et al., 1982). In the Trois Rivières Basin and
Northern Plain in Haiti, the Oligocene consists of clastic and argillaceous limestones (Vila et al., 1986b).
From the San Juan Basin towards the south and west, the Oligocene is characterized by exclusively
calcareous facies; massive limestones in the Plateau Central and well stratified chalky and cherty
limestones in the Chaîne des Matheux and Sierra de Neiba (Butterlin, 1960; Desreumaux, 1987;
Bourgueil et al., 1988). The Oligocene age attributed to limestones in the Chaîne des Matheux and
Sierra de Neiba is questionable, and the Oligocene possibly missing (Van den Berghe, 1983a; Andreieff
et al., 1987). The Oligocene – Pliocene is absent in the Artibonite-I (Crux) well offshore western Haiti,
with Pleistocene reefal limestones resting directly on Eocene limestones (Bourgueil et al., 1988).

The Southern Peninsula is characterized by continued pelagic chalky limestone deposition, with a
locally subtidal depositional environment in the northern part and exclusively pelagic conditions in the
south (Desreumaux, 1987), suggesting a period of relative tectonic quiescence. In the central Massif

67
Chapter 3: Geological review of Hispaniola

de la Hotte (Camp Perrin) region the Oligocene has a minor basal unconformity (Calmus, 1983;
Bourgueil et al., 1988) and was prone to local instabilities resulting in deposition of thin turbidite beds
(Maurrasse, 1980).

3.3.5 Miocene
In the northern Dominican Republic, the early Miocene records the onset of subsidence in the Cibao
Basin (Erikson et al., 1998). It consists of a mildly faulted and northwards-thickening (Edgar, 1991)
sequence of shallow marine sediments (Vokes, 1979; Saunders et al., 1986). Infill of the basin is related
to incipient activity on the SFZ (Erikson et al., 1998). Continued subsidence and infill of the Cibao Basin
throughout the Miocene is recorded by; 1) middle to upper Miocene shallow-water limestones in the
east (Mann et al., 1991b), 2) upper Miocene platform carbonates in the west (de Zoeten and Mann,
1999), and 3) upper Miocene (Palmer, 1979) littoral conglomerates and siltstones (Biju-Duval et al.,
1980) that unconformably onlap older formations to the southwest of the Cibao Basin (Saunders et al.,
1986; Erickson et al., 1998).

To the SW of the Cibao Basin activity on the HFZ became transpressive, resulting in uplift of the
northern Cordillera Central (Lewis et al., 2006). En-echelon faults on the northern side of the HFZ
indicate that fault movement continued through the late Miocene and possibly early Pliocene
(Coleman and Winslow, 1999). Offset of the Amina-Maimón schists suggests at least 120 km of lateral
displacement along the HFZ (Draper, 1999). In the late Miocene the locus of strike-slip deformation
jumped northwards onto the Septentrional Fault system (Draper, 1999).

Sedimentation in northern and central Haiti from the early Miocene onwards evolved from calcareous
to more detrital, recording a progressive decrease in depositional depth and increase in basin infill
(Bourgueil et al., 1988). This diachronous evolution propagated from northeast to southwest and
records the activation of thrust faults in the Haitian fold-and-thrust belt (Pubellier et al., 2000). The
upwards coarsening sedimentary sequence associated with this evolution consists of three formations,
which are; 1) the Madame Joie Formation, characterized by a succession of calcareous marls and flysch,
which records the first occurrence of clastics derived from the uplifted thrust sheets, which is overlain
by 2) the Thomonde Formation, containing calcareous silts and sandstones, in turn overlain by 3) the
Las Cahobas Formation, consisting of a coarse detrital sandstones and conglomerates (Woodring et
al., 1924; Van den Bold, 1981; Dubreuilh, 1982; Desreumaux, 1987). Deposition of the Madame Joie
Formation occurred in the Trois Rivières Basin and northern Plateau Central during the Burdigalian,
the southern Plateau Central in early Langhian, the Montagnes Noires and Artibonite Valley during the
Serravallian, the northern Chaîne des Matheux in the Tortonian, and the Cul-de-Sac plain in the latest

68
Chapter 3: Geological review of Hispaniola

Tortonian (Van den Bold, 1981; Dubreuilh, 1982; Feinberg and Vila, 1982; Vila et al., 1986b; Bourgueil
et al., 1988; Pubellier et al., 1991, 2000).

On the Southern Peninsula the lower and middle Miocene is represented by almost exclusively pelagic
limestones deposited under deep marine basinal conditions. Calcareous – silty flysch-type Madame
Joie facies are recorded in the Sierra de Bahoruco during the Tortonian (Maurrasse et al., 1980, 1979a).
In the Massif de la Selle, facies similar to the Madame Joie Formation are termed the Rivière Gris
Formation by Butterlin (1950), which arrive in the northern Massif de la Selle during the Messinian
(Andreieff et al., 1987; Desreumaux, 1987), and in the southern Massif de la Selle during earliest
Pliocene (Van den Berghe, 1983a; Desreumaux, 1987). Other regions in the Southern Peninsula
experienced pelagic limestones sedimentation throughout the Miocene. Exceptions to this are the
Burdigalian to Serravallian continental Camp Perrin (Calmus, 1983; Bizon et al., 1985) and L’Asile basins
(Bien-Aime Momplaisir, 1986) in the western Southern Peninsula. Bourgueil et al. (1988) dispute this
early Miocene age arguing that biostratigraphic dating is probably performed on samples containing
reworked material.

3.3.6 Pliocene – Recent


In the late Miocene in the northern Dominican Republic, the SFZ undergoes major strike-slip motion
coupled with regional subsidence in the Cibao Basin, which continued until early Pliocene times (~11
– 4 Ma). During the early Pliocene (~4 – 2.5 Ma), compression and uplift of the Cordillera Septentrional
resulted in accelerated subsidence of the Cibao Basin, which was followed by late Pliocene to recent
uplift (Edgar, 1991; Erikson et al., 1998). Lateral strike-slip related offset along the SF is estimated at
85 km (Erikson et al., 1998). The SFZ, passing to the south of Tortue Island (Figure 3 - 1), became active
after a southward jump from a currently inactive strike-slip segment north of Tortue (Erikson et al.,
1998). The onset of motion along this active southern segment is estimated at 1.8 Ma, which is based
on the 16.5 km offset of incising valleys to the southeast of Tortue Island coupled with present-day
modelled GPS velocities (Leroy et al., 2015).

During the Pliocene Haiti is subject to widespread terrigenous sedimentation (Van den Bold, 1981;
Maurrasse, 1982). Deposition of coarsening-upwards Pliocene littoral sediments continued into the
Pleistocene on the southern part of the Southern Peninsula (Jean Poix, 1980; Desreumaux, 1987). The
source for the detrital fractions differs between northern and southern Haiti. In the Plateau Central it
consists of alkali feldspars sourced from the Cordillera Central, while in the Massif de la Selle the
detrital material is Ca-plagioclase (labradorite) sourced from eroding CLIP basalts (Desreumaux, 1987;

70
Chapter 3: Geological review of Hispaniola

Bourgueil et al., 1988). Most of the Southern Peninsula was emerged by Pliocene times, with
continental deposition in onshore strike-slip basins (Calmus, 1983; Mann et al., 1983; Bien-Aime
Momplaisir, 1986) .

Uplifted Pleistocene to Present-day reefal limestone terraces can reach elevations of some 600m, and
are found on Tortue Island, the NW Peninsula, Gonâve Island, and the western extremity of the Chaîne
des Matheux (Saint Marc Peninsula) (Mann et al., 1995). Using the elevations of <133Ka (Th-U on
aragonite, Dodge et al., 1983; Mann et al., 1995) raised terraces, Mann et al. (1995) calculated
resurrection rates for the NW Peninsula at 0.37 mm yr-1, the Saint Marc Peninsula at 0.19 mm yr-1, and
Gonâve Island at 0.05 mm yr-1.

3.3.7 Evolution of the Peralta Belt


The Peralta Belt spans the whole island from NW to SE (Figure 3 - 7) and evolved from a back-arc basin
for the Great Arc, with deposition of latest Cretaceous to Eocene turbidites, through an Oligocene
strike-slip phase, to a transpressive fold-and-thrust belt in Miocene times (Pindell and Barrett, 1990;
Witschard and Dolan, 1990; Dolan et al., 1991; Hernaiz Huerta and Pérez-Estaún, 2002). Progressive
NW closure of this basin over time causes a shift of its depocenter towards the SE (Heubeck et al.,
1991). From NW to SE the belt becomes progressively wider, younger, less metamorphosed, less folded
and faulted and lower in elevation (Heubeck et al., 1991). The Los Pozos – San Juan Fault Zone (LPSJFZ)
delineates the southwestern boundary of the Peralta Belt, which on the northern side is marked by
the San José – Restauración Fault Zone (SJRFZ) (Mann et al., 1991b). Activity on the LPSJFZ started in
the early (?) to late Eocene with syn-sedimentary deformation related to closure of the basin
(Witschard and Dolan, 1990; Hernaiz Huerta and Pérez-Estaún, 2002). During late Eocene to early
Miocene sedimentation to the north of the LPSJFZ records abrupt lateral facies change, extremely rapid
basin subsidence and complex paleoflow patterns, interpreted to result from deposition in basins
within a broad strike-slip system (Dolan et al., 1991). Based on the facies discrepancy in the Peralta
belt in the north and the San Juan Basin in the south, Pindell and Barrett (1990) propose 350 km of
lateral offset along the LPSJFZ during late Eocene – early Miocene times. The system becomes
transpressive in the early Miocene, with the Peralta belt overthrusting the San Juan Basin to the south.
This is postulated to be in response to the collision between CLIP-dominated southern Hispaniola and
island-arc terranes of northern Hispaniola (Heubeck and Mann, 1991b). A similar timing is noted for
south-verging thrust along the southern border of the Massif du Nord, where they affect the Miocene
series of the Plateau Central (Berthier and Janjou, 1984; Boisson and Vila, 1984; Boisson, 1987). Late
Miocene uplift of the Cordillera Central is concomitant with uplift in the Sierra de Neiba, with both
ranges providing a sediment source for the eastern San Juan Basin (McLaughlin and Sen Gupta, 1991).

71
Chapter 3: Geological review of Hispaniola

3.4 Deformation phases

The deformation phases that affected Hispaniola during the Cenozoic are summarized and discussed
in this section. The main events are listed in the tectonostratigraphic chart in Figure 3 - 6. The
morphotectonic terranes are located on Figure 3 - 1, while the geological formations are show on the
simplified geological map in Figure 3 - 7.

3.4.1 Laramide phase


In northern and central Haiti, the Laramide tectonic phase is recorded by a regional unconformity
separating the Cretaceous basement from overlying Paleocene conglomerates and basal Eocene
deposits in the Montagnes Noires, NW Peninsula, Terre Neuve, Massif du Nord and Tortue Island
(Bourgueil et al., 1988), and at the base of the NW Peralta Belt (Butterlin, 1960). This deformation was
accompanied by low-grade metamorphism, the formation of a schistosity and kink-folding (Boisson
and Vila, 1984; Michel, 1986), which is also observed in the Cordillera Central in the Dominican
Republic (Vila and Feinberg, 1982).

On the Southern Peninsula, the existence of a widespread Laramide phase of deformation mainly relies
on the occurrence of angular unconformities separating the Cretaceous from the Paleocene (Mercier
de Lépinay et al., 1979; Calmus, 1983; Van den Berghe, 1983a; Bien-Aime Momplaisir, 1986). In certain
regions on the Southern Peninsula the existence of these unconformities is however questioned
(Maurrasse, 1980; Desreumaux, 1987). Chapter 4 is partly dedicated to resolving the problem of the
existence, extent, and intensity of this Laramide deformation phase on the Southern Peninsula.

3.4.2 Eocene deformation


In northwest Haiti, instabilities within the Eocene carbonate platform are recorded by the occurrence
of NW – SE trending chaotic sedimentary complexes (Vila et al., 1987). In the Northern Plain and Terre
Neuve these are related to strike-slip tectonics (Vila et al., 1987), while in the Ennery Valley the Perodin
Formation is affected by middle Eocene reverse thrusts related to transpressive displacements
(Pubellier et al., 1991).

Contradictions about the geodynamic interpretation of the Perodin and El Aguacate Formation basalts
are already pointed out in section 3.2.5. Pubellier et al. (1991, 2000) argue that these volcanics result
from fissural activity related to rifting in the Cayman Trough. The trend of these basalts – limestone
intercalations appears to follow the main NW – SE trend of the Haitian fold-and-thrust belt (Vila et al.,
1987), as can be deduced from the geological map (Figure 3 - 7; Boisson and Pubellier, 1987). This

72
Chapter 3: Geological review of Hispaniola

trend is however perpendicular to the NE – SW trending tilted fault blocks observed in the eastern part
of the Cayman Trough (Leroy et al., 2000). The predominantly middle Eocene age for the basalts of the
Perodin and El Aguacate Formations (Biju-Duval et al., 1982; Bourgueil et al., 1988; Hernaiz Huerta et
al., 2007) also post-dates rifting in the Cayman Trough, since spreading initiates at 49 Ma (Leroy et al.,
2000). It is therefore unlikely that these basalts are related to rifting in the Cayman Trough. Hernaiz
Huerta et al. (2004) propose that the Perodin and El Aguacate Formations result from intra-plate
volcanism related to a mantle plume. However, the ages and geochemical data provided by Hernaiz
Huerta et al. (2007) differs significantly from other plume-related basalts in the Caribbean, all of which
are of pre-Cenozoic age (Sinton et al., 1998; Kerr et al., 2003; Escuder-Viruete et al., 2011a). Bourgueil
et al. (1988) propose that the Perodin Formation is extruded in a back-arc environment. Closing of the
back-arc basin of the Great Arc (Peralta Belt in Hispaniola) is concomitant with extrusion of the Perodin
and El Aguacate Formations (Biju-Duval et al., 1982; Bourgueil et al., 1988; Witschard and Dolan, 1990;
Hernaiz Huerta and Pérez-Estaún, 2002; Hernaiz Huerta et al., 2007). Therefore it is possible that the
Perodin and El Aguacate Formations record the latest phase of back-arc magmatism during the arrest
of subduction in northern Hispaniola in the early to middle Eocene (Lidiak and Jolly, 1996; Rojas-
Agramonte et al., 2004; García-Casco et al., 2008).

3.4.3 Oligocene - Miocene deformation


Several intra-Oligocene and early to middle Miocene unconformities are found in Hispaniola, most of
which appear to be related to local events (Bourgueil et al., 1988). A base Oligocene (angular)
unconformity is encountered on the southern side of the Montagnes Noires (Butterlin, 1956), the Trois
Rivières Basin (Vila et al., 1986a, b), and on the borders of the Plateau Central (Dubreuilh, 1984;
Boisson, 1987), the latter related to minor compressive tectonics (Dubreuilh, 1984). In the Chaîne des
Matheux it is unclear if the Oligocene is (partly) eroded or not (Van den Berghe, 1983b; Bourgueil et
al., 1988). A middle Miocene unconformity is recognized in the Terre Neuve region (Pubellier et al.,
1991), which in the Trois Rivières Basin represents a ravinement surface (Vila et al., 1986b). An intra-
Miocene deformation event is also postulated for the northern part of the Plateau Central (Boisson,
1987). The existence and extent of an early Miocene deformation phase affecting the Southern
Peninsula is addressed in Chapter 4.

3.4.4 Miocene – Recent deformation


The Miocene marks the epoch in which the major structures that characterize Haiti today started to
develop. The timing, structural style and evolution of the EPGFZ on the Southern Peninsula is discussed
in chapter 4. The evolution of the Haitian fold-and-thrust belt is extensively discussed in chapter 5.

73
Chapter 3: Geological review of Hispaniola

74
Chapter 4: Geology of the Southern Peninsula

sedimentation in pull-apart basins along its trace (Mann et al., 1983; Calmus, 1983), left-lateral offset
of Quaternary alluvium (Mann et al., 1995) and left-lateral offset of stream channels (Prentice et al.,
2010).

This chapter focusses on the structural and stratigraphic evolution of the Southern Peninsula from the
Late Cretaceous through the Cenozoic era. It addresses the number of deformation phases, the
relation between sedimentation and deformation, the responsible paleo-stress patterns and structural
control on deformation, timing and development of the EPGFZ, and the geometry of the EPGFZ south
of the Cul-de-Sac plain (Figure 4 - 1). The results are presented in a paper submitted to Tectonophysics,
which is included in section 4.2 of this chapter.

The structural and temporal constraints provided in the submitted paper (section 4.2) are used in the
discussion on the palinspastic position of the Southern Peninsula at the Cretaceous – Paleocene
boundary and the amount of offset along the EPGFZ.

The cumulative amount of left-lateral displacement of the Southern Peninsula with respect to central
Haiti and the amount of shortening is addressed in the discussion in Chapter 7. In the same chapter
the paleogeographic evolution of the Southern Peninsula through the Miocene is discussed.

Stratigraphic observations on the Southern Peninsula that are used as an input for the stratigraphic
columns shown in the submitted paper “Polyphase tectonic history of the Southern Peninsula, Haiti:
From folding-and-thrusting to transpressive strike-slip” are found in Appendix 1.

76
Chapter 4: Geology of the Southern Peninsula

4.2 Submitted paper to Tectonophysics

Polyphase tectonic history of the Southern Peninsula, Haiti: From folding-and-


thrusting to transpressive strike-slip

Richard J.F. Wessels1, 2, Nadine Ellouz-Zimmermann2, Nicolas Bellahsen1, Youri Hamon2, Claudio

Rosenberg1, Remy Deschamps2, Roberte Momplaisir3, Dominique Boisson3, Sylvie Leroy1

1. Sorbonne Université, CNRS-INSU, Institut des Sciences de la Terre Paris, ISTeP UMR 7193, F-

75005 Paris, France.

2. IFP Energies nouvelles, IFPEn, Geosciences, Rueil-Malmaison, France.

3. Université d’Etat d’Haïti (UEH), Port-au-Prince, Haiti.

Abstract

The Southern Peninsula of Haiti is a seismogenically active east – west trending transpressive structure.

Present-day deformation is mainly partitioned along the left-lateral Enriquillo – Plantain Garden Fault

Zone (EPGFZ) and associated oblique reverse faults and/or folds. The configuration of these faults,

their respective timing, and the role of structural heritage on their development is poorly understood.

To address these questions we present the results of extensive field campaigns, combined with

satellite imagery interpretation and a compilation of literature data, which allows us to constrain the

Cenozoic evolution of the Southern Peninsula. Our results show a polyphase tectonic history consisting

of three major tectonic events: 1) Maastrichtian to early Paleocene crustal-scale folding that developed

coevally with predominantly south-to-southwest verging thrusts, which resulted in uplift and erosion

of the Cretaceous sedimentary cover in large parts of the Southern Peninsula, 2) early Miocene uplift

and erosion, which was strongest in the southwestern part of the peninsula and decreased eastwards,

leaving the Massif de la Selle unaffected. Folding in response to NE – SW shortening was possibly

amplified by the reactivation of older thrust faults. This deformation phase was likely related to a

77
Chapter 4: Geology of the Southern Peninsula

contemporaneous reconfiguration of the plate boundaries in the northern Caribbean, 3) late Miocene

to present-day deformation and uplift. Spatially distributed strike-slip that during the late Miocene

became progressively focused along the EPGFZ in the latest Miocene. Oblique and thrust faults locally

post-date strike-slip activity from the Pliocene onwards. Increase in compressional deformation from

west to east is reflected by a change in structural style, with predominantly strike-slip faults in the west

and oblique-slip faults in the east, the latter rooted on the EPGFZ at depth. Paleo-stresses associated

with a strike-slip regime are at a significantly high angle to the trace of the EPGFZ, indicating that the

EPGFZ is a mechanically weak fault.

Keywords

Southern Peninsula Haiti, EPGFZ, folding-and-thrusting, transpressive strike-slip, paleo-stress

1. Introduction

Haiti is located on the western side of the island of Hispaniola and occupies a central position in the

northern Greater Antilles arc in the Caribbean (Fig. 1a). The Southern Peninsula of Haiti experienced

devastating earthquakes in the past (McCann, 2006; Ali et al., 2008; Bakun et al., 2012). These historical

earthquakes occurred proximal to a N80°E trending onshore lineament stretching from Tiburon in the

west up to Pétionville in the east (Fig. 1b). This segmented lineament is easily traced on satellite

images and is known as the Enriquillo – Plantain Garden Fault Zone (EPGFZ), left-lateral strike-slip

system (Fig. 2a; Mann and Burke, 1984) (Fig. 2a; Mann and Burke, 1984). The seismogenic nature of

the Southern Peninsula was once more demonstrated by the M7.0 2010 Leogâne earthquake that

struck close to the capital of Port-au-Prince (Hayes et al., 2010; Bilham, 2010b). This event likely did

not involve slip along the steeply south-dipping EPGFZ segment in this region (Prentice et al., 2010),

78
Chapter 4: Geology of the Southern Peninsula

Miocene deformation phase did (Calmus, 1983; Van den Berghe, 1983a; Bizon et al., 1985; Bien-Aime

Momplaisir, 1986) or did not (Desreumaux, 1985b, 1987; Bourgueil et al., 1988) affect the southern

Peninsula and if it is related to strike-slip activity along the EPGFZ (Calmus, 1983; Bizon et al., 1985;

Calmus and Vila, 1988). The configuration and timing of faults and deformation in the eastern Southern

Peninsula is also subject to debate. Some authors argue in favor of a continuation of the EPGFZ into

the Cul-de-Sac plain (Mann et al., 1995) while others propose that active thrust faults along the

southern border of the Cul-de-Sac truncate the EPGFZ (Bourgueil et al., 1988; Symithe and Calais,

2016), or that normal faulting dominates the southern border (Pubellier et al., 2000).

The aim of this paper is to understand the structural evolution of the Southern Peninsula during the

Cenozoic era and the temporal and spatial evolution of the EPGFZ and associated faults. Of particular

interest are the next questions: 1) how many episodes of deformation has the Southern Peninsula

experienced during the Cenozoic, 2) what was the structural style associated with the successive

deformation phases, and 3) how are paleo-stresses distributed along the EPGFZ and what is their

timing.

Three study areas are selected, based on results from field campaigns conducted in 2014, 2015, and

2017. These are the western Massif de la Hotte, l’Asile Basin, and Massif de la Selle regions of the

Southern Peninsula (Fig. 2b). Detailed mapping of these three areas is based on the interpretation of

satellite imagery, which is controlled and tested against field data. Stratigraphic columns derived from

literature are updated with field observations and serve as a first-order constraint for our cross

sections. Paleostresses are deduced from the inversion of fault kinematic slip data obtained at

outcrops. The before mentioned data allows us to present a coherent structural model for the

evolution of the Southern Peninsula, which shows a polyphase tectonic history with deformation

phases in the latest Cretaceous - Paleocene, early Miocene, and late Miocene – present-day.

80
Chapter 4: Geology of the Southern Peninsula

2. Geological setting

2.1 Geodynamic setting

Haiti is located in a geodynamically complex area, at the boundary between the North American and

Caribbean plates (Fig. 1a). This boundary zone comprises an amalgamation of microplates and blocks

bounded by fault zones (Fig. 1b; Benford et al., 2012a; Calais et al., 2016).

The northern plate boundary between the Gonâve and North American plates is defined by the left-

lateral Oriente Fault Zone (OFZ), which passes through the Santiago Deformed Belt (SDB) south of Cuba

to link up with the Septentrional Fault Zone (SFZ) in Hispaniola (Fig. 1b; Heubeck and Mann, 1991).

The boundary between the Gonâve and Hispaniola blocks possibly corresponds to the northeastern

boundary of the Haitian fold-and-thrust belt (Fig. 1b; Benford et al., 2012a; Symithe et al., 2015), which

is a southwest-verging, forward propagating thrust system (Pubellier et al., 2000). The southern

boundary between the Gonâve microplate and the Caribbean plate coincides with the Walton Fault

Zone (WFZ) - Enriquillo-Plantain Garden Fault Zone (EPGFZ) system (Fig. 1b; Benford et al., 2012a;

Symithe et al., 2015). In the west, the WFZ marks the southern border of the Cayman Trough

(Rosencrantz and Mann, 1991; Leroy et al., 2000), which is connected with the EPGFZ through Jamaica

(Benford et al., 2015). East of Jamaica the EPGFZ crosses the Jamaica Passage (Leroy et al., 2015;

Corbeau et al., 2016b) and continues onshore the Southern Peninsula (Duplan, 1975; Calmus, 1983;

Van den Berghe, 1983b; Mann et al., 1995; Prentice et al., 2010). In the eastern Southern Peninsula

the EPGFZ either continues through the Cul-de-Sac plain (Mann et al., 1995) or is abutting against E-

to ESE-trending thrust faults (Fig. 2a; Saint Fleur et al., 2015; Symithe and Calais, 2016).

Present-day deformation along the northern Caribbean plate boundary changes from almost pure

strike-slip along the WFZ and the Jamaica Passage in the west (Benford et al., 2012a; Leroy et al., 2015;

Symithe et al., 2015; Corbeau et al., 2016a), to increasingly more transpressive along the EPGFZ in the

Southern Peninsula in the east (Ali et al., 2008; Calais et al., 2016).

82
Chapter 4: Geology of the Southern Peninsula

2.2 Stratigraphy of the Southern Peninsula

The distribution of the formations described below is found on 1:250.000 geological maps (Published

by the Bureau des Mines et de l'Energie d'Haïti (BME) after Boisson and Bien-Aime Momplaisir, 1987;

Bien-Aime Momplaisir et al., 1988), which are simplified and presented in Fig. 2b. The purpose of this

section is to describe the general stratigraphy of the Southern Peninsula. A corresponding schematic

stratigraphic column is presented in Fig. 3. A detailed discussion of our stratigraphic observations is

given in section 4.

The oldest rocks exposed on the Southern Peninsula are Cretaceous in age and composed of an

alternation of tholeiitic basalts, limestones and radiolarites (Fig. 3; Dumisseau Formation of Maurrasse

et al., 1979). This succession is part of the Caribbean Large Igneous Province (CLIP) that extruded in a

submarine setting (Sinton et al., 1998). The lowermost part contains Aptian – Albian rudists indicative

of platform facies (Bien-Aime Momplaisir, 1986; Bourgueil et al., 1988). Laterally and upwards in the

stratigraphy pelagic facies dominate until late Campanian times (Maurrasse et al. (1979), in Bourgueil

et al. (1988); Mann et al., 1991). The thickness of this basaltic complex is at least 1500m (Bourgueil et

al., 1988), with a similar thickness estimated offshore using seismic refraction (Mauffret and Leroy,

1997) and submersible sampling (Mauffret et al., 2001). The Cretaceous sedimentary cover

corresponds to the calcareous Macaya Formation of Butterlin (1954). It consists of well stratified

pelagic limestones with chert nodules and occasional claystone and radiolarite beds deposited under

bathyal conditions (Calmus, 1983). The age of this formation is well constrained from Campanian to

middle Maastrichtian, although the base could be located in the Coniacian (Ayala, 1959; Maurrasse,

1982; Calmus, 1983; Van den Berghe, 1983; Desreumaux (1987), in Bourgueil et al., 1988). The facies

indicate a general deepening of the depositional environment throughout the Late Cretaceous.

83
Chapter 4: Geology of the Southern Peninsula

The Cretaceous basalts and limestones are unconformably overlain by a diachronous transgressive

fining-upwards sequence that becomes more calcareous through time (Bourgueil et al., 1988). This

Paleocene sequence commences with shallow water conglomerates and volcaniclastic breccias, clay-

silt- and sandstones overlain by turbiditic silty limestones (Van den Berghe, 1983a; Bourgueil et al.,

1988; Amilcar, 1997), and locally possibly platform limestones (Calmus, 1983). The Paleocene

sequence is occasionally interbedded with transitional to calc-alkaline basalts (Calmus, 1987). The

temporal and lateral equivalents of the Paleocene detrital series on the Southern Peninsula are known

by the following formation names; Marigot in the south-east (Butterlin, 1960), Beloc in the central-

east (Desreumaux, 1987) and Rivière Glace in the west (Calmus, 1983). Locally the transition from

Maastrichtian to Paleocene is conformable and characterized by an upper Maastrichtian ravinement

surface (Desreumaux, 1985a; Bourgueil et al., 1988) or K-T impact ejecta material (Maurrasse and Sen,

1991).

The Eocene is characterized by deeper depositional facies on the southernmost part of the peninsula

compared to the northern part (Desreumaux, 1985b). In the north, lower Eocene platform limestones

rich in benthic foraminifera progressively change to middle to upper Eocene outer platform limestones

(Calmus, 1983). In the southernmost part the Eocene stratigraphy is dominated by cherty pelagic slope

limestones (Bourgueil et al., 1988). Calc-alkaline basalts interbedded in lower to middle Eocene

limestones are observed in the western part of the Southern Peninsula (Calmus, 1987).

The Oligocene through late Miocene facies are characterized by deep marine chalks with cherts,

pelagic mudstones, chalky limestones and marls (Maurrasse, 1980; Van den Berghe, 1983;

Desreumaux (1987), in Bourgueil et al. 1988). Local highs represented by platform facies are also

observed (Bien-Aime Momplaisir, 1986). A notable exception to these submarine depositional

environments is found west of Camp Perrin in the west-central part of the Southern Peninsula (Fig.

2b). Fining upwards detrital deposits of Burdigalian to Langhian age commence with basal littoral

conglomerates containing Cretaceous basalts and limestones pebbles, which are overlain by lacustrine

85
Chapter 4: Geology of the Southern Peninsula

sands and clays interbedded with lignite levels (Calmus, 1983). The age of these deposits is disputed

by Bourgueil et al. (1988) who argued that the sediments could be reworked and therefore much

younger. The detrital series possibly has a lateral equivalent in the L’Asile region (Fig. 2b; Bien-Aime

Momplaisir, 1986), although in this basin the age is less well constrained. The implication of these

littoral and lagoonal facies on the Miocene evolution of the Southern Peninsula are further addressed

in the discussion.

The late Miocene marks a change in depositional environment. Pelagic limestones and marls are

progressively overlain by flysch-type detrital deposits, whose facies display a shallowing upwards trend

from the late Messinian onwards (Maurrasse, 1982; Bourgueil et al., 1988). The detrital feldspar

fraction in Pliocene deposits consists of labradorite that signals erosion and sourcing from CLIP-type

basalts in the area (Bizon et al., 1985; Desreumaux, (1987), in Bourgueil et al., 1988). Quaternary

continental facies were deposited in basins along-strike and proximal to the trace of the EPGFZ, and in

the Les Cayes, Leogâne and Cul-de-Sac basins (Fig. 2b), while reefal platform limestones were confined

to the coastal areas (Mann et al., 1995).

2.3 Tectonic evolution of the Southern Peninsula during the Cenozoic

The existence and effects of a Late Cretaceous to Paleocene tectonic event on Haiti’s Southern

Peninsula is still subject to debate. Authors in favor such a deformation event (Mercier de Lépinay et

al., 1979; Calmus, 1983; Van den Berghe, 1983a; Bien-Aime Momplaisir, 1986; Calmus and Vila, 1988)

note: (1) the presence of an angular unconformity between the Paleocene and Cretaceous formations,

(2) the existence of deformed Cretaceous limestone fragments in olistoliths encountered in the

Paleocene Rivière Glace formation, (3) outcrop-scale isoclinal folding of the Cretaceous Macaya

limestone with a vergence towards the NE, (4) Cretaceous Macaya limestones thrusted over lower

Paleocene detrital deposits, both of which are unconformably covered by upper Paleocene limestones,

(5) a difference in fold-wavelength and deformation intensity between the Cretaceous units and post-

Paleocene sediments. Other authors (Maurrasse, 1982; Desreumaux, 1985a; Bourgueil et al., 1988;

86
Chapter 4: Geology of the Southern Peninsula

Amilcar, 1997) disagree with this interpretation, and argue that: (1) the transition from Cretaceous to

Paleocene is conformable and (2) the structuration observed results only from Neogene tectonics. The

latter would imply that the difference in fold-wavelength between the Cretaceous and Cenozoic units

is caused by: (i) differences in the mechanical response of the two units to folding, (ii) proximity to

Neogene tectonic structures or (iii) a combination of both.

Angular unconformities of early Miocene age are locally observed in the southwestern (Calmus, 1983;

Amilcar, 1997), central (Bien-Aime Momplaisir, 1986) and eastern (Van den Berghe, 1983a) Southern

Peninsula. The importance, geographical extent, and even existence of these unconformities is

disputed by Desreumaux (1985b) and Bourgueil et al. (1988). A homogenization of pelagic

sedimentation marks a phase of tectonic quiescence during the middle to late Miocene (Bourgueil et

al., 1988). The latest phase of uplift of the Southern Peninsula started during the late Miocene with

erosion reaching the Cretaceous basalts from the late Messinian onwards (Bizon et al., 1985; Bourgueil

et al., 1988).

The most dominant structural feature on the Southern Peninsula is the N080°E trending Tiburon –

Pétionville lineament (Duplan, 1975), known as the Enriquillo-Plantain Garden Fault Zone (EPGFZ) (Fig.

2a; Mann et al., 1984). Present-day activity of this left-lateral fault is documented by: 1) Quaternary

sedimentation within subsiding pull-apart basins along its trace, most notably the Miragoane, Clonard

and smaller unnamed basins in the Tiburon Valley (Mann et al., 1983), 2) left-lateral offset of

Quaternary alluvium (Mann et al., 1995) and stream channels (Prentice et al., 2010), and 3) a

restraining bend north of Camp Perrin affecting Quaternary deposits (Mann et al., 1995). The timing

of onset of strike-slip activity is still debated, with ages ranging from the Eocene (Calmus, 1984) to the

Pliocene (Bourgueil et al., 1988; Mann et al., 1995; Symithe and Calais, 2016).

87
Chapter 4: Geology of the Southern Peninsula

3. Methodology

Biostratigraphic data from literature (Calmus, 1983; Van den Berghe, 1983a; Bien-Aime Momplaisir,

1986; Bourgueil et al., 1988; Amilcar, 1997) was used to create compilation maps that highlight the

Maastrichtian – early Paleocene (Fig. 4a) and early Miocene (Fig. 4b) unconformities. None of the

above literature provides coordinates for sample sites, but rather specifies sample locations relative

to a village, road, river, or cross section. Although care has been taken to re-localize the samples, their

location as presented in Fig. 4a and Fig. 4b should be viewed with caution. Taxa quoted in the

publications was checked against the Mikrotax database and placed in the correct nannofossil (Young

et al., 2018a) and planktonic (Young et al., 2018b) zones. The planktonic taxa were cross-checked,

when possible, against the data provided in BouDagher-Fadel (2015), while the benthic taxa were

checked against the data in BouDagher-Fadel (2008), unless otherwise stated. The extended

compilation of data can be found in Appendix 1 for the Maastrichtian - early Paleocene and Appendix

2 for the early Miocene unconformities. The correlation between the planktonic, benthic, or

nannofossil and the geologic time scale is after Hilgen et al. (2012) for the Neogene, Vandenberghe et

al. (2012) for the Paleogene and Ogg et al. (2012) for the Cretaceous periods.

We re-mapped three parts of the island (Fig. 9, Fig. 11 and Fig. 13) based on field observations

combined with satellite imagery. These maps contain a reference grid with 0.1° spacing. Background

shaded relief is obtained from ASTER DEM. Elevation contour lines with 50m are from the Haitian

Centre National de l’Information Géo-Spatiale (CNIGS). These contours are slightly shifted to match

the DEM and the course of rivers although the fit is not perfect. Offshore bathymetry contours are

from GEBCO, with the first two contour lines at 50m increment and all deeper contours at 250m

increment. The geological maps were created using a workflow (Appendix 3) involving: 1)

implementation of stratigraphic and structural data derived from literature supplemented with our

field observations, 2) georeferenced mapping with MAPublisher® using the DEM and elevation

88
Chapter 4: Geology of the Southern Peninsula

contours, 3) exportation and transparent overlay of the mapped units in Google Earth® for quality

control, and 4) re-iteration of the above steps to arrive at the final result.

Stratigraphic columns used to guide the geological mapping and the cross sections can be found in

Appendix 4, whose locations are displayed on the cross sections of Fig. 10, Fig. 12 and Fig. 14. The

formation name, age, lithology, depositional environment, thickness, and associated references can

be found in Appendix 5.

Kinematic fault slip data was analyzed using the multiple inverse software package from Yamaji et al.

(2005). The orientation of shear fractures (sensu Twiss and Moores, 2007; Fossen, 2010) and the sense

and orientation of kinematic indicators such as slickenlines (i.e. striations, ridges, grooves),

slickenfibres (i.e. calcite steps) and slickolites on these fracture planes are used as input parameters

for the program. The numerical inversion method allows separation of different paleostress states

from heterogeneous fault slip data. The fault dataset and obtained reduces stress tensor were back-

tilted to obtain an Andersonian configuration (Anderson, 1905) for the stress states, with one stress

axis perpendicular and two parallel to the earth’s surface. When present, we used the strike of the

bedding to guide the back-tilting. If back-tilting the stress state using the bedding as a guiding plane

did not result in an Andersonian configuration, the original dataset was preserved. Even though faults

and shear fractures can form in non-Andersonian configurations, for instance resulting from

topographic effects, heterogeneities within the rock, great depth of initiation, or zones of weakness

(Simpson, 1997), an Andersonian configuration offers good first-order constraints for analyzing the

data. Any stress inversion results that, in their present setting, deviate strongly from an Andersonian

configuration are discussed separately.

89
Chapter 4: Geology of the Southern Peninsula

4. Results

4.1 Maastrichtian – early Paleocene unconformity

The compilation map of biostratigraphic data for the Maastrichtian – early Paleocene unconformity

based on the work by Mercier de Lépinay et al. (1979), Calmus (1983), Van den Berghe (1983), Bien-

Aime Momplaisir (1986), Bourgueil et al. (1988) and Amilcar (1997) is presented in Fig. 4a, while an

extensive overview can be found in Appendix 1. The Campanian to Maastrichtian (Ayala, 1959;

Maurrasse, 1982; Calmus, 1983; Van den Berghe, 1983a; Desreumaux, 1985a) Macaya Formation is

eroded in the western Massif de la Hotte, the eastern l’Asile region and in large parts of the Massif de

la Selle (Fig. 4a). In these areas, for instance west of Morne la Visite in the Massif de la Selle (Fig. 2b),

the Cretaceous CLIP basalts are unconformably overlain by lower Paleocene conglomerates and

clastics containing erosional products derived from the Cretaceous basalts and Macaya Formation (Fig.

5a; Calmus, 1983; Bourgueil et al., 1988). The sequence generally displays a deepening and fining

upwards trend and becomes more calcareous upwards in the sequence (Calmus, 1983; Van den

Berghe, 1983a; Bien-Aime Momplaisir, 1986; Amilcar, 1997). In the eastern Massif de la Hotte and

western Massif de la Selle (Fig. 4a) the Macaya Formation is not completely eroded and the contact

with the lower Paleocene is marked by a ravinement surface (Desreumaux, 1985a; Bourgueil et al.,

1988) or a minor erosional unconformity (Calmus, 1983; Van den Berghe, 1983a). Even though the

lower Paleocene is directly overlying the Macaya Formation in these areas, the uppermost

Maastrichtian is never observed in-situ and only found as erosional products (Bourgueil et al., 1988).

Because of the lithological homogeneity (Bourgueil et al., 1988) and similarity in facies (Calmus, 1983)

of the Macaya Formation throughout the Southern Peninsula, it is unlikely that the absence of this

formation in certain areas can be ascribed to local non-deposition. It is more likely that the Macaya

Formation was deposited throughout the Southern Peninsula during the Campanian to Maastrichtian

and locally became uplifted and eroded.

90
Chapter 4: Geology of the Southern Peninsula

-74° -73° -72°


A Jeremie Cretaceous - Paleocene unconformity map

Anse-d’Hainault PORT-AU-PRINCE
Baradères
18.5°

18.5°
Rampe des Lions
Grand Goâve
MASSIF DE LA HOTTE L’Asile
Bonne Fin Grand Caille Beloc
MASSIF DE LA SELLE
Tiburon Camp Perrin Marbial
Port-a-Piment
Erosion;
Morne Orangers Jacmel Marigot strong
Les Cayes Gross Caye Bainet
Chantal mild

q SOUTHERN PENINSULA conformable


Port Salut
Île-a-Vache
18°

18°
0 20 40 80
Kilometers
-74° -73° -72°
-74° -73° -72°
B Jeremie Miocene unconformity map
Roseaux
PORT-AU-PRINCE
18.5°

18.5°
MASSIF DE LA HOTTE Camp Perrin Beloc
Tiburon MASSIF DE LA SELLE
L’Asile
Maniche
Platon Besace
Port-a-Piment Touya Jacmel Marigot
Les Cayes Erosion;

q
strong
SOUTHERN PENINSULA medium
mild
18°

18°
0 20 40 80 conformable
Kilometers
-74° -73° -72°

Fig. 4a: Compilation map of the Cretaceous – Paleocene unconformity based on biostratigraphic ages for
Paleocene deposits and underlying formations. Color coding indicates erosion. Red; strong, erosion down into
Cretaceous basalts. Yellow; mild, erosion down into Cretaceous Macaya Formation limestones. Green;
conformable, biostratigraphically complete and concordent transition from Cretaceous to Paleogene. Fig. 4b:
Compilation map of the Cretaceous – Paleocene unconformity based on biostratigraphic ages for lower to middle
Miocene deposits and underlying formations. Color coding indicates erosion. Red; strong, erosion down to
Cretaceous. Orange; medium, erosion down to Paleocene or Eocene. Yellow; mild, erosion down to Oligocene or
intra-Miocene. Green; conformable, biostratigraphically complete and concordent transition from Paleogene to
Neogene.

91
Chapter 4: Geology of the Southern Peninsula

The areas unaffected by this uplift phase experienced continued sedimentation of deep marine chalks

and limestones during early Miocene times (Van den Berghe, 1983a; Bizon et al., 1985; Desreumaux,

1987). The regions affected by the uplift experienced a rapid return to deep marine sedimentation

during the middle Miocene. This is characterized by a homogenization of deep marine facies without

detrital erosional products that are found throughout the Southern Peninsula (Calmus, 1983; Van den

Berghe, 1983a; Bizon et al., 1985; Bien-Aime Momplaisir, 1986; Desreumaux, 1987; Bourgueil et al.,

1988; Amilcar, 1997). The fact that these marine deposits are presently emerged requires another,

post-middle Miocene, phase of uplift and erosion.

4.3 Structural style of deformation

4.3.1 Western Massif de la Hotte

The Cretaceous basalts and limestones are strongly fractured and folded. Northwest of Les Anglais (Fig.

2b) a 500 x 200 x 100m block of silicified limestones and radiolarites forms a wedge within the

Cretaceous basalts. Limestone intercalations within these basalts are boudinaged (Fig. 6a). Halfway

Tiburon and Les Anglais (Fig. 2b), blocks of Macaya limestone are embedded as tectonic lenses within

the basalts, with intra-basalt faults thrusting folded silicified limestones and radiolarites over

Cretaceous basalts with top-to-N motion. In the Tiburon Valley (Fig. 2b), Macaya Formation limestones

developed tight disharmonic folds along WNW-trending axes (Fig. 6b).

Eocene platform limestones in this valley are gently folded along W-trending axes (Fig. 6c). The Eocene

limestones north of Rampe des Lions (Fig. 2b) are also characterized by gentle to open folds along

WNW-trending axes. The Miocene to Quaternary sediments between Port-a-Piment and Tiburon (Fig.

2b) are tilted 35° to the west and are transected by N – S trending normal faults (Fig. 6d).

93
Chapter 4: Geology of the Southern Peninsula

The core of the EPGFZ south of Port-a-Prince (Fig. 2b) consists of a more than 200m wide zone of re-

cemented limestone breccia (Fig. 8b). This tectonically pulverized rock is used as building material and

mined in quarries that are located along the EPGFZ. At the quarries of La Boule and Toto (Fig. 2b),

predominantly south-dipping and east – west trending polished fault mirrors range from a few meters

to 100m in size. Some of the fault surfaces are undulating along sub-vertical and sub-horizontal axes

(Fig. 8b). The sense of motion on oblique faults that are in a Riedel configuration to the E – W trending

fault mirrors is mainly sinistral on sub-horizontal kinematic indicators and reverse on sub-vertical ones.

The sub-horizontal trends are overprinted by the sub-vertical trends.

At Boutillier quarry (Fig. 2b) large fault mirrors with undulating surfaces are dipping steeply to the

north. Kinematic indicators show reverse motion with a north-side-up component.

At Les Cayettes quarry (Fig. 2b) a 45° east-dipping thrust fault separates a lower to middle Miocene

hanging wall from an upper Miocene – lower Pliocene footwall (Fig. 8c). East-west trending sub-

vertical faults with inclined striations are only observed in the hanging wall and not in the footwall.

These striations dip 45° to the east, similar to the orientation and dip of the thrust. This indicates that

the vertical strike-slip faults are older and displaced by the thrust.

96
Chapter 4: Geology of the Southern Peninsula

4.4 Map-scale deformation structures

4.4.1 Western Massif de la Hotte

The largest first-order fold is an E – W trending 10 km half-wavelength anticline that continues from

Les Irois in the west through the Macaya Massif in the east (Fig. 9). This fold is tighter in the east (Fig.

10b) than in the west (Fig. 10a). In the map area where the sedimentary cover is lacking (Fig. 10), it is

difficult to observe any structuration within the Cretaceous basalts on satellite imagery. The second

order folding shown on cross section A (Fig. 10a), with wavelengths of around 5 kilometers, is

interpretative and based on the topography, folding observed on the onshore cross section B (Fig. 10b)

and on offshore seismic profiles (for instance line CT2-28 from Mann et al. (1995)).

The E – W trending topographic lineament that is expressed in the morphology east of Tiburon is

considered to be the trace of the EPGFZ (Fig. 9; Mann et al., 1984). This lineament consists of at least

five segments in the mapped area. Close to Tiburon the valley widens and the morphological

expression of the EPGFZ becomes less pronounced. In map view we cannot identify any clear lateral

offset of lithological units. Based on the difference in elevation of the Macaya limestones on both sides

of the Tiburon Valley (Fig. 10a) a vertical offset is inferred and estimated at around 1000m. Another

east – west trending topographic lineament, although less prominently expressed in the morphology

compared to the EPGFZ, originates at Les Irois, continues east for around 40 kilometers, and is

displacing the course of the Rivière Bras à Droite and Rivière Bois Mahot (Fig. 9). Further east this trace

undulates and it is unclear if a single fault is still present north of the Pic Macaya. These lineaments are

interpreted as the surface traces of strike-slip faults. Based on field observations, the deformation

accommodated on each of them appears however scarce.

98
Chapter 4: Geology of the Southern Peninsula

Between Anse-d’Hainault and Source Chaudes an E – W trending, south-verging thrust is inferred

based on the mapping by Calmus (1983) and Amilcar (1997) and the measured dip of the lower

Paleocene to the south of the fault (Fig. 9). This fault, if mapped correctly, runs through the area of

hot springs at Sources Chaudes, displaces lower Paleocene clastics and Cretaceous basalts (Calmus,

1983), but based on satellite images does not displace the upper Paleocene or younger lithologies. This

indicates that this thrust fault is of pre- late Paleocene age (Fig. 10).

4.4.2 L’Asile region

The EPGFZ in the mapped area (Fig. 11) consists of a continuous northern segment and a shorter

southern segment, both of which are acting as boundaries to the Clonard Basin. The EPGFZ offsets

mapped lithological units north and south of its trace. A change in vertical motion occurs along the

trace of the EPGFZ in this area, with the l’Asile Basin (section C, Fig. 12c) displaying south-side up and

the Clonard Basin (section A, Fig. 12a) displaying north-side up motion. The Clonard Basin is trending

E – W and is bounded by faults to the north and south (Mann et al., 1983, 1995). Neither on satellite

data nor in the field could we identify a fault acting as the northern boundary to the WNW-trending

l’Asile Basin, which appears only bounded to the south by the EPGFZ (Fig. 11).

Several second-order WNW – ESE trending synclines and anticlines with wavelengths in the order of 5

km were observed on satellite images in the region north of the EPGFZ (Fig. 11 and Fig. 12). The

wavelengths of these structures are comparable to the wavelengths of fault-controlled folds observed

on seismic data offshore between the Southern Peninsula and Gonâve Island (Bien-Aime Momplaisir,

1986; Mann et al., 1995). We have no field or satellite image based evidence to support that the

onshore folds are also fault-controlled.

101
Chapter 4: Geology of the Southern Peninsula

4.4.3 Massif de la Selle

The two dominant structures on the map are the N080°E trace of the EPGFZ and the roughly E – W

trending axis of the Massif de la Selle anticline (Fig. 13). This anticline has a wavelength of around 40

km and is mildly asymmetric, with a gentle southern limb and a slightly steeper northern limb. Second-

order folding with a 5 km wavelength is more open compared to the other cross sections (Fig. 13 and

Fig. 14). Within the Cretaceous basalts in the central part of the anticline we were unable to identify

any folding based on satellite images or topography.

West of Obleon, on the northern flank of the anticline, the Paleocene is missing, with a fault marking

the contact between the Eocene and the Cretaceous (Fig. 13 and Fig. 14). The sense of motion on this

fault could not be determined. An upper Maastrichtian to lower Paleocene series is identified in the

Beloc region on the northern flank of the anticline (Fig. 2b; ‘Jacmel Blocks’ of Mercier de Lépinay et

al., 1979). This indicates that the Paleocene is deposited on the northern flank and is the reason for

interpreting it in Fig. 13.

The labelled faults bordering the Cul-de-Sac plain (Fig. 13) are taken from the work of Saint Fleur et

al. (2015) and verified using satellite imagery. According to these authors and to Symithe and Calais

(2016), the folds in the southern Cul-de-Sac plain are fault-propagation folds bounded by south-dipping

thrusts. The previously unnamed fault between Fermate and Kenscoff (Van den Berghe, 1983a) is

labelled Redoute after the corresponding river valley, while the Chauou fault is named after the town

on the crest of the anticline directly south (Fig. 13). We interpret the Redoute and Chauou faults as

south-dipping thrusts, since they juxtapose an anticline of older sequences to the south against

younger sequences to the north (Fig. 13 and Fig. 14). In map view the thrust faults terminate against

the EPGFZ. This indicates that these thrusts are either; 1) synchronous or post-dating the EPGFZ and

are abutting against it (scenario 1, Fig. 14a), or 2) that they are older and subsequently offset by the

EPGFZ (scenario 2, Fig. 14b).

104
Chapter 4: Geology of the Southern Peninsula

4.5 Paleo-stresses and relative timing

4.5.1 Western Massif de la Hotte

Within the Macaya Formation limestones south of Tiburon (Fig. 9.1), set 1A shear fractures were

formed by a strike-slip regime with NNW - SSE trending σ1 axis. Set 1B resulted from a reverse regime

with north – south trending σ1 axis. In the present-day orientation set 1A is in an Andersonian

configuration. Back-tilting the bedding to a horizontal position restores the set 1B strike-slip stress

regime to an Andersonian configuration. This suggests that set 1B originated prior to tilting of the

bedding and formed before set 1A shear fractures, and that reverse faulting preceded strike-slip.

Lower Eocene platform limestones in the Tiburon valley, proximal to the trace of the EPGFZ, have

pervasive and regularly spaced orthogonal joints that are reactivated as shear fractures (Fig. 9.2).

Reactivation resulted from a strike-slip regime with NNW – SSE trending σ1 axis, which is similar to set

1A, possibly indicating that both were created post- early Eocene. Middle Miocene chalky limestones

southeast of Tiburon at Haut Fort yielded one set of shear fractures that resulted from a reverse regime

with NNE – SSW trending σ1 axis (Fig. 9.3). Back-tilting over an E – W trending axis restores the stress

regime to an Andersonian configuration, which indicates that the bedding was mildly tilted to the west

before being affected by the reverse stress regime.

Two sets of normal faults are observed. One set is located in the middle Miocene (Langhian?) chalky

limestones northwest of Port-a-Piment (Fig. 9.4), while another set is situated in the Plio-Quaternary

conglomerates south of Tiburon (Fig. 9.5). Both sets indicate E – W oriented extension.

The above results can be summarized as follows: (1) reverse regime with N – S σ1 axis between post-

Maastrichtian and probably pre-Eocene times, followed by (2) strike-slip regime with NNW – SSE σ1

axis in post- early Eocene times, (3) reverse regime with NNE – SSW σ1 axis during post- early Miocene

times, and (4) post- Langhian through Plio-Quaternary extensional regime with E – W σ3 axis. No

relative timing could be established between the last three deformation phases.

107
Chapter 4: Geology of the Southern Peninsula

4.5.2 L’Asile region

Conjugate shear fractures with variably plunging striations were found in the steeply dipping

calcareous beds of the lower Paleocene volcaniclastic sequence along the Rivière Dose (Fig. 11.1).

Back-tilting this dataset resulted in an Andersonian extensional stress regime with E – W trending σ3

axis.

Paleocene clastic limestones between Mornes and Bontemos (Fig. 11.2) contain shear fractures with

sub-horizontal striations. Stress inversion on these fractures in their present-day orientation results in

an Andersonian strike-slip regime with east – west trending σ1 axis.

NW – SE trending normal faults were found in lower Eocene limestones west of Cavaillon (Fig. 11.3).

Stress inversion and subsequent back-tilting of the data results in a NE – SW trending σ3 axis.

No relative timing could be established between the different data sets, with the results summarized

as follows: (i) post- early Paleocene extensional regime with E – W trending σ3 axis, (ii) post- Eocene

extensional regime with NE – SW trending σ3 axis, and (iii) post- Paleocene strike-slip regime with E –

W trending σ1 axis. Although there is no relative timing of deformation, the strike-slip regime (Fig.

11.2) is probably a young deformation event, since it took place when the bedding was already roughly

in its present-day orientation.

4.5.3 Massif de la Selle

Lower Eocene limestones east of Ristache have NE – SW trending left-lateral shear fractures resulting

from a back-tilted strike-slip regime with NNW – SSE trending σ1 axis (Fig. 13.1). Lower Paleocene

sediments west of Morne La Visite are fractured by a back-tilted strike-slip regime with NNW – SSE

trending σ1 axis (Fig. 13.2). Both these stress regimes are similar, suggesting that they initiated post-

early Eocene.

108
Chapter 4: Geology of the Southern Peninsula

On the northern flank, in both the Cretaceous basalts and Eocene limestones directly adjacent to the

faulted contact west of Obleon, left-lateral shear fractures resulted from a strike-slip regime with NNE

– SSW trending σ1 axis (Fig. 13.3). At this locality it is unclear if the strike-slip regime pre- or post-dates

the faulted contact. A strike-slip regime with NE – SW trending σ1 axis fractured Eocene limestones

directly west (Fig. 13.4). It is not possible to back-tilt this dataset to an Andersonian configuration

using only the bedding, possibly indicating a local variation in the stress regime.

In the core of the EPGFZ at La Boule quarry we identified distinct sets of shear fractures, whose

inversion points to two distinct paleo stress states (Fig. 13.5). Because the measurements were taken

within the core of the fault the obtained stresses may reflect the local rather than the far field stress

state. Set 5A and 5B have an E – W and NE – SW trending σ1 axis, respectively. Reverse striations that

dominate set 5A overprint strike-slip striations of set 5B, indicating that reverse faulting is younger

than strike-slip at this location. The NE – SW orientation of the σ1 axis at Toto Quarry (Fig. 13.6) is

similar to set 5B at La Boule, and both measurement sites are located in the core of the EPGFZ. The

difference is that stress state 6 at Toto is reverse, while 5B at La Boule is strike-slip. At Boutillier quarry

thrusting resulted from a reverse regime with N – S trending σ1 axis (Fig. 13.7). At Les Cayettes quarry,

striations on strike-slip faults, which are only observed in the lower to middle Miocene hanging wall

(Fig. 8c), have a similar plunge as the dip of the thrust fault. This suggests that the strike-slip faults are

older than the thrust and were subsequently transported onto the upper Miocene – lower Pliocene

footwall. This indicates that strike-slip activity occurred roughly pre- latest Miocene, with thrusting

roughly post- earliest Pliocene.

The above results can be summarized as follows for the southern flank: (i) Post- early Eocene strike-

slip regime with NNW – SSE trending σ1 axis. For the northern flank: Reverse stress regimes with

variably oriented σ1 axes are generally post-dating strike-slip stress regimes with NE – SW and NNE –

SSW trending σ1 axes. At Boutillier quarry the timing is constrained to (1) Pre- late Miocene strike-slip

and (2) post- earliest Pliocene thrusting.

109
Chapter 4: Geology of the Southern Peninsula

4.6 Compilation of kinematic data

Stress inversion of kinematic fault slip data obtained at 22 sites resulted in a total of 26 reduced stress

tensors. The spatial distribution of paleostresses along the EPGFZ are plotted on the map in Fig. 15a.

A -74° -73° -72° Map symbols;


= Sigma1 axis
* = Sigma 1 VDB83
PORT-AU-PRINCE
= Sigma3 axis
18.5°

18.5°
Léogâne (normal)
* = Outcrop fold axis
MASSIF DE LA HOTTE *
EPGFZ
MASSIF DE LA SELLE = Thrust
= Anticline
= Syncline

q
= EPGFZ
SOUTHERN PENINSULA = focal mechanism
for the 2010
earthquake
18°

18°
0 20 40 80
Kilometers = focal mechanism
-74° -73° -72°
of 2010 aftershocks

B [deg]
EPGFZ
100 between σ1 or σ3 and the
70 to 100
80 EPGFZ-trace azimuth
60
40

σ1/σ
+100°
20

3
GFZ
0 trace EP -20°
50 100 150 200 250 [km] σ1/σ3
-20
Age of lithologies; Plio.-Quat. Miocene Oligocene Eocene Paleocene Camp-Maast Stress regime; Normal Strike-slip Reverse Core of EPGFZ

Fig. 15a: Plot of trends of sigma 1 (black) and sigma 3 (blue) axes from stress inversion results. Fold axes derived
from outcrops are plotted in green. Fig. 15b: Chart showing angular difference (degrees) between trend of the
sigma 1 or sigma 3 axes and the trace of the EPGFZ. Colors indicate age of lithologies from which the data was
obtained. Square indicates normal stress regime, circle indicates strike-slip stress regime and hashed circle
indicates reverse stress regime. Results on the right-hand side are constructed using the N095°E average trend
of thrusts in this region. Red circle is data from the core of the EPGFZ. Plotted results are at less than 15km
distance from the EPGFZ trace. Angular difference varies between -20 and +100 degrees (see inset on right).

4.6.1 Along-strike variation of stress tensors

The maximum principal stress axes can be separated into three groups (Fig. 15b). The first group has

an angle with the EPGFZ varying between 70 and 100°. The second group has an angle between -20°

and 20° to the EPGFZ. The third group has an angle of approximately 45° to the EPGFZ. An angle of 45°

or less is obtained from data collected in the brecciated fault core of the EPGFZ. West of Léogâne a

reverse regime is only found in Cretaceous and lower Paleocene formations. East of Léogâne reverse

regimes are found in lower to middle Miocene limestones.

The minimum principal stress axes for normal regimes are either at around 45° to strike of the EPGFZ

or grouped between -20° to 10°, hence sub-parallel to the EPGFZ.

110
Chapter 4: Geology of the Southern Peninsula

5. Discussion

5.1 Maastrichtian – early Paleocene deformation phase; local folding, thrusting, and erosion

The existence, extent, and structural implications of a Late Cretaceous to Paleocene tectonic event on

the Southern Peninsula are subject to debate, with authors arguing in favor (Mercier de Lépinay et al.,

1979; Calmus, 1983; Van den Berghe, 1983a; Bien-Aime Momplaisir, 1986; Calmus and Vila, 1988) or

against (Maurrasse, 1982; Desreumaux, 1985a; Bourgueil et al., 1988; Amilcar, 1997) such a

deformation phase.

The first two arguments of the authors in favor of this deformation phase focus on the stratigraphic

unconformity between the Cretaceous and Paleogene formations and the presence of Cretaceous

erosional products within the lower Paleocene formations. The compilation map of Cretaceous and

Paleocene biostratigraphic data (Fig. 4a) clearly shows the presence of a Maastrichtian to early

Paleocene unconformity. Erosion was strongest in the Anse-d’Hainault area, along an E – W line from

Port-a-Piment to Morne Orangers, and in the central Massif de la Selle region. The area around Platon

Besace is not affected by this deformation, while north of Camp Perrin and south of Baraderes erosion

only removed part of the Cretaceous sedimentary cover. The upper Maastrichtian to Danian clastic

sequence overlying the unconformity often starts with a basal conglomerate containing clasts of

Cretaceous basalts and Macaya Formation limestones (Mercier de Lépinay et al., 1979; Calmus, 1983;

Van den Berghe, 1983a; Bien-Aime Momplaisir, 1986; Calmus and Vila, 1988).

The other arguments focus on a difference in deformation style and intensity between Cretaceous to

lower Paleocene formations on the one hand, and the upper Paleocene and more recent formations

on the other. The tightness of folds may also depend on the mechanical behavior of the lithology to

compression, which is influenced by layer thickness and the type of lithology. There are however no

first-order lithological differences between the Cretaceous Macaya Formation and the middle Eocene

to early Miocene formations. Both consist of dm-bedded chalky limestones with cherts. The early

112
Chapter 4: Geology of the Southern Peninsula

Eocene limestones have a similar bed thickness, but consist of platform facies without cherts (Calmus,

1983; Van den Berghe, 1983a; Desreumaux, 1985a; Bourgueil et al., 1988).

The Cretaceous limestones in the Tiburon Valley are tightly folded around WNW-trending axes, while

lower Eocene limestones in this valley are only mildly folded and deformed around ENE-trending axes

(Fig. 6b and Fig. 6c). Any deformation resulting from activity on the EPGFZ, which runs through this

valley as well, can be expected to have a similar effect on both lithologies. The difference in style and

intensity of deformation is thus not related to activity on the EPGFZ. It is probable that this difference

in style results from a deformation event occurring prior to deposition of the lower Eocene limestones,

and post-dating deposition of the Macaya Formation limestones. The most likely candidate is the

Cretaceous to early Paleocene deformation that is also responsible for the associated unconformity.

Calmus (1983) and Calmus and Vila (1988) also described a difference in the style and intensity of

deformation between Cretaceous to lower Paleocene, and upper Paleocene to younger formations.

These authors observed a NE-vergence for structures within the Cretaceous and early Paleocene

formations. Our own field observations contradict this, with most structures displaying a vergence

towards the south or southwest (Fig. 9a and Fig. 9b), with a minority showing top-to-north or

northeast motion (Fig. 8c).

Calmus (1983) and Calmus and Vila (1988) also proposed that the Macaya Formation was transported

as a thin-skinned thrust sheet over the Cretaceous basalts towards the NE, over distances up to 30 km.

We observed that in large parts of the Southern Peninsula the Macaya Formation is not eroded, only

mildly deformed and in a stratigraphic contact with the underlying Cretaceous basalts. We found no

evidence supporting the transport of the Macaya Formation as a thrust sheet over the Cretaceous

basalts and the lower Paleocene units. Deformation and uplift probably resulted from a crustal-scale

buckling-style of folding (Fig. 10 and Fig. 12), which can explain the lateral differences in the amount

of erosion of the Cretaceous formations.

113
Chapter 4: Geology of the Southern Peninsula

5.2 Early Miocene deformation phase; local folding, minor thrusting, and erosion

The existence of this deformation phase strongly hinges on the existence of early Miocene erosional

unconformities. Our compilation map of biostratigraphic data (Fig. 4b) shows an E – W trend in the

amount of erosion of Aquitanian and older units, which is decreasing eastwards. Around 1000m of

erosion occurred in the southwestern Massif de la Hotte, while sedimentation was continuous

throughout most of the Massif de la Selle. This compilation also shows a general correlation between

the amount of erosion and the onset of post-erosion sedimentation; the stronger the erosion, the

younger the age of the overlying units. The timing of deformation responsible for this erosion is

probably Aquitanian in age, as indicated by the intra-Aquitanian unconformity at Platon Besace (Fig.

4b). The paleo-topography created by the uplift in the western Southern Peninsula only became

submerged during middle Miocene times.

Calmus (1983) and Bizon et al. (1985) have related this early Miocene phase of uplift to strike-slip

activity along the EPGFZ, because they noted that some of the unconformities occur proximal to the

trace of the EPGFZ. Although the unconformities trend along-strike of the EPGFZ (Fig. 4b), we doubt

that strike-slip activity on the EPGFZ caused early Miocene uplift and erosion. Firstly, erosion gradually

decreases eastwards and the effects of erosion are known to affect the area up to at least 15 km

distance to the trace of the EPGFZ, for instance at Port-a-Piment (Fig. 4b). This contrasts with offshore

observations along the active trace of the EPGFZ between Haiti and Jamaica (Leroy et al., 2015;

Corbeau et al., 2016a; b), where transpressional and transtensional structures are irregularly

distributed along, but always within a few kilometers distance to, the trace of the EPGFZ. The observed

pattern for the early Miocene unconformity onshore (Fig. 4b) better matches with large-wavelength

folding as shown on the cross sections (Fig. 10 and Fig. 12). Secondly, the middle to early late Miocene

is characterized by homogeneous marine sedimentation without evidence for tectonic activity. Early

Miocene activity on the EPGFZ should thus have ceased during the middle Miocene, only to become

active again in the late Miocene. There is no evidence to support such behavior for the EPGFZ.

114
Chapter 4: Geology of the Southern Peninsula

5.3 Late Miocene – present-day; strike-slip to transpressive deformation and island-wide uplift

The most recent phase of uplift of the Southern Peninsula is recorded by a change in sedimentary

facies from pelagic marls and limestones that dominated the Miocene to detrital sedimentation of

marls, clays and calcareous sandstones from the latest Messinian onwards (Bourgueil et al., 1988).

Labradorite found as the detrital fraction records erosion from the Cretaceous basalts (Desreumaux,

(1987) in, Bourgueil et al., 1988). This indicates that during the latest Messinian, exhumation already

reached the structural levels of the Cretaceous basalts in the core of the Massif de la Selle.

Present-day compression in the Southern Peninsula is shown by GPS velocity modeling (Benford et al.,

2012a; Calais et al., 2016). Using the modelled GPS velocity vectors from Calais et al. (2016) and the

trend of the EPGFZ, we can derive the fault parallel and fault perpendicular GPS velocities (Fig. 2a). In

the western Massif de la Hotte shortening is partitioned via 8.8 mm yr-1 of EPGFZ-parallel and 4.8 mm

yr-1 of fault-perpendicular motion. In the l’Asile region shortening is partitioned via 8.7 mm yr-1 EPGFZ-

parallel and 6.0 mm yr-1 perpendicular motion, while in the eastern Port-au-Prince region shortening

is partitioned via 8.6 mm yr-1 EPGFZ-parallel and 7.7 mm yr-1 EPGFZ-perpendicular motion (Calais et al.,

2016). This indicates that the EPGFZ parallel velocity is relatively constant over the Southern Peninsula,

but that EPGFZ perpendicular velocities increases from west to east by around 40%.

Present-day shortening between the Massif de la Selle and the Cul-de-Sac plain is partitioned over the

EPGFZ and the thrust faults bordering the Cul-de-Sac plain in the south. The EPGFZ is presently active

(Mann et al., 1983, 1995; Prentice et al., 2010) and rooted in the upper mantle (Benford et al., 2012c).

Activity on the thrust faults is documented by active folding of Quaternary alluvium along the Lamentin

thrust (Saint-Fleur et al., 2015) and reverse faulting in Quaternary alluvium at the Dumay thrust along

the Rivière Grise (Terrier et al., 2014; Saint-Fleur et al., 2015). Based on our geological map (Fig. 13)

and cross section (Fig. 14) we have two possible scenarios for interpreting the geometry of these faults

at depth (Fig. 17).

115
Chapter 4: Geology of the Southern Peninsula

Fig. 17e) either abut against the EPGFZ or are left-stepping and geographically proximal to the EPGFZ.

We note that this scenario (#1) is essentially the same as what is modelled by Symithe and Calais

(2016). The main difference is that in our conceptual model the EPGFZ is bending as it approaches the

Cul-de-Sac plain and remains active as an oblique fault at depth. Another possible scenario is that the

thrust faults are older and offset by the EPGFZ (#2; Fig. 17b and Fig. 17d). This scenario (#2) is in

conflict with our observations showing that vertical striations with a reverse motion on fault planes

are generally younger and cross-cutting horizontal lineations. This scenario (#2) does also not explain

that the thrust faults are only found in the vicinity of the EPGFZ. We strongly favor scenario #1 where

the EPGFZ is the dominant fault in the region with the thrust faults rooted on, and thus associated

with, the EPGFZ.

5.4 Timing of paleostress states and along-strike variation

The reverse stress regimes found in Cretaceous and early Paleocene formations on the western part

of the Southern Peninsula indicate north-south compression. At Tiburon the reverse regime is older

than the strike-slip regime (conform section 4.5.1 and 5.1). Although the orientation of σ1 for the

reverse and strike-slip regimes in this area is similar (Fig. 9.1), it is probable that these reverse regimes

record the Maastrichtian – Paleocene deformation phase, since no reverse regime is observed in

younger lithologies in this region. Thrust faults associated with this deformation phase trend E – W to

NW – SE, indicating a general NNE – SSW directed compression (Fig. 18).

South of Maniche, Eocene to Oligocene limestones are folded along N130°E trending axes and are

unconformably overlain by middle Miocene sediments (Calmus, 1983). These trends indicate a local

NE – SW oriented shortening prior to middle Miocene times. Fold axes with a similar trend are found

in group C (Fig. 16b). This trend is only observed in lithologies that are Eocene or older, and

predominantly in the western two-thirds of the Southern Peninsula. We therefore relate these folds

to the early Miocene deformation phase (Fig. 18).

117
Chapter 4: Geology of the Southern Peninsula

stresses that have a significantly high angle to the trace of the EPGFZ are related to a strike-slip stress

regime (Fig. 15b). Present-day stresses, as recorded by the N021°E trend of the P-axis of the 2010

Léogâne earthquake, are at a 64° angle to the trace of the EPGFZ. This event did not involve slip on the

EPGFZ (Prentice et al., 2010) however, but resulted from oblique slip on the NNW dipping Leogâne

fault (Calais et al., 2010; Mercier de Lépinay et al., 2011; Douilly et al., 2013). Similar high angles of

maximum principal stress are observed at the San Andreas Fault (SAF) system in California. There, the

axis of maximum principal stress obtained from focal mechanisms, stress-induced wellbore breakouts

and field measurements is also sub-perpendicular to the trace of the SAF, with fold axes and thrust

faults trending sub-parallel to the trace (Zoback et al., 1987). These results are comparable with the

data from the Cajon Pass (Zoback and Healy, 1992) and SAFOD (Hickman and Zoback, 2004) boreholes

along the SAF, which in addition show considerable variation of the angular difference with depth. The

high angle of σ1 to the trend of the SAF is interpreted as the result of stress rotations around a weak

fault zone (Zoback et al., 1987). In similar fashion, the 80° angular difference between maximum

horizontal stress and the trace of the Median Tectonic Line and the Rokko-Awaji segment in Japan is

also interpreted to result from stress rotations proximal to a weak fault zone (Famin et al., 2014). The

sub-perpendicular trend of σ1 axes to the trace of the EPGFZ could also indicate that this strike-slip

fault is mechanically weak.

The east – west trending topographic lineament originating at Les Irois in the Massif de la Hotte (Fig.

9) is morphologically less expressed than the EPGFZ, and possibly represents an inactive strike-slip fault

segment. An east – west trending strike-slip fault zone without a morphological expression is observed

on the plateau southwest of Miragoane (Fig. 2b; Bourgueil et al., 1988). Locally short-lived strike-slip

activity during the late Miocene is documented at Les Cayettes quarry (section 4.5.3). Present-day

strike-slip activity is mainly focused on the active segments of the EPGFZ (Mann et al., 1983, 1995;

Prentice et al., 2010). During the development of a fault system, displacement may progressively

become focused onto newly linked fault segments, with activity terminating on smaller faults (Walsh

et al., 2002). Such a process possibly played a role in the development of the EPGFZ. In this scenario

119
Chapter 4: Geology of the Southern Peninsula

initially distributed strike-slip activity progressively became focused through time along the

morphologically well express but strongly segmented EPGFZ, with activity terminating on older fault

segments at a distance from the EPGFZ.

From the above discussion it follows that; 1) strike-slip activity became established during the late

Miocene (Fig. 18) and was characterized by spatially distributed deformation, which 2) became

progressively more focused along the EPGFZ from the latest Miocene or Pliocene onwards, possibly

concomitant with EPGFZ-parallel shortening and compression, while 3) N – S directed thrusting, which

locally post-dates strike-slip activity, became active from the Pliocene onwards.

5.5 Evolutionary cross section through the central Southern Peninsula

The deformation phases described above are summarized in a series of schematic cross sections,

illustrating the retro-deformation of the north-south cross section D of Fig. 12. The inferred amount of

shortening shown in Fig. 19 is constructed using the line-length of the uppermost formation in the

proceeding image. This gives a conservative, minimum estimate of shortening between the Cretaceous

and Present of ~11%, or ~2.3 km on this cross section. The total amount of shortening is probably

larger due to overthrusting or layer-parallel shortening hidden by subsequent erosion.

 The top cartoon (Fig. 19a) shows the configuration at the end of the Cretaceous prior to the

first deformation phase. Following emplacement of the Cretaceous flood basalts, pelagic

sediments of the Macaya Formation were deposited. Intercalated sediments within the

Cretaceous basalts and the Macaya Formation limestones indicate a general deepening trend

from the Albian to Maastrichtian (Maurrasse et al., 1979a; Bien-Aime Momplaisir, 1986;

Bourgueil et al., 1988; Mann et al., 1991b).

 The first deformation phase (Fig. 19b) occurred during the Maastrichtian and early Paleocene,

when the Southern Peninsula was locally subjected to folding with S- to SW-vergence. Based

on the extent of the erosional unconformity (Fig. 4a and Fig. 11) the first-order wavelength is

estimated at around 20 km. This folding was probably aided by thrusts, similar to the one at

120
Chapter 4: Geology of the Southern Peninsula

Anse-d’Hainault (Fig. 9 and Fig. 10). Deformation in the l’Asile region (Fig. 11) resulted in

erosion of the Cretaceous formations, which was more severe in the eastern part of this area.

Maastrichtian erosion is also documented offshore southern Haiti on the northern Beata Ridge

(Mauffret et al., 2001).

 Subsidence and a deepening of facies characterized the late Paleocene to Aquitanian period

(Fig. 19c). The global rise in sea level during this period of around 100m (Haq et al., 1988) is

insufficient to account for the relative sea level rise and thickness of Paleogene sediments,

which is at least 1000m (Bourgueil et al., 1988). Regional controls on the subsidence could

involve: 1) a flexural response to deformation in the north, 2) a response to rifting and

subsequent spreading in the Cayman Trough, or 3) cooling and contraction of the lithosphere

following the Late Cretaceous flood basalt event, as proposed for other LIPs (Greene et al.,

2010).

 Renewed folding and uplift during early Miocene times was more pronounced in the central

region of the section (Fig. 19d). Erosion locally removed 500 to 1000m of Paleogene

sediments. Tectonic inversion was possibly aided by reactivation of older faults. The cause for

this deformation event remains enigmatic. The Oligocene to early Miocene corresponds to a

number of major tectonic events in the region: 1) at a large scale, convergence between North

and South America accelerated during the Oligocene resulting in increased compressional

deformation along the Caribbean Plate boundaries (Somoza, 2007), 2) the Cayman spreading

center grew south between 26 and 20 Ma, which initiated activity on the Walton Fault (Leroy

et al., 2000), 3) Hispaniola started to separate from Cuba at around 20 Ma (Pindell and Barrett,

121
Chapter 4: Geology of the Southern Peninsula

1990; Calais and Mercier de Lépinay, 1995), contemporaneous with forward propagating

folding and thrusting in central Haiti (Pubellier et al., 2000) and deformation at the Beata Ridge

offshore southern Haiti (Mauffret and Leroy, 1999). The deformation observed onshore the

Southern Peninsula during the late Oligocene – early Miocene is therefore possibly related to

increased compression in the northern Caribbean realm.

 Renewed subsidence started in the Burdigalian and lasted until the Messinian (Fig. 19e). The

l’Asile Basin became progressively deeper, with lacustrine and lagoonal sedimentation in the

Burdigalian giving way to marine conditions that dominated from the Langhian onwards. The

arrest of compressive deformation on the Southern Peninsula is possibly related to the

initiation of the Septentrional – Oriente fault system, along which Hispaniola separated from

Cuba from 20 Ma onwards (Pindell and Barrett, 1990; Calais and Mercier de Lépinay, 1995).

The eastward escape of Hispaniola shifted the area of compressional deformation from the

Southern Peninsula to the Haitian Fold-and-Thrust belt in the north.

 Distributed strike-slip activity started during the late Miocene (Fig. 19e), which became

focused along the present-day EPGFZ in the latest Miocene or Pliocene (Fig. 19f). A late

Miocene timing for the onset of strike-slip activity along the EPGFZ is also found in Jamaica,

which is recorded by the inversion of older Paleocene rift-related basins (Abbott et al., 2013;

James-Williamson et al., 2014). Strike-slip deformation became progressively more

transpressive during the late Miocene, which resulted in uplift and shortening of the l’Asile

Basin. Uplift is documented by a shallowing of facies and by erosion cutting down to the

Cretaceous basalts during the Messinian. CLIP-derived erosion products are re-deposited in

sediments from the Pliocene onwards.

123
Chapter 4: Geology of the Southern Peninsula

5.6 Structural style of deformation

The Southern Peninsula is a large anticlinorium bounded by offshore oblique thrusts towards the north

and south (Fig. 20). The half-wavelength of this structure increases from 55 km in the west at cross

section A, to around 70 km at cross section C, and 90 km at section D (Fig. 20). In the northwestern

part of the Southern Peninsula the general trend of this anticlinorium is east – west (Fig. 20). Towards

the southeast this trend progressively changes to a NW – SE orientation (Fig. 20). In the Dominican

Republic, the structural trend in the Sierra de Bahoruco, which is the eastward continuation of the

Massif de la Selle, also becomes progressively NW – SE oriented (Van den Berghe, 1983a). This change

in trend from west to east and north to south is likely related to the indentation of the Southern

Peninsula - Beata Ridge basement block into central Hispaniola, which started in Miocene times

(Pubellier et al., 1991; Mauffret and Leroy, 1999). The first-order folds (brown axes, Fig. 20) have

similar trends as the main anticlinorium and wavelengths between 15 and 30 km (Fig. 20). The axes of

these folds are displaced, but their structural trend not disrupted by, the EPGFZ. Only the trends of the

second-order folds (blue axes, Fig. 20) are controlled by the EPGFZ as well as other, minor faults. This

is evidenced by the sub-parallel trend of the second-order folds in proximity to the segments of the

EPGFZ (Fig. 20). The wavelengths of these folds varies between 3 and 10 km. We interpret these

second-order folds as being controlled by strike-slip and associated thrust faults, whose spacing and

dip angles are responsible for the variations in fold wavelength and orientation. In our interpretation

these faults mainly penetrate the CLIP basalts and sedimentary intercalations and are rooted in the

crust (Fig. 20), as opposed to shallow decollement structures. The first-order folds appear unrelated

to the mapped faults (Fig. 20). The 15 to 30 km wavelength of these folds is in the same order of

magnitude as the 16 to 30 km crustal thickness of the Southern Peninsula (Corbeau et al., 2017). This

could indicate that these folds are not directly controlled by faults, but are the result of large-

wavelength buckling of the crust itself in response to the regional stress regime. An anti-correlation

exists between regions of high elevation and basins on the Southern Peninsula (Fig. 20). In the west

the highest elevations are directly north of the EPGFZ with the basins located to the south, which is

124
Chapter 4: Geology of the Southern Peninsula

6. Conclusions

The Cenozoic tectonic history of the Haitian Southern Peninsula is polyphase and we distinguish three

major tectonic events.

1. Deformation and uplift during the Maastrichtian to early Paleocene. This deformation event

is expressed by crustal-scale folds that developed coeval with predominantly south or

southwest verging thrusts. Associated uplift caused partial erosion of the Upper Cretaceous

sedimentary cover and of the CLIP basalts in large parts of the Southern Peninsula.

Deformation resulted from NNE-directed compression. Following this deformation phase the

Southern Peninsula experienced general subsidence and a transgressive period, associated

with a deepening of depositional facies that continued until the Aquitanian.

2. Deformation and uplift in the early Miocene. This deformation phase mainly affected the

southwestern parts of the Southern Peninsula. It is characterized by an eastwards decrease in

uplift and erosion, and is not expressed in the eastern Southern Peninsula. Uplift resulted from

NE – SW shortening and contemporaneous folding, which was possibly aided by reactivation

of older thrusts in the western part of the peninsula. This event is likely related to a

contemporaneous reconfiguration of the plate boundaries in the northern Caribbean. The

depositional environment from the Burdigalian onwards is characterized by transgressive

facies and a return to marine conditions, which lasted until the Messinian.

3. Strike-slip followed by transpressional activity from late Miocene to recent. Strike-slip

activity along the EPGFZ started in the late Miocene. Deformation was initially distributed and

became progressively more focused along the EPGFZ from the Messinian or Pliocene onwards.

This process created a new plate boundary which separated the Gonâve microplate from the

Caribbean plate. The paleo- σ1 principal stress axis was sub-perpendicular to the EPGFZ,

indicating that the fault zone is mechanically weak. Initial strike-slip motion along the EPGFZ

became more transpressive through time. This transpressional phase is recorded by 1) uplift

of the Southern Peninsula form the Messinian onwards, 2) an increase in compressional

126
Chapter 4: Geology of the Southern Peninsula

deformation from West to East, possibly related to progressive indentation of the Beata Ridge

into central Hispaniola, which is manifested by 3) a change in structural style along the EPGFZ,

with predominantly strike-slip faults in the west and oblique-slip faults in the east. In the

Massif de la Selle thrust faults developed from the Pliocene onwards. These faults are spatially

associated with the EPGFZ and are rooted on this dominant fault at depth.

Acknowledgements

The authors are grateful to the Bureau de Mines et de l’Energie d’Haïti (BME) and the Université d’Etat

d’Haïti (UEH) for providing invaluable support during the field campaigns. This work greatly benefitted

from an IFPEn internal report on the stratigraphy of Haiti.

127
Chapter 4: Geology of the Southern Peninsula

Figure captions

Fig. 1a: Geodynamic setting of the Caribbean. GPS velocities (black arrows) from DeMets et al. (2010)
in a Caribbean reference frame. Fig. 1b: Geodynamic setting of Haiti. GPS velocity vectors (black
arrows) from Calais et al. (2016). GPS velocity vectors indicate motion of block to the south (west) of
fault with respect to block to the north (west). Faults modified after Leroy (1995). MCSC = Mid-Cayman
Spreading Center; OFZ = Oriente Fault Zone; SDB = Santiago Deformed Belt; SFZ = Septentrional Fault
Zone; NHDB = North Hispaniola Deformed Belt; HFTB = Haitian Fold-and-thrust belt; EPGFZ = Enriquillo-
Plantain Garden Fault Zone; WFZ = Walton Fault Zone; D.R. = Dominican Republic; P.R. = Puerto Rico;
V.I. = Virgin Islands; PRVI = Puerto Rico – Virgin Islands block.............................................................. 79
Fig. 2a: Structural map of the Southern Peninsula. Onshore shaded relief is 30m resolution ASTER
DEM, offshore bathymetry is at 25m resolution and was collected during the Haiti-SIS mission. Offshore
faults from Bien-Aime Momplaisir (1986) and Mauffret and Leroy (1997). Onshore structures from
Mann et al. (1995), Saint Fleur et al. (2015), Symithe and Calais (2016). GPS velocity vectors constructed
from Calais et al. (2016). Earthquake focal mechanisms from the International Seismological Centre
(2014) (GCMT). Fig. 2b: Compilation and simplification of 1:250.000 geological maps published by the
Bureau des Mines et de l'Energie d'Haïti (BME) after Boisson and Bien-Aime Momplaisir, 1987; Bien-
Aime Momplaisir et al., 1988). Black rectangles show extent of small scale geological maps and
corresponding figures. Black lines are cross sections. .......................................................................... 81
Fig. 3: Synthetic stratigraphic column for the Southern Peninsula. Relative sealevel for northeast
(blue), central (brown) and southwest (green). Data from references in text. Chronostratigraphic chart
from the International Commission on Stratigraphy (ICoS) based on the geologic time scale from
Gradstein et al. (2012). .......................................................................................................................... 84
Fig. 4a: Compilation map of the Cretaceous – Paleocene unconformity based on biostratigraphic ages
for Paleocene deposits and underlying formations. Color coding indicates erosion. Red; strong, erosion
down into Cretaceous basalts. Yellow; mild, erosion down into Cretaceous Macaya Formation
limestones. Green; conformable, biostratigraphically complete and concordent transition from
Cretaceous to Paleogene. Fig. 4b: Compilation map of the Cretaceous – Paleocene unconformity based
on biostratigraphic ages for lower to middle Miocene deposits and underlying formations. Color coding
indicates erosion. Red; strong, erosion down to Cretaceous. Orange; medium, erosion down to
Paleocene or Eocene. Yellow; mild, erosion down to Oligocene or intra-Miocene. Green; conformable,
biostratigraphically complete and concordent transition from Paleogene to Neogene. ..................... 91
Fig. 5a: Unconformable contact between CLIP basalts and calcareous clay-, silt- and sandstones of the
lower Paleocene, overlain by upper Paleocene clastic limestones. Top of ridge consists of lower Eocene

128
Chapter 4: Geology of the Southern Peninsula

cherty limestones. Northwest of Morne La Visite. Fig. 5b: Unconformable contact between CLIP basalts
and middle Miocene limestones. Haut Fort, south of Tiburon. ............................................................ 92
Fig. 6a: Boudinage of intercalated limestones within CLIP basalts. Northwest of Les Anglais. Fig. 6b:
Disharmonically folded Macaya Formation limestones with WNW-trending fold axes. Tiburon Valley.
Fig. 6c: Lower Eocene platform limestone gently folded along ENE-trending fold axis. Tiburon Valley.
Fig. 6d: North – south trending normal faults in middle Miocene chalks. NW of Port-a-Piment......... 94
Fig. 7a: Thrust contact between Macaya Formation limestones and CLIP basalts. Contact is brecciated
and underlying basalts are hydrothermally altered. SE of Mornes. Fig. 7b: Tight southwest verging folds
within Paleocene calcarenites. SE of Mornes. ...................................................................................... 95
Fig. 8a: Fault with unknown sense of motion separating Eocene platform limestones from CLIP basalts.
West of Obleon, northern flank of Massif de la Selle. Fig. 8b: Fault core of the EPGFZ. South-dipping,
east – west trending fault mirrors. La Boule quarry. Fig. 8c: Southwest verging thrust separating upper
Miocene – lower Pliocene marlstones in the footwall from lower to middle Miocene chalky limestones
in the hanging wall. Les Cayettes quarry............................................................................................... 97
Fig. 9: Geological map of the Tiburon area. For location see Fig. 2b. Black bedding symbols with dip in
degrees are measured in the field. Black fold axes represent folds constructed using outcrop
measurements. Bedding symbols in grey are obtained using satellite imagery. Solid black lines are
observed structures, dashed black lines denote structures from literature, while dashed grey lines are
structures based on satellite image interpretation. ............................................................................. 99
Fig. 10: One-to-one scale cross sections of the western Massif de la Hotte. Locations of cross sections
in Fig. 9. The topography is derived using ASTER DEM data. Faults shown as solid red lines are observed
while dashed red lines are from satellite interpretation. Grey lines within the formations denote the
inferred general trend of the bedding. Only the large-scale structures are shown. Dip of EPGFZ
unknown. Abbreviations; App. = Appendix, EPGFZ = Enriquillo – Plantain Garden Fault Zone. ......... 100
Fig. 11: Geological map of the l’Asile area. For location see Fig. 2b. Symbols are the same as described
in the caption of Fig. 9. ....................................................................................................................... 102
Fig. 12: One-to-one scale cross sections of the l’Asile Basin area. Locations in Fig. 11. Symbols are the
same as described in the caption of Fig. 10. Dip of EPGFZ unknown. ............................................... 103
Fig. 13: Geological map of the Marigot area. For location see Fig. 2b. Symbols are the same as
described in the caption of Fig. 9. ...................................................................................................... 105
Fig. 14: One-to-one scale cross section of the Massif de la Selle. Locations in Fig. 13. Symbols are the
same as described in the caption of Fig. 10. ...................................................................................... 106
Fig. 15a: Plot of trends of sigma 1 (black) and sigma 3 (blue) axes from stress inversion results. Fold
axes derived from outcrops are plotted in green. Fig. 15b: Chart showing angular difference (degrees)
between trend of the sigma 1 or sigma 3 axes and the trace of the EPGFZ. Colors indicate age of

129
Chapter 4: Geology of the Southern Peninsula

lithologies from which the data was obtained. Square indicates normal stress regime, circle indicates
strike-slip stress regime and hashed circle indicates reverse stress regime. Results on the right-hand
side are constructed using the N095°E average trend of thrusts in this region. Red circle is data from
the core of the EPGFZ. Plotted results are at less than 15km distance from the EPGFZ trace. Angular
difference varies between -20 and +100 degrees (see inset on right). .............................................. 110
Fig. 16a: Plot showing trend of sigma 1 axes vs distance to the trace of the EPGFZ for measurements
within 15 km of this fault. Arrows indicate dominant trends. Measurements obtained from the fault
core are within the red hashed zone. Fig. 16b: Same principle as figure 16a but for fold axes derived
from outcrops. Fig. 16c: Comparison between groups A, B and C from averaged σ1 trend and averaged
axes of contraction. ............................................................................................................................. 111
Fig. 17: Block cartoons showing the fault geometries corresponding to the different scenarios for the
faults bordering the southern Cul-de-Sac plain. Fig. 17a and Fig. 17b correspond to scenario 1 as shown
in Fig. 14a. Fig. 17c and Fig. 17d correspond to scenario 2 as shown in Fig. 14b. Fig. 17e is a simplified
geological map from Fig. 13. Black box corresponds to the geographical extent of the block cartoons
in Figs. 17a and 17c. ............................................................................................................................ 116
Fig. 18: Compilation chart for the Southern Peninsula displaying relative sea level, stratigraphy,
deformation event, uplift and subsidence, and the associated stress regime. For discussion see text.
............................................................................................................................................................. 118
Fig. 19: Evolutionary cross section through the l’Asile Basin, based on cross section D from Fig. 12d.
............................................................................................................................................................. 122
Fig. 20: Top map is a westwards facing perspective view of the Southern Peninsula. Background
bathymetry is 30 arc-second interval data from GEBCO, high-resolution overlay bathymetry is at 25m
resolution collected during the Haiti-SIS mission, onshore shaded relief is 30m resolution ASTER DEM.
Elevation and bathymetry of the map are 2x exaggerated. Offshore faults modified from Bien-Aime
Momplaisir (1986) and Mauffret and Leroy (1997) to match the bathymetry. Fold axes interpreted
using the above data, the BME geological map (Fig. 2b) and our small-scale geological maps (Fig. 9,
Fig. 11 and Fig. 13). Cross sections are on a 1-to-1 scale, with the relative spacing reflecting the
distance between them: A and B are from the western Massif de la Hotte (A and B in Fig. 10), C is from
the l’Asile region (C in Fig. 12), and E is from the Massif de la Selle (Fig. 14). Offshore bathymetry
constructed from GEBCO data. ........................................................................................................... 125

130
Chapter 4: Geology of the Southern Peninsula

Appendix 1a Spatial distrubution of latest Cretaceous - Paleocene unconformity and erosion


Lithologies; Erosion
Basalts Marl strong
Silty limestone Chalky/marly limestones Platform limestone Breccia
mild
Volcaniclastic Cherty pelagic limestones Chalk Silt/clay Calcareous siltstone Conglomerate Microbreccia conformable
unknown
E Than. - Ypr. 16 Ypr. - Lut.
Sel. - Than. 7 Dan. - Ypres. 64 - 58 Ma D Maastrichtian Than. - Ypr.
Danian 27 Dan.-Than. Than. - Lut.
Camp. 39A Maastrichtian 24 C Maastrichtian
15
9 Danian Dan. - 10 Dan. 6 Danian? 8 Danian
38 Danian? 25 31 Dan. - Than. Sant. - Camp.- 39B Camp. - Maast. Camp.- 14 Con. - Camp.
30 Than. Camp. - Maast. 26 Pal.? 40 Maast. 44 Maast. 43 Maast. B 13
Cretaceous Cret. 5 Camp. Cretaceous Cretaceous Cretaceous Cretaceous Camp. - Maast. Santonian
-74° -73° -72°

Jeremie
Dame Marie Roseaux PORT-AU-PRINCE
Léogâne
18.5°

18.5°
9
Baradères
30 10 38
26 40 Miragoane
Anse-d’Hainault 27 16 Malpasse
7 Grand Goâve 15
8
39
MASSIF DE LA HOTTE 6
4
5 25
31
44 24
13 14
3 Maniche 43
Beloc MASSIF DE LA SELLE
Tiburon Grand Caille
29 1 2
Clonard L’Asile 32 45 19
18
Seguin
28 Platon Besace 17
20 22 23
Camp Perrin 48 Marbial
Les Anglais 34
41 46 21
Touya 47
Port-a-Piment 37 Chantal
12 42 Jacmel Marigot
35
Les Cayes
Bainet

q
33

Port Salut 11 36 SOUTHERN PENINSULA


18°

18°
0 20 40 80
Kilometers
-74° -73° -72°
Maast. Dan.-Than. Dan. Dan. - Than. 32A Cret.
Sel. - 41 Camp. - Maast. Dan. - Pal.? 20 Danian 23 Ypresian
28 37 12 33 Dan. 48 47 Than. Cretaceous 46 42 Maast.
Cret.
Cret. Danian <U. Maast.?
32B
Ypr. Sel.
Cret. 19 Maastrichtian 22 Dan. - Than.
Cretaceous 34 18 Campanian 21 Danian
Tur. - Maast. 17 Con. - Camp. Cretaceous
Camp. 4 Maastrichtian Dan. - Dan. <U. Maast.? 45C Than. - Ypr.
29 3 Sant. - Camp. 11 Than. 36 35 45B Dan. - Than.
Cret.
2 Santonian Cret. Cretaceous 45A Maast.?
1 Con. - Sant. Cretaceous

131
Chapter 4: Geology of the Southern Peninsula

Appendix 1b
Range Combined Range
Ma / Zone Age Ma / Zone Age
ID Author Location Lithology Fauna Ref. FAD LAD FAD LAD FAD LAD FAD LAD Description
1 Cal83 Ravine du Sud Chalky limestone and silicified limestone Dicarinella Concavata 1 89,9 84 Coniacian Santonian 89,9 84 Coniacian Santonian Macaya
Globotruncana coronata 1 89,9 79,2 Coniacian Campanian
Globotruncana cyrenaica 1 89,9 84,2 Coniacian Santonian
Globotruncana fornicata 1 89,9 67,3 Coniacian
2 Cal83 Base Rampe de Marcelline limestone, pelagic silicified limestone 86,7 84,2 Santonian Santonian Macaya
Globotruncana arca convexa 1 86,7 67,3 Santonian
Globotruncana sigali 1 93,5 84,2 Turonian Santonian
Globotruncana concavata 1 89,9 84 Coniacian Santonian
Globotruncana s 1 86,7 67,3 Santonian
Globotruncana tricarinata 1 93,9 83,6 Turonian Santonian
3 Cal83 Top Rampe de Marcelline Argillaceous limestone and pelagic silicified limestone 83,6? 74? Santonian? Campanian? Macaya
Globotruncana stuar 1 73 67,3 Campanian
Globotruncana calcarata 1 80 74 Campanian Campanian
Globotruncana fornicata 1 89,9 67,3 Coniacian
Massive limestone with argillaceous intervals, radiolarites Globotruncana arca convexa 1 86,7 67,3 Santonian
4 Cal83 Tombeau Cheval 69 66 n Macaya
and tuffs Micula staurophora 2 89,8 66 Coniacian Maastrichtian
Tetralithus murus 2 69 66 Maastrichtian Maastrichtian
Globotruncana stuar 1 73 67,3 Campanian
5 Cal83 Plaine Limestone Globotruncana fornicata 1 89,9 67,3 Coniacian 73 67,3 Campanian Macaya
Globotruncana arca convexa 1 86,7 67,3 Santonian
Globigerina <Cretaceous 3 P0 P2 Danian Danian
Rivière Glace bridge, road Camp Heterohelix 1 & 3 100 66 Cenomanian
6 Cal83 Siltstone, claystone, 66 62 Danian Danian Rivière Glace 75m thick sequence
Perrin - Jeremie Hedbergella 3 113 62,5 Albian? Danian
NO Globotruncana 3 <66Ma <66Ma <Cret. <Cret.
Base; limestone with detrital plagioclase Globorotalia pusilla pusilla 1 P3a P4 Danian
62 61 Danian Selandian
1km north of Rivière Glace bridge, intervals Globorotalia uncinata 1 P2 P3a Danian Danian
7 Cal83 Overlying the Rivière Glace
road Camp Perrin - Jeremie Overlying base; limestone overlain by marls/clays Globorotalia pseudomenardii 1 P4 P4 Selandian 60,5 57 Selandian
Overlying are conglomerates and limestones Globorotalia aequa 1 P4c P5 Thanetian Ypresian 57,8 50,7 n Ypresian
Globotruncana stuar 1 73 67,3 Campanian
Microbreccia overlain by limestones (Baradères
From clasts in microbreccia Globotruncana arca convexa 1 86,7 67,3 Santonian 73 67,3 Campanian
8 Cal83 3.5 km SE of Baradères capped by basalts. Probably unconformable on
Globotruncana lapparen 1 86,7 67,3? Santonian
Macaya
Overlain by limestone capped by basalt Polyps, gasteropods, algae, mollusks K/Ar on basalts 64 58 Danian
Base, south; deformed Cretaceous basalts Cretaceous Cretaceous Deformed Cretaceous basalts unconformably overlain by
Between Dame-Marie and Anse-
9 Cal83 Overlying; detrital series with intervals Danian? Danian? undeformed and (Rivière Glace), in turn
d'Hainault
Top, north; limestone Globorotalia pseudomenardii 1 P4 P4 Selandian 61 57 Selandian overlain by limestones
Base; deformed Cretaceous basalts none Cretaceous Cretaceous
Between Anse-d'Hainault and Source Hedbergella 3 113 62,5 Albian? Danian Deformed Cretaceous basalts unconformably overlain by
10 Cal83 Overlying; Detrital series with intervals -
Chaude; Ravine Bras à Droite Globigerina <Cretaceous 3 P0 P2 Danian Danian <66 62,5 <Cret Danian undeformed and sequence (Rivière Glace)
samples are reworked fragments
Globotruncana s 1 86,7 67,3 Santonian
Globorotalia angulata P3 P4 Danian Thanetian
along route Port Salut to Les
11 Cal83 Chalky limestone Chiasmolithus danicus NP3 NP6 Danian Thanetian 62,3 59 Danian Selandian chalky limestones
Cayes
Ellipsolithus macellus NP4 NP11 Danian Ypresian
12 Cal83 Les Cayes plain, 1 km W of Chantal Calcarenites and marls overlying Cret basalts no fauna given, only zones NP1 NP2 Danian Danian 66 64,7 Danian Danian Calcarenites unconf. overlying deformed Cret. Basalts
Globotruncana lapparen 1 86,7 67,3? Santonian
Globotruncana carinata 1 86,7 84,2 Coniacian Santonian
Globotruncana fornicata 1 89,9 67,3 Coniacian in Dumisseau deposited under pelagic
13 VdB83 Between Obleon and Furcy Marly within basalts 86,7 84,2 Santonian Santonian
Heterohelix 1&3 100 66 Cenomanian Maastrichtian open marine CLIP
Watznaueria bernesea 2 168,2 66 Bathonian
Micula staurophora 2 89,8 66 Coniacian Maastrichtian
14 VdB83 Between Obleon and Furcy Basalt none K/Ar on basalts 88 82 Coniacian Campanian Basalt of the Dumisseau CLIP
limestones; Lepidocyclina, Melobesia algae, Ranikothalia sp. (bermudezi) 3 SBZ4 SBZ5 n Ypresian limestones separated by fault from underlying
15 VdB83 SW of Obleon 58 55,5 n Ypresian
Miliolida Discocyclina barkeri 4 SBZ4? SBZ7? Thanetian Ypresian Cretaceous basalts of 13/14
Discoaster barbadiensis 2 NP11 NP19-20 Ypresian Priabonian
Discoaster lodoensis 2 NP12 NP14 Ypresian Lutetian
Cyclococcolithus formosus 2 NP12 NP21 Ypresian Rupelian
Massive chalky limestones; overlying limestones
16 VdB83 SW of Obleon Sphenolithus radians 2 NP11 NP23 Ypresian Rupelian 53,7 46,3 Ypresian n Massive chalky limestone overlying limestones
of #15
Globorotalia aspensis 1 E6 E8 Ypresian Lutetian
Globorotalia bullbrooki 1 E7a E11 Ypresian Bartonian
Truncorotaloides rohri 1 E6 E13 Ypresian Bartonian
Manivitella pemmatoides 2 147,7 66 Tithonian
Watznaueria bernesea 2 168,2 66 Bathonian
Rivière Gosselin Directly NE Massif basalts, tuff and sedimentary of the
17 VdB83 Marly within basalts Micula staurophora 2 89,8 66 Coniacian Maastrichtian 89,8 72 Coniacian Campanian
of Marbial Dumisseau CLIP
Marthasterites furcatus 2 93,7 72 Turonian Campanian
Eiffellithus turriseiffeli 2 103,3 66 Albian

132
Chapter 4: Geology of the Southern Peninsula

Globotruncana elevata 1 84,2 68? Campanian


Globotruncana fornicata 1 89,9 67,3 Coniacian
Rivière Gosselin Directly SW Globotruncana s 1 86,7 67,3 Santonian
18 VdB83 Massive biomicrite with cherts 79,2 74 Campanian Campanian Base Macaya
of Marbial Globotruncana arca convexa 1 86,7 67,3 Santonian
Globotruncana calcarata 1 80 74 Campanian Campanian
Globotruncana falsostuar 1 79,2 67,3 Campanian
Globotruncana elevata 1 84,2 67,3? Campanian
Globotruncana stephensoni 1 83,6 67,3 Campanian
Rivière Gosselin SW of
19 VdB83 Massive biomicrite with cherts Globotruncana contusa 1 70,1 67,3 n 70,1 67,3 n Top Macaya (total thickness 300m)
Marbial
Globotruncana lugeoni-gansseri 1 71 67,3 n
Globotruncana gagnebini 1 74 67,3 Campanian
Globigerina pseudobulloides 1 Pa P3 Danian Selandian
Rivière Gosselin SW of Basal conglomerate overlain by clay/silts with volcanic Globigerina triloculinoides 1 P1b P4 Danian Marigot overlying Macaya by
20 VdB83 63,9 62,6 Danian Danian
Marbial debris and marls Globorotalia edita 1 Pa P1c Danian Danian (transgressive?) unconformity (160m thick; BRGM88)
Globorotalia compressa 1 P1c P3 Danian Selandian
Macary?; between Marigot and Silt/marlstone with sandstone containing grains of basalt Globigerina pseudobulloides 1 Pa P3 Danian Selandian Marigot elements of CLIP basalts within marl-clay-silt-
21 VdB83 66 62,6 Danian Danian
Seguin (bathyal, open marine) Globigerina daubjergensis 1 Pa P1c Danian Danian sandstone sequence (>250m thick)
Globorotalia angulata 1 P3 P4 Danian
22 BRGM88 North of Macary?; overlying #19 Argillaceous limestones 60,7 52,2 Danian n Overlying Marigot (125m thick)
Globorotalia pseudomenardii 1 P4 P4 Selandian
23 BRGM88 Seguin Platform limestones Globorotalia edgari 1 E2 E3 Ypresian Ypresian 55,9 54,6 Ypresian Ypresian Top of sequence in this region (100m thick)
Blocks within the sequence; containing basalts, dolerites, Globotruncana concavata 1 89,9 84 Coniacian Santonian
24A cherts, silicified limestones, massive limestones, sandy Globotruncana sigali 1 93,5 84,2 Turonian Santonian 89,9 84,2 Coniacian Santonian Series of blocks on the road from Leogâne to Jacmel.
limestones Globotruncana coronata 1 89,9 79,2 Coniacian Campanian
Globotruncana tricarinata (reworked?) 1 93,9 83,6 Turonian Santonian
Globotruncana s 1 86,7 67,3 Santonian
Globotruncana arca convexa 1 86,7 67,3 Santonian
24B Matrix from the blocks Globotruncana globigerinoides 1 92,6 71,8 Turonian 73 71,8 Campanian n Matrix for the series of blocks
Globotruncana marginata (reworked?) 1 93,5 84,2 Turonian Santonian
Globotruncana stephensoni 1 83,6 67,3 Campanian
Around Grand Savanne?; Route
MdL79 Globotruncana stuar 1 73 67,3 Campanian
Leogâne to Jacmel
Globotruncana s 1 86,7 67,3 Santonian
Overlying Series of blocks; Greywackes with conglomerate Globotruncana elevata 1 84,2 67,3? Campanian Greywackes and conglomerates unconformably(?) overlying
24C 70,1 67,3 n
intervals Globotruncana contusa 1 70,1 67,3 n Series of blocks (50m thick roughly)
Globotruncana stuar 1 73 67,3 Campanian
Globotruncana contusa 1 70,1 67,3 n
Overlying greywackes&conglomerates; Argillaceous Globotruncana gansseri 1 71 67,3 n Argillaceous limestones overlying greywackes&conglomerate
24D 70,1 67,3 n
limestones Globotruncana lugeoni 1 71 67,3 n interval (at least 450m thick)
Globotruncana stuar 1 73 67,3 Campanian
24E Overlying argillaceous limest.; limestones Ranikothalia sp. 3 SBZ4 SBZ5 Thanetian Ypresian 58 55,5 n Ypresian Overlying limestones
25 Am97 Policard Biomicrite with chert thrusted over Globotruncana calcarata 1 80 74 Campanian Campanian 80 74 Campanian Campanian Macaya Fm. thrusted over Rivière Glace Fm.
Globotruncana calcarata 1 80 74 Campanian Campanian
Globotruncana arca convexa 1 86,7 67,3 Santonian
26 Am97 Ramp towards Fond Tortue Massif marly limestone 80 74 Campanian Campanian Macaya
Globotruncana elevata 1 84,2 67,3? Campanian
Globotruncana bulloides 1 86,7 67,3 Santonian
Globotruncana a 1 93,5 92,6 Turonian Turonian
Clasts in conglomerate Heterohelix sp. 1&3 100 66 Cenomanian Maastrichtian 93,5 92,6 <Turonian >Turonian Age of clasts in conglomerate discordantly overlying Macaya Fm.
Hedbergella 3 113 62,5 Albian? Danian
27 Am97 Between Baradères and Nicolas(?)
Dictyoconus sp. 3 ?? 23 Danian? Chattian
limestones overlying conglomerate 66? 56? Danian? Limestone overlying conglomerate
Planorbulina cretae 5 ?? ?? Danian? Thanetian?
Basalts overlying limestones none limestones capped by basalts
Globotruncana arca convexa 1 86,7 67,3 Santonian
Globotruncana s 1 86,7 67,3 Santonian
Around Fonds Charlemagne (N of Les
28 Am97 Silicious biomicrite Globotruncana contusa 1 70,1 67,3 n 70,1 67,3 n Macaya Fm in faulted contact with CLIP basalts
Anglais)
Globotruncana lapparen 1 86,7 67,3? Santonian
Globotruncana bulloides 1 86,7 67,3 Santonian
Globotruncana arca convexa 1 86,7 67,3 Santonian
Dm-bedded limestones with basal (micro?) conglomerate
Globotruncana elevata 1 84,2 67,3? Campanian Macaya Fm. overlying CLIP basalts, separated by a (micro?)
29 Am97 Chanterelle (NE of Les Anglais) containing Macaya Fm. and CLIP basalt clasts, overlying 80 74 Campanian Campanian
Globotruncana calcarata 1 80 74 Campanian Campanian conglomerate containing clasts of both
CLIP basalts
Globotruncana lapparen 1 86,7 67,3? Santonian
30 Am97 Franklin (E of Anse-d'Hainault) Dacites none K/Ar on basalts 63 56 Danian n Dacites proximal to of the Rivière Glace Fm.
Between Sinaï and Maquisard (N of Rivière Glace Fm. with clasts of Macaya Fm. and
31 Am97 Calcareous siltstones with clasts of Macaya Fm Planorbulina cretae 5 ?? ?? Danian? Thanetian? 66? 56? Danian Thanethian
l'Asile) lavas, unconformably overlying CLIP basalts
32A Marly limestones Globotruncana sp. 1 >93,9 <67,3 >Turonian <Danian <93,9 >67,3 <Turonian >Danian
Between Vieux Bourg d'Aquin and
Am97 Calcareous marls/silts with and Macaya Fm. limestone thrusted over Rivière Glace Fm.
l'Asile 66 61,6 Danian Danian
32B Globigerina <Cret 3 P0 P2 Danian Danian
Base; Clay, silt and sandstone with volcanic grains Globigerina <Cret 3 P0 P2 Danian Danian 66 61,6 Danian Danian
Well- limestones concordantly overlying detrital
33 Am97 Brieux (western Les Cayes plain) Globigerina <Cret 3 P0 P2 Danian Danian
Top; Biomicrite 63,9 57,1 Danian sequence
Globorotalia sp. 1 P1c P4 Danian Thanethian

133
Chapter 4: Geology of the Southern Peninsula

Base: Basalts with marly Globotruncana sp. 1 >93,9 <67,3 >Turonian <Danian <93,9 >67,3 <Turonian >Danian Basalts
Microconglomerate basalts from overlying silty
Middle; Microconglomerate none limestones
34 Am97 Tibarra (E of Cavaillon)
Micula staurophora 2 89,8 66 Coniacian Maastrichtian
(Paleocene?) Silty limestones containing reworked Cretaceous
Top; Silty limestones Watznaueria bernesea 2 168,2 66 Bathonian 67,3 66 n
fauna unconformably overlying basalts
Micula prinsi 2 67,3 66 Maastrichtian Maastrichtian
Micula murus 2 69 66 Maastrichtian Maastrichtian
Rivière Glace Fm. with basal conglomerate unconformably
35 Am97 Gross Caye (Island S of Aquin) facies with chalky intervals Micula staurophora 2 89,8 66 Coniacian Maastrichtian 69 66 n
overlying CLIP basalts
Tetralithus obscurus 2 93,9 66 Turonian Maastrichtian
Rivière Glace Fm. siltstones and calcarenites unconformably
36 Am97 L'Île à Vache Siltstones and calcarenites Globigerina <Cret 3 P0 P2 Danian Danian 66 61,6 Danian Danian
overlying CLIP basalts
Pelites and calcarenites overlying polymict breccia of basalts, in
37 Am97 Morne Verdun (SE of Port-a-Piment) Silts-clays with calcarenites Chiasmolithus danicus 2 NP3 NP6 Danian n 63,5 59,2 Danian
turn overlying CLIP basalts
Ranikothalia sp. (bermudezi) 3 SBZ4 SBZ5 n Ypresian
Caracoli (between Baradères and limestones with basal conglomerate unconformably
38 Am97 limestones Morozovella aragonensis 2 E5 E9 Ypresian 52,5? 41,2 ?
-Trou-de-Nippes overlying CLIP basalts
Eoconuloides sp. 3 SBZ4 SBZ20 Thanetian Priabonian
Globotruncana gansseri 1 71 67,3 n
39A RM86 South of Goâve Siliceous biomicrite Globotruncana contusa 1 70,1 67,3 n 70,1 67,3 n n Top Macaya
Globotruncana stuar 1 73 67,3 Campanian
Globotruncana stuar 1 73 67,3 Campanian
39B RM86 South of Goâve Biomicrites 73 67,3 Campanian n Base (?) Macaya
Globotruncana s 1 86,7 67,3 Santonian
Rivière Brossard (2km south of Anse-a-
40 RM86 Limestones with chert nodules Globotruncana s 1 86,7 67,3 Santonian n 86,7 67,3 Santonian n Macaya
Veau?)
Globotruncana arca convexa 1 86,7 67,3 Santonian
Globotruncana stephensoni 1 83,6 67,3 Campanian
41 RM86 South of Fond-des-Blancs Limestones 83,6 67,3 Campanian n Macaya overlying CLIP basalts
Globotruncana s 1 86,7 67,3 Santonian
Globotruncana fornicata 1 89,9 67,3 Coniacian
Globotruncana contusa 1 70,1 67,3 n
Globotruncana stuar 1 73 67,3 Campanian
42 RM86 Rivière Bas-de-la-Croix Calcirudites Globotruncana arca convexa 1 86,7 67,3 Santonian 70,1 67,3 n n Macaya
Globotruncana gansseri 1 71 67,3 n
Globotruncana calciformis 1 79,2 67,3 Campanian
Globotruncana arca convexa 1 86,7 67,3 Santonian
Globotruncana stuar 1 73 67,3 Campanian
43 RM86 Road towards Palmes Biomicrites with chert nodules 73 67,3 Campanian n Macaya
Globotruncana elevata 1 84,2 67,3? Campanian
Globotruncana stephensoni 1 83,6 67,3 Campanian
Globotruncana stephensoni 1 83,6 67,3 Campanian
Globotruncana s 1 86,7 67,3 Santonian
44 RM86 8km west of road towards Palmes Biomicrites with chert nodules Globotruncana arca convexa 1 86,7 67,3 Santonian 73 71,8 Campanian n Macaya
Globotruncana lapparen 1 86,7 67,3? Santonian
Globotruncana conica 1 73 71,8 Campanian
Watznaueria bernesea 2 168,2 66 Bathonian
45A Marlstone 89,8 66 Coniacian Sequence; top of calcarenites overlying silts in turn overlying
Micula staurophora 2 89,8 66 Coniacian Maastrichtian
RM86 Morne Madame Toussaint marls in turn overlying breccia in turn
45B Siltstone Globigerina triloculinoides 1 P1b P4 Danian 65,2 57,2 Danian
unconformably overlying CLIP basalts
45C Calcarenites Discocyclina barkeri 4 SBZ4? SBZ7? Thanetian Ypresian 58 54,5 n Ypresian
46 RM86 Bas-de-la-Croix Biomicrites (silty?) pseudobulloides 1 Pa P3 Danian Selandian 66 60,7 Danian Selandian (Silty?) biomicrites in faulted contact with CLIP basalts
Globigerina triloculinoides 1 P1b P4 Danian
47 RM86 Rivière Bois d'Orme Calcareous siltstones (?) Globorotalia velascoensis 1 P3b E2 Selandian Ypresian 57,7 57 n n Dm-bedded calcaroues siltstones (?)
Globorotalia aequa 1 P4c E5 Thanetian Ypresian
Globorotalia aequa 1 P4c E5 Thanetian Ypresian
48 RM86 Rivière Mombrun Calcarenites 61,3 55 Selandian Ypresian Calcarenites
Globorotalia velascoensis 1 P3b E2 Selandian Ypresian

134
Chapter 4: Geology of the Southern Peninsula

Appendix 2a Spatial distribution of early Miocene unconformity and erosion


Lithologies; Erosion
Basalts Marl strong
Silty limestone Chalky/marly limestones Platform limestone Breccia
medium
mild
Volcaniclastic Cherty pelagic limestones Chalk Silt/clay Calcareous siltstone Conglomerate Microbreccia conformable
unknown
13D Serravallian
25C Lang. - Tort. 7C Serravallian Lang. Burd. - Serr.? 10C Lang. - Serr.
13C Langhian 14B Serr. - Tort.
Lacustrine
Eoc. 5B Tort. - Mess. Aqui.- 25B <Burd. 7B Aqui. - Burd. 3 Aqui.- 27 Mio.? 10B Burd. - Lang. Serr. - 13B Chat. - Aqui. Conformable
24 Burd.-
Serr. 5A Aqui? - Burd. 26 Mess.-
Zanc. 6 Chat. 25A Oligo.? ? 7A Chat. - Aqui. Cret.?
2 Tort.
Conglomerate
Cret. - Pal.?
28
Bart. 10A Aqui. - Burd. 11 Tort. 12 Serr.
13A Pria. - Rup. 14A Lut. - Rup.
-74° -73° -72°

6
Jeremie
Roseaux
Dame Marie 26
PORT-AU-PRINCE
25 7
5 Léogâne
18.5°

18.5°
24
Baradères
Miragoane 12 13 Malpasse
Anse-d’Hainault Grand Goâve 10
11 14
MASSIF DE LA HOTTE 27 28
Tiburon
Camp Perrin Grand Caille Beloc MASSIF DE LA SELLE
Clonard L’Asile
3 2 Platon Besace Seguin
23 1 29 8
Maniche 9
4 17
30 Marbial
Les Anglais 22
21 18
Touya
15
Port-a-Piment Marigot
Jacmel
Les Cayes
Chantal Bainet

q
16

Port Salut
20 19
SOUTHERN PENINSULA
18°

18°
0 20 40 80
Kilometers
-74° -73° -72°
Serr.? Lang.- Serr.? 4B Serravallian Burd.- 1B Serr. Burd.- 30D Lang. - Serr. 8B Rup. - Burd. 9C Burdigalian Rup.- Rup.-
23 22 Serr. 21 4A 20 Serr. 19 Priab. 18 Burd. 17 Lang. 1A 29 Serr.? 30C Burd. - Lang. 9B 15 Aqui. 16 Chat.
Cret. Cret. Cret. Aqui. - Burd. Bart. 8A Rup. - Chat. Aquitanian
Eocene 30B Aquitanian 9A Rup. - Aqui.
30A Aquitanian

135
Chapter 4: Geology of the Southern Peninsula

Appendix 2b ŝŽƐƚƌĂƟŐƌĂƉŚLJĞĂƌůLJDŝŽĐĞŶĞƵŶĐŽŶĨŽƌŵŝƚLJ DĂ / Zone


RanŐe
Őe DĂ / Zone
Coŵbined ZĂŶŐĞ
ŐĞ
ID AƵtŚor LoĐation LitŚoůoŐLJ FaƵna ReĨ. FAD LAD FAD LAD FAD LAD FAD LAD DesĐription
Globigerinatheka semiinvoluta 1 E14 E15 Priabonian Priabonian
Globigerinatheka tropicalis 1 E8 E16 >ƵƚĞƟĂn Priabonian
Globorotalia cerroazulensis 1 E7? E16 >ƵƚĞƟĂn Priabonian
Globorotalia cocoaensis 1 E13 E16 Bartonian Priabonian
1A Pelagic mudstones 40 38,3 Bartonian Bartonian
Globorotalia spinulosa 1 E7b E13 Ypresian Bartonian
Truncorotaloides rohri 1 E6 E13 Ypresian Bartonian
Globugerina ƚƌŝƉĂƌƟƚĂ 1 E13 O6? Bartonian ŚĂƫĂŶ
Folded Eocene limestones unconformably overlain by 200m of
Cal83 South of Maniche Globigerinita dissimillis 1 E13 N6 Bartonian Burdigalian
Miocene marls
Orbulina universa 1 N9 Extant Langhian Present
Globigerinoides subquadratus 3 N4b N15 Aquitanian Tortonian
Globigerinoides obliquus 3 N5b PT1a Burdigalian Pleistocene
1B Marls with intercalated calcarenites Globigerina druryi 1 N7 N15 Burdigalian Tortonian 13,8 11,8 Serravallian Serravallian
Globorotalia praemenardii 3 N11 N13a Langhian Serravallian
Globorotalia fohsi 1&3 N9? N12 Langhian Serravallian
Globorotalia peripheroronda 1 M3 M12 Burdigalian Tortonian
Globigerinoides trilobus 3 N4b Extant Aquitanian Present
2 Cal83 Raymond; Camp Perrin region Marls 22,3 7,3 Aquitanian Tortonian Marine marls
Globigerinoides subquadratus 3 N4b N15 Aquitanian Tortonian
Globigerinoides trilobus 3 N4b Extant Aquitanian Present
Globigerinoides bisphericus 3 N7 N8 Burdigalian Langhian
Ravine Bras de Gauche; W of Camp Globigerinoides subquadratus 3 N4b N15 Aquitanian Tortonian Lagoonal/lacustrine marls and lignites with basal conglomerate
3 Cal83 Marls, lignite, silts, clays, minor argillaceous limestone 16 14 Langhian Langhian?
Perrin Globorotalia mayeri 3 P22b N13 ŚĂƫĂn Serravallian unconformably overlying Cretaceous limestones
Globorotalia fohsi (?) 1&3 N9? N12 Langhian Serravallian
Globorotalia peripheroronda 1 M3 M12 Burdigalian Tortonian
Globigerina ƚƌŝƉĂƌƟƚĂ (reworked?) 1&3 P14 P22a Bartonian ŚĂƫĂŶ
Globigerinoides ĂůƟĂƉĞƌƚƵƌƵƐ 1&3 N4b N8a Aquitanian Burdigalian
Globigerinoides subquadratus 3 N4b N15 Aquitanian Tortonian
4A Marls ŝŶƚĞƌƐƚƌĂƟĮĞĚ with conglomerates 22,4 17,4 Aquitanian Burdigalian
Globigerinoides trilobus 3 N4b Extant Aquitanian Present
Globorotalia mayeri 3 P22b N13 ŚĂƫĂn Serravallian Conglomerate with clasts of Macaya Fm. and CLIP basalts
Globigerinita dissimillis 1 E13 N6 Bartonian Burdigalian unconformably overlying folded Eocene limestones. Towards the
Cal83 ^ĞĐƟŽŶ Monville - Touya
Orbulina universa 1 N9 Extant Langhian Present top reefal limestones topped by silts and clays with
Orbulina bilobata 3 N9 Extant Langhian Present ŵŝĐƌŽĐŽŶŐůŽŵĞƌĂƟĐ intervals
Globorotalia mayeri 3 P22b N13 ŚĂƫĂn Serravallian
4B Clays and silts 13,4 11,8 Serravallian Serravallian
Globorotalia lenguaensis 1&3 N12 N17 Serravallian Messinian
Globorotalia fohsi 1&3 N9? N12 Langhian Serravallian
Globigerinoides subquadratus 3 N4b N15 Aquitanian Tortonian
5A Cal83 Monoré Marls none given N5 N6 Aquitanian? Burdigalian 21,2 17,6 Aquitanian? Burdigalian Marls conformably overlying limestones
5B Cal83 Rivière Bourdon Marls none given N16 N17 Tortonian Messinian 9,7 6,5 Tortonian Messinian Marls
Globigerina venezuelana 1&3 P10 N19 >ƵƚĞƟĂn Zanclean
Globigerina ƚƌŝƉĂƌƟƚa 1&3 P14 P22a Bartonian ŚĂƫĂŶ
Globigerinita unicavus 3 P14 N7 Bartonian Burdigalian
6 Cal83 Jeremie Chalky limestones with cherts 25,2 21 ŚĂƫĂn Aquitanian Bathyal marine chalky limestones with cherts
Globorotalia kugleri 3 N4 N4 Aquitanian Aquitanian
Globorotalia mendacis 1&3 P21 N4 ŚĂƫĂn Aquitanian
Globorotalia mayeri 3 P22b N13 Chattian Serravallian
Globigerina ƚƌŝƉĂƌƟƚa 1&3 P14 P22a Bartonian ŚĂƫĂŶ
Globigerinoides trilobus 3 N4b Extant Aquitanian Present
Globigerinoides subquadratus 3 N4b N15 Aquitanian Tortonian
7A Chalky limestones with cherts 25,2 21 ŚĂƫĂŶ Aquitanian Bathyal marine chalky limestones with cherts
Globorotalia kugleri 3 N4 N4 Aquitanian Aquitanian
Globorotalia mayeri 3 P22b N13 Chattian Serravallian
Globorotalia acrostoma 1&3 P22b N12a ŚĂƫĂn Serravallian
Globigerinoides trilobus 3 N4b Extant Aquitanian Present
Globigerinoides subquadratus 3 N4b N15 Aquitanian Tortonian
Globorotalia mayeri 3 P22b N13 Chattian Serravallian
7B Cal83 South of Roseaux Marls and limestones 23 17,6 Aquitanian Burdigalian Marls and limestones
Globorotalia peripheroronda 1 M3 M12 Burdigalian Tortonian
Globigerinoides ĂůƟĂƉĞƌƚƵƌƵs 1&3 N4b N8a Aquitanian Burdigalian
Globigerinita stainforthi 1 P22 N5 ŚĂƫĂn Burdigalian
Orbulina universa 1 N9 Extant Langhian Present
Globigerinoides trilobus 3 N4b Extant Aquitanian Present
Globigerinoides subquadratus 3 N4b N15 Aquitanian Tortonian
7C Marls with intercalated siltstones 13,4 11,6 Serravallian Serravallian Marls with intercalated siltstones
Globigerinoides obliquus 3 N5b PT1a Burdigalian Pleistocene
Globorotalia mayeri 3 P22b N13 Chattian Serravallian
Globorotalia menardii 1 N12 Extant Serravallian Present

136
Chapter 4: Geology of the Southern Peninsula

Globigerinita dissimillis 1 E13 N6 Bartonian Burdigalian


Globigerinita unicavus 3 P14 N7 Bartonian Burdigalian
8A Chalks Globugerina a 1 E13 P22 Bartonian 29,8 25,2 Rupelian n Chalks
Cassigerinella chipolensis 1&3 P18 N13 Rupelian Serravallian
Morne à Boite; Between Blockauss Globorotalia opima sp. 3 P20b P21b Rupelian
VdB83
and Trouin Coccolithus abisectus 2 NP23 NN6 Rupelian Serravallian?
Coccolithus floridanus 2 NP15 NN6 n Serravallian
8B Chalks Discoaster deflandrei 2 NP13 NN7 Ypresian Tortonian 32 22,8 Rupelian? Burdigalian? Chalks
Sphenolithus belemnos (??) 2 NN2 NN4 Burdigalian Burdigalian
Sphenolithus moriformis 2 NP5 NN10 Selandian Tortonian
Sphenolithus moriformis 2 NP5 NN10 Selandian Tortonian
Cyclicargolithus floridanus 2 NP15 NN6 n Serravallian
9A Chalks Coccolithus abisectus 2 NP23 NN1 Rupelian Aquitanian? 32 22,8 Rupelian Aquitanian Chalks
Discoaster deflandrei 2 NP13 NN7 Ypresian Tortonian
Coccolithus eopelagicus 2 NP14 NP23 Ypresian Rupelian
Globorotalia kugleri 3 N4 N4 Aquitanian Aquitanian
Globigerinoides trilobus 3 N4b Extant Aquitanian Present
9B VdB83 Between Jacmel and Trois Palmiste Pelagic chalky limestones Triquetrorhabdulus carinatus 2 NP25 NN2 n Burdigalian 23 22,8 Aquitanian Aquitanian Pelagic chalky limestones
Sphenolithus conicus 2 NP25 NN3 Chattian Burdigalian
Coccolithus abisectus 2 NP23 NN1 Rupelian Aquitanian?
Discoaster deflandrei 2 NP13 NN7 Ypresian Tortonian
Sphenolithus heteromorphus 2 NN4 NN6? Burdigalian Serravallian
Marly limestones with interbedded conglomerates; bathyal open
9C Marly limestones with interbedded conglomerates Cyclicargolithus floridanus 2 NP15 NN6 n Serravallian 18 14,8 Burdigalian Burdigalian
marine environment
Helicosphaera ampliaperta 2 NN2 NN4 Burdigalian Langhian
pseudoumbilica 2 NN4 NN15 Burdigalian Zanclean
Globigerinoides subquadratus 3 N4b N15 Aquitanian Tortonian
Globigerinita unicavus 3 P14 N7 Bartonian Burdigalian
10A Chalks with marly 22 16,4 Aquitanian Burdigalian Chalks with marly
Globorotalia mayeri 3 P22b N13 Chattian Serravallian
Globoquadrina a 3 N4b N21a Aquitanian Piacenzian
Globigerinoides trilobus 3 N4b Extant Aquitanian Present
Globigerinoides subquadratus 3 N4b N15 Aquitanian Tortonian
Globigerinoides obliquus 3 N5b PT1a Burdigalian Pleistocene
Globoquadrina a 3 N4b N21a Aquitanian Piacenzian
10B Limestone and marls 17,4 14,3 Burdigalian Langhian Limestone and marls
Globorotalia mayeri 3 P22b N13 Chattian Serravallian
VdB83 Between Tomasin and Fermate Globorotalia peripheroronda 1 M3 M12 Burdigalian Tortonian
Globigerinoides diminutus 1&3 N7 N9 Burdigalian Langhian
Globigerinoides bisphericus 3 N7 N9 Burdigalian Langhian
Orbulina bilobata 3 N9 Extant Langhian Present
Orbulina universa 1 N9 Extant Langhian Present
Globorotalia fohsi 1&3 N9? N12 Langhian Serravallian
10C Mudstones (marly limestones) Globorotalia peripheroronda 1 M3 M12 Burdigalian Tortonian 15,2 11,8 Langhian Serravallian Mudstones (marly limestones)
Globigerina druryi 1 N7 N15 Burdigalian Tortonian
Globigerinoides subquadratus 3 N4b N15 Aquitanian Tortonian
Sphaeroidinellopsis seminulina 1 N7 Pl3 Burdigalian Piacenzian
Sphaeroidinellopsis seminulina 1 N7 Pl3 Burdigalian Piacenzian
Globigerina nepenthes 1 M11 PL1 Serravallian Zanclean
Globoquadrina a 3 N4b N21a Aquitanian Piacenzian
Rivière Grise and Rivière Helicosphaera carteri 2 NN1 Extant Aquitanian Present
11 VdB83 Marly limestones with interbedded conglomerates 11,8 8,4 Serravallian Tortonian Marly limestones with interbedded conglomerates
Redoute Cyclococcolithus macintyrei 2 NN3? NN19 Burdigalian? Calabrian
pseudoumbilica 2 NN2 NN15 Aquitanian? Zanclean
Sphenolithus abies 2 NN5 NN16 Serravallian Gelasian
coalitus 2 NN7 NN10 Serravallian Tortonian
Globorotalia menardii 1 N12 Extant Serravallian Present
Orbulina suturalis 1 M6 N21 Langhian Gelasian
Helicosphaera carteri 2 NN1 Extant Aquitanian Present
Sphenolithus abies 2 NN5? NN16 Serravallian Gelasian
12 VdB83 Rivière Grise; N of Morne Desgourdes Marly and chalky limestones Sphenolithus heteromorphus 2 NN4 NN6? Burdigalian Serravallian 13,5 11,8 Serravallian Serravallian Marly and chalky limestones
Coccolithus abisectus 2 NP23 NN6 Rupelian Serravallian?
Coccolithus floridanus 2 NP15 NN6 n Serravallian
Discoaster exilis 2 NN4 NN9 Burdigalian Tortonian
Discoaster brouweri 2 NN6? NN18 Serravallian? Gelasian

137
Chapter 4: Geology of the Southern Peninsula

Lepidocyclina sp. 4 P10 N13 Lutetian Serravallian


Discocyclina sp. 4 SBZ3 SBZ20 Selandian Priabonian
13A Biomicrite (337-339) 37 28 Priabonian Rupelian?
Fabiania sp. 4 SBZ4 SBZ20 Thanetian Priabonian
Nummulites sp. 4 P3 P21a Selandian Rupelian
Lepidocyclina sp. 4 P10 N13 Lutetian Serravallian
Fabiania sp. 4 SBZ4 SBZ20 Thanetian Priabonian
Discocyclina sp. 4 SBZ3 SBZ20 Selandian Priabonian
13B Biomicrite (340) 28 20 ? Aquitanian?
Dictyoconus sp. 4 >P0 P22 >Danian Chattian
Asterocyclina 4 SBZ3 SBZ20 Selandian Priabonian
Globoquadrina a 3 N4b N21a Aquitanian Piacenzian
Praeorbulina sp. 3 N8 N10 Langhian Langhian
VdB83 Rivière Blanche valley Conformable sequence from Eocene to Miocene
Globorotalia fohsi 1&3 N9? N12 Langhian Serravallian
13C Chalky limestones (341) 15 13,8 Langhian Langhian
Globorotalia mayeri 3 P22b N13 Chattian Serravallian
Globigerinoides trilobus 3 N4b Extant Aquitanian Present
Globorotalia mayeri 3 P22b N13 Chattian Serravallian
Globorotalia fohsi 1&3 N9? N12 Langhian Serravallian
Sphaeroidinellopsis seminulina 1 N7 Pl3 Burdigalian Piacenzian
Globigerinoides subquadratus 3 N4b N15 Aquitanian Tortonian
13D Marly limestones 13,4 11,6 Serravallian Serravallian
Globigerinoides obliquus 3 N5b PT1a Burdigalian Pleistocene
Globorotalia menardii 1 N12 Extant Serravallian Present
Sphaeroidinellopsis seminulina 1 N7 Pl3 Burdigalian Piacenzian
Globigerina nepenthes 1 M11 PL1 Serravallian Zanclean
Cyclicargolithus floridanus 2 NP15 NN6 n Serravallian
Sphenolithus moriformis 2 NP5 NN10 Selandian Tortonian
14A Siliceous biomicrites 46,2 29,6 n Rupelian Siliceous biomicrites
VdB83 Rivière Fond Parisien Lanternithus minutus 2 NP14 NP23 Lutetian Rupelian
Globorotalia cerroazulensis (reworked) 1 E7? E16 n Priabonian
XB Siliceous limestones no reliable fauna Oligocene? Miocene? Siliceous limestones
Globorotalia menardii 1 N12 Extant Serravallian Present
Globoquadrina a 3 N4b N21a Aquitanian Piacenzian
14B VdB83 Rivière Fond Parisien Marly limestones 13,4 9,7 Serravallian Tortonian Marly limestones
Globigerinoides obliquus 3 N5b PT1a Burdigalian Pleistocene
Globigerinoides subquadratus 3 N4b N15 Aquitanian Tortonian
Sphenolithus moriformis 2 NP5 NN10 Selandian Tortonian
Coccolithus abisectus 2 NP23 NN6 Rupelian Serravallian?
Coccolithus floridanus 2 NP15 NN6 n Serravallian
Pedernales River, Mare Jauffrey
15 VdB83 Chalky limestones with cherts Helicosphaera s 2 NP18 NN4 Priabonian Langhian 32 22,8 Rupelian Aquitanian Chalky limestones with cherts
village
Sphenolithus belemnos 2 ?? NN4 ? Burdigalian
Ericsonia fenestrata 2 NP12 NN1 Ypresian Aquitanian
Discoaster deflandrei 2 NP13 NN7 Ypresian Tortonian
Sphenolithus moriformis 2 NP5 NN10 Selandian Tortonian
Sphenolithus distentus 2 NP23 NP24 Rupelian
16 VdB83 Close to Banane village Pelagic limestones with cherts Cyclicargolithus floridanus 2 NP15 NN6 n Serravallian 32 26,8 Rupelian n Pelagic limestones with cherts
Discoaster deflandrei 2 NP13 NN7 Ypresian Tortonian
Ericsonia fenestrata 2 NP12 NN1 Ypresian Aquitanian
Sphenolithus heteromorphus 2 NN4 NN6? Burdigalian Serravallian
Helicosphaera kamptneri 2 NN1 Extant Aquitanian Present
Helicosphaera s 2 NP18 NN4 Priabonian Langhian
pseudoumbilica 2 NN2 NN15 Aquitanian? Zanclean
17 Am97 Lake north of La Borde town Marls 17,8 15 Burdigalian Langhian Marls
Discoaster exilis 2 NN4 NN9 Burdigalian Tortonian
Cyclococcolithus macintyrei 2 NN3? NN19 Burdigalian? Calabrian
Triquetrorhabdulus carinatus (rewrkd?) 2 NP25 NN2 n Burdigalian
Dictyococcites bisectus (reworked?) 2 NP17 NN1 Bartonian Aquitanian
Triquetrorhabdulus carinatus 2 NP25 NN2 n Burdigalian
Around the town of Galais; 8 km NNW Cyclicargolithus floridanus 2 NP15 NN6 n Serravallian
18 Am97 Marls with 20 17,3 Burdigalian Burdigalian Marls with
of Les Cayes Discoaster deflandrei 2 NP13 NN7 Ypresian Tortonian
Sphenolithus belemnos 2 NN2 NN4 Burdigalian Burdigalian
19 Am97 2km south of Carrefour Masson Chalks Globigerapsis semiinvoluta 1 E14 E15 Priabonian Priabonian 38,3 34,6 Priabonian Priabonian Pelagic chalks
Globigerina praebulloides 3 P13 N17a Bartonian Tortonian
20 Am97 Ca Dace; S of Port Salut Marls Globorotalia lobata 1 M9a M9b Serravallian Serravallian 13,4 11,9 Serravallian Serravallian Pelagic marls
Orbulina universa 1 N9 Extant Langhian Present
Helicosphaera carteri 2 NN1 Extant Aquitanian Present
Morne Desravines; 3 km SE of Port-a- Discoaster surculus 2 NN6 NN16 Serravallian? Gelasian
21 Am97 Marls 13,5 3,6 Serravallian? Zanclean Marls unconformably overlying CLIP basalts
Piment pseudoumbilica 2 NN2 NN15 Aquitanian? Zanclean
Sphenolithus abies 2 NN5? NN16 Serravallian Gelasian
Orbulina universa 1 N9 Extant Langhian Present
Morne Rouge; between Sandy clays with biodetrital unconformably
22 Am97 Sandy clays with biodetrital Orbulina bilobata 3 N9 Extant Langhian Present 14,2 13,2 Langhian Serravallian
Chardonnières and Port-à-Piment overlying CLIP basalts
Globorotalia peripheroacuta 1&3 M7 M9a Langhian Serravallian

138
Chapter 4: Geology of the Southern Peninsula

Helicosphaera carteri 2 NN1 Extant Aquitanian Present


Discoaster surculus 2 NN6 NN16 Serravallian? Gelasian Marls and sandy limestones with basal conglomerate containing
23 Am97 Haut Fort; 4 km SE of Tiburon Marls and sandy limestones 13,5 3,6 Serravallian? Zanclean
pseudoumbilica 2 NN2 NN15 Aquitanian? Zanclean basalt clasts, unconformably overlying CLIP basalts
Sphenolithus abies 2 NN5? NN16 Serravallian Gelasian
Sphenolithus heteromorphus 2 NN4 NN6? Burdigalian Serravallian
Around Chambellan; 6 km E of Dame
24 Am97 Marls with intercalated calcarenites Helicosphaera kamptneri 2 NN1 Extant Aquitanian Present 17,8 11,9 Burdigalian Serravallian Marls in faulted contact with Eocene limestones
Marie
pseudoumbilica 2 NN2 NN15 Aquitanian? Zanclean
Catapsydrax dissimilis 1 E14 N6 Priabonian Burdigalian
25A Chalky limestones with cherts 37,8 17,5 Priabonian Burdigalian
Globigerina praebulloides 3 P13 N17a Bartonian Tortonian
Morne Beaumont region; 5 km SSW Globigerina praebulloides 3 P13 N17a Bartonian Tortonian Chalky limestones with cherts overlain (unconformable?) by
25B Am97 Marls; matrix of conglomerates 37,8 17,5 Priabonian Burdigalian
of Jeremie Globigerina venezuelana 1&3 P10 N19 n Zanclean marls with a basal conglomerate
Globigerina praebulloides 3 P13 N17a Bartonian Tortonian
25C Marls with intercalated calcarenites 15,2 17,5 Langhian Tortonian
Orbulina universa 1 N9 Extant Langhian Present
Globorotalia margaritae 1 M14 PL2 Messinian Zanclean
Globigerinoides trilobus 3 N4b Extant Aquitanian Present
26 Am97 Gélin; 8 km SW of Jeremie Marls; matrix of conglomerates 6 3,8 Messinian Zanclean Conglomerates overlying marls (?)
Globigerinoides sacculifer 1 N4b Extant Aquitanian Present
Orbulina universa 1 N9 Extant Langhian Present
Sphenolithus heteromorphus 2 NN4 NN6? Burdigalian Serravallian Marls and limestones overlying lacustrine deposits with basal
27 RM86 North of l'Asile Marls with limestones Helicosphaera kamptneri 2 NN1 Extant Aquitanian Present 17,8 11,9 Burdigalian Serravallian? conglomerate in turn unconformably overlying Cretaceous (or
Discoaster exilis 2 NN4 NN9 Burdigalian Tortonian Paleocene?)
Globigerina senni 1 E5 E13 Ypresian Bartonian
Along Rivière Serpente; between Globigerinatheka barri 1 E10 E15 n Priabonian Eocene (brecciated) limestones unconformably overlain by marls
28 RM86 Limestones 42 38,2 Bartonian Bartonian
Fonds des Negres and l'Asile Globigerapsis kugleri 1 E9 E13 Lutetian Bartonian (Miocene?)
Orbulinoides beckmanni 1 E11 E13 Bartonian Bartonian
Helicosphaera kamptneri 2 NN1 Extant Aquitanian Present
pseudoumbilica 2 NN2 NN15 Aquitanian? Zanclean
Cyclococcolithus macintyrei 2 NN3? NN19 Burdigalian? Calabrian
Between Lhomond and Bélivert; 8 km Marls, contains some labradorite (i.e. CLIP-derived erosional
29 RM86 Marls Coccolithus leptoporus 2 NN2 Extant Aquitanian Present 17,8 11,9 Burdigalian Serravallian?
NE of Fond des Blancs products)
Discoaster exilis 2 NN4 NN9 Burdigalian Tortonian
Helicosphaera kamptneri 2 NN1 Extant Aquitanian Present
Sphenolithus heteromorphus 2 NN4 NN6? Burdigalian Serravallian
Globorotalia obesa 3 P21b Extant Chattian Present
Catapsydrax dissimilis 1 E13 N6 Bartonian Burdigalian
Aquitanian (cf.
30A Limestones with large benthic forams Globigerina praebulloides 3 P13 N17a Bartonian Tortonian 22 >21.2 Aquitanian
overlying)
Globigerinoides trilobus 3 N4b Extant Aquitanian Present
Miogypsina sp. 4 P22 N13 Chattian Serravallian
Globigerinoides trilobus 3 N4b Extant Aquitanian Present
Globorotalia kugleri 3 N4 N4 Aquitanian Aquitanian
30B RM86 Platon Besace - Fort Gari Pelagic limestones 22 21,2 Aquitanian Aquitanian Pelagic limestones unconformably overlying limestones
Catapsydrax dissimilis 1 E13 N6 Bartonian Burdigalian
Globigerina praebulloides 3 P13 N17a Bartonian Tortonian
Globigerinoides trilobus 3 N4b Extant Aquitanian Present
30C Calcarenites Praeorbulina glomerosa 1 N8 N9 Burdigalian Langhian 16,3 14,3 Burdigalian Langhian
Globugerina (reworked?) 1 E13 P22 Bartonian
Globorotalia praemenardii 3 N11b N13a Langhian Serravallian
30D Detrital limestones 13,8 11,6 Langhian Serravallian
Orbulina universa 1 N9 Extant Langhian Present

139
Chapter 4: Geology of the Southern Peninsula

Appendix 3

Google Earth Adobe Illustrator

• Literature •
• Fieldwork • Structural
DATA
• Satellite • Contours
• AsterDEM • Waterways

Google Earth

BME Comparison Our work

140
Chapter 4: Geology of the Southern Peninsula

Appendix 5
Western Massif de la Hotte Depositional Max.
Lithology Reference Reference
Name Age environment thickness
Turonian - Tholeiitic flood basalts intercalated with limestones, Bourgueil et al., 1988;
Cretaceous CLIP Open marine Calmus, 1983; Amilcar, 1997 1500m
Campanian radiolarites, shales and turbiditic sands Mauffret and Leroy, 1997
Sant. - Camp. - Well stratified pelagic limestones with cherts, minor marly Maurrasse, 1982; Calmus, Calmus, 1983, Bourgueil et
Macaya Fm. Bathyal 1000m
m. Maast. limestones 1983; Bourgueil et al., 1988 al., 1988
Basal conglomerate, clay-, silt,- and sandstones, volcaniclastic Shallow marine to
Rivière Glace Fm. Danian Calmus, 1983 150m Calmus, 1983
intervals, interbedded dacites slope
Outer
Upper Paleocene Late Paleocene Silty calcarenite, marls, minor sandstone and conglomerate Calmus, 1983 200m Calmus, 1983
platform/slope
Massive to dm-stratified limestones with large benthic Calmus, 1983; Bourgueil et Calmus, 1983; Bourgueil et
Lower Eocene Early Eocene Platform 200m
foraminifera al., 1988 al., 1988
Middle to late Outer Calmus, 1983; Bourgueil et Calmus, 1983; Bourgueil et
M. - u. Eocene Pelagic limestones with cherts 300m
Eocene platform/slope al., 1988 al., 1988
Oligocene - Maurrasse, 1982; Calmus,
Jeremie Fm. Chalky pelagic limestones with cherts Slope/bathyal Maurrasse 1982 400m
Aquitanian 1983, Amilcar, 1997
Langhian - Transgressive to Calmus, 1983; Bizon et al., Calmus, 1983; Bizon et al.,
Miocene Basal conglomerate, marls and chalky limestones >300m
Messinian bathyal 1985; Amilcar, 1997 1985; Amilcar, 1997
L'Asile region
Calmus, 1983; Bien-Aime
Albian - Tholeiitic flood basalts intercalated with limestones, Platform (Base) to Bourgueil et al., 1988;
Cretaceous CLIP Momplaisir, 1986; Amilcar, 1500m
Campanian radiolarites, shales and turbiditic sands open marine (Top) Mauffret and Leroy, 1997
1997
Campanian - Bien-Aime Momplaisir, 1986; Calmus, 1983, Bourgueil et
Macaya Fm. Well stratified pelagic limestones with cherts Bathyal 1000m ?
Maastrichtian Amilcar, 1997 al., 1988
Rivière Glace Fm. U. Maast. (?) - Basal conglomerate, clay-, silt,- and sandstones, volcaniclastic Shallow marine to
Amilcar, 1997 300m ? Amilcar, 1997
(south/central) Danian intervals, basalts slope
Baradères Fm. Danian -
Neritic limestones with polyps and algae overlain by basalts Outer platform Calmus, 1983 100m Calmus, 1983
(north) Selandian
Selandian - Outer Bourgueil et al., 1988;
Upper Paleocene Siltstones, silty limestones, calcarenites Amilcar, 1997 100 - 250m
Thanethian platform/slope Amilcar, 1997
Massive to dm-stratified limestones with large benthic Calmus, 1983; Bourgueil et Calmus, 1983; Bourgueil et
Lower Eocene Early Eocene Platform 200m?
foraminifera al., 1988 al., 1988
M. - Eoc. - Middle to late Outer platform Calmus, 1983; Bourgueil et
Pelagic limestones with cherts 500m Calmus, 1983
Aquitanian Eocene /slope/bathyal al., 1988
Burdigalian - Transgressive, Woodring et al., 1924; Bien-
Miocene Basal conglomerate, lacustrine deposits, marine marls >300m Bien-Aime Momplaisir, 1986
Messinian lagoonal to bathyal Aime Momplaisir, 1986

142
Chapter 4: Geology of the Southern Peninsula

Massif de la Selle Depositional Max.


Name Age Lithology environment Reference thickness Reference
Sayeed et al., 1978; Van den Desreumaux, 1987 and
Aptian - Tholeiitic flood basalts intercalated with limestones, Platform (Base) to 1500 -
Dumisseau Fm. Berghe, 1983; Bellon et al., Maurrasse et al., 1979; in
Campanian radiolarites, shales and turbiditic sands open marine (Top) 2000m
1985; Sinton et al., 1998 Bourgueil et al., 1988
Campanian - Van den Berghe, 1983; Van den Berghe, 1983;
Macaya Fm. Well stratified pelagic limestones with cherts Bathyal <500m
Maastrichtian Desreumaux, 1985 Desreumaux, 1985
U. Maast. - Clay-, silt-, and sandstones, fragments of CLIP basalts and Mercier de Lépinay et al., Van den Berghe, 1983;
Marigot Fm. Slope 150m
Danian Macaya Fm. 1979; Van den Berghe, 1983 Bourgueil., 1988
Outer Van den Berghe, 1983; Van den Berghe, 1983;
Upper Paleocene Late Paleocene Silty limestones 200m
platform/slope? Bourgueil et al., 1988 Bourgueil., 1988
Massive to dm-stratified limestones with large benthic Van den Berghe, 1983; Van den Berghe, 1983;
Lower Eocene Early Eocene Platform 200m?
foraminifera Bourgueil et al., 1988 Bourgueil., 1988
M. Eoc. - Outer Van den Berghe, 1983;
M. Eoc. - Oligocene Chalky limestones with occasional cherts 400m Van den Berghe, 1983
Oligocene platform/slope? Bourgueil et al., 1988
Aquitanian - Aquitanian -
White chalky mudstones Bathyal Van den Berghe, 1983 <700m Van den Berghe, 1983
Tortonian Tortonian
(Late) Mess. - Calcareous marls and sandstones, labradorite fragments from
Rivière Grise Fm. Flysch, regressive Bourgueil et al., 1988 <800m Bourgueil et al., 1988
Zanclean CLIP erosion
Conglomerates and sandstones, interbedded mollusk-rich Lagoonal/continent Jean Poix, 1980, in Bourgueil
Morne Delmas Fm. Piacenzian <800m Bourgueil et al., 1988
limestones al et al., 1988; Maurrasse, 1982

Cul-de-Sac-1 (Atlantic) well (TD = 2408m)


M. - U. Eocene M. - l. Eocene Limestones with occasional cherts n/a Bourgueil et al., 1988 250m Bourgueil et al., 1988
Oligocene Oligocene Limestones with occasional cherts n/a Bourgueil et al., 1988 150m Bourgueil et al., 1988
Lower Miocene Early Miocene Cherty limestones n/a Bourgueil et al., 1988 500m Bourgueil et al., 1988
Limestones with occasional cherts (lower 150m); argillaceous
Middle Miocene Middle Miocene n/a Bourgueil et al., 1988 400m Bourgueil et al., 1988
limestones (upper 250m)
Silty and marly limestones (lower 300m); silts and clays (upper
Upper Miocene Late Miocene n/a Bourgueil et al., 1988 700m Bourgueil et al., 1988
400m)
Sands, silts and clays (lower 250m); conglomerates and sands
Pliocene Pliocene n/a Bourgueil et al., 1988 400m Bourgueil et al., 1988
(upper 150m)

143
Chapter 4: Geology of the Southern Peninsula

Bibliography
Amilcar, H., 1997: Etude géologique du Massif de La Hotte occidental, Presqu’île du Sud d’Haïti (Grandes Antilles). Implications géodynamiques: 188 p.
Bellon, H., Vila, J.M., and Mercier de Lépinay, B., 1985: Chronologie 40K - 40Ar et affinites geochemique des manifestations magmatiques au Cretace et au Paleogene dans l’ile d’Hispaniola. Symp.
geodyn. Caraïbes, Paris, Ed. Technip,, 329–340.
Van den Berghe, B., 1983: Evolution sedimentaire et structurale depuis le Paleocene du secteur “Massif de la Selle” (Haïti) - “Bahoruco” (République Dominicaine) au nord de la ride de Beata, dans
l’orogene nord Caraïbe (Hispaniola-Grandes Antilles): 305 p.
Bien-Aime Momplaisir, R., 1986: Contribution à l’étude géologique de la partie orientale du Massif de la Hotte (Presqu’île du Sud d’Haïti); Synthèse structurale des marges de la Presqu’île à partir des
données sismiques: 261 p.
Bizon, G., Bizon, J.J., Calmus, T., Muller, C., and Van den Berghe, B., 1985: Stratigraphie du Tertiaire du sud d’Hispaniola (Grandes Antilles). Influence de la tectonique déchrochante sur la
paléogéographie et l’histoire sédimentaire. Symp. Geodyn. Caraïbes, Paris, Ed. Technip,, 371–380.
Bourgueil, B., Andreieff, P., Lasnier, J., Gonnard, R., Le Métour, J., and Rançon, J.-P., 1988: Synthèse géologique de la République d’Haïti, in Technical report- ATN/SF 2506 HA, Port-au-Prince, Bureau
des Mines et de l’Énergie d’Haïti, 332.
Calmus, T., 1983: Contribution à l’étude géologique du Massif de Macaya (Sud-ouest d’Haïti, Grandes Antilles). Sa place dans l’évolution de l’orogène Nord-Caraibe: Universite Pierre et Marie Curie,
222 p.
Desreumaux, C., 1987: Contribution a l’etude de l’histoire geologique des regions centrale et meridional d’Haïti (Grandes Antilles) du Cretace à l’actuel: 507 p.
Jean Poix, C., 1980: Bilan des formations d’Haïti. 1 ère Coll. Géologie d’Haïti, Impr. Le Natal, P-au-P,, 171–183.
Mauffret, A., and Leroy, S., 1997: Seismic stratigraphy and structure of the Caribbean igneous province. Tectonophysics, 283, 61–104.
Maurrasse, F., Husler, J., Georges, G., Schmitt, R., and Damond, P., 1979: Upraised Caribbean sea-floor below acoustic reflector B" at the Southern Peninsula of Haïti. Geologie en Mijnbouw, 58 (1),
71–83.
Maurrasse, F., 1982: Survey of the geology of Haïti: Guide to the field excursions in Haïti. Miami Geological Society,, 103.
Mercier de Lépinay, B., Labesse, B., Sigal, J., and Vila, J.M., 1979: Sedimentation chaotique et tectonique tangentielle Maestrichtienne dans la Presqu’île du Sud d’Haïti (île d’Hispaniola, Grandes
Antilles). C.R. Ac. Sc. Paris, 289 (D), 887–890.
Sayeed, V., Maurrasse, F., Husler, J., and Schmitt, R., 1978: Geochemistry and petrology of some mafic rocks from Dumisseau, southern Haiti. EOS Transactions of the American Geophysical Union, 59
(4), 403.
Sinton, C.W., Duncan, R.A., Storey, M., Lewis, J., and Estrada, J.J., 1998: An oceanic flood basalt province within the Caribbean plate. Earth and Planetary Science Letters, 155, 221–235.

Woodring, W.P., Brown, J.S., and Burbank, W.S., 1924: Geology of the Republic of Haïti. Geological Survey of the Republic of Haïti,, 631.

144
Chapter 4: Geology of the Southern Peninsula

mechanism involves a change of stress perpendicular to the systematic joints after their creation.
Because the creation of open joints will release crack-normal tensile stresses (Pollard and Segall, 1987),
it is possible that the least compressive stress will change from being perpendicular to the principal
joints to being parallel, resulting in the creation of cross joints by biaxial extension (Ghosh, 1988; Simon
et al., 1988). On a regional scale, strain relaxation results from the release of contractional strain
acquired during creation of the systematic joints, which will promote orthogonal cross joints once the
initial stresses are released (Rives et al., 1994). A process responsible for this release of stress could be
uplift and erosion, following deformation and creation of systematic joints at depth during burial
(Engelder, 1985). Stress rotation requires a 90° rotation of regional stress after the formation of the
systematic joints, to explain the creation of orthogonal cross joints (i.e. Bai et al 2002). This rotation
can result from regional fluctuations of tectonic stresses (Bahat and Grossmann, 1988; Dunne and
North, 1990), or from regional uplift and erosion (Friedman, 1972).

The similarity in orientation of the joints implies that they resulted from a similar stress regime. The
youngest lithologies in which these joints are found are middle Miocene in age, which indicates that
the joints formed after the middle Miocene. This timing corresponds to the period of maximum burial
of the formations and initiation of uplift of the Southern Peninsula during late Miocene times (section
4.2). In this hypothesis, the systematic joints probably developed in response to layer-parallel
shortening in the late Miocene, while the cross joints develop by the release of stress by strain
relaxation (Rives et al., 1994) during initiation of uplift (Engelder, 1985). Unfortunately no systematic
cross-cutting relationships could be established for the joints, thus it is unclear which of the two trends
represent the systematic joints and which the cross joints. Alternatively, the orthogonal joint sets could
result from stress release after creation of the systematic joints during burial (Pollard and Segall, 1987;
Ghosh, 1988; Simon et al., 1988).

147
Chapter 4: Geology of the Southern Peninsula

4.5 Shortening of the Southern Peninsula


This section aims at quantifying the amount of shortening on the Southern Peninsula from the Eocene
onwards by line-length restoration of a key horizon as indicated on the cross sections presented in
section 4.2 and figure Figure 4 - 4. In Chapter 7 the results are compared with post-Eocene shortening
in Haiti and the southern Dominican Republic.

Line-length restoration of cross sections works best when; 1) cross sections are parallel to the direction
of shortening, 2) no out-of-plane transport of material occurs, and 3) bedding length is preserved
during deformation (Woodward et al., 1989). These three conditions are not perfectly met on the
Southern Peninsula, but their relative impact is expected to be minor, as discussed below.

The cross sections in section 4.2 are constructed perpendicular to the dominant east – west structural
trend and thus parallel to the shortening direction. Total amount of shortening is underestimated due
to out-of-plane deformation resulting from the change in trend of fold axes along-strike of the cross
section. Fortunately the trend of the fold axes mainly varies between WNW- to W-striking (section
4.2), and the effect of this on the measured amount of shortening is thus only small. Out-of-plane
transport of material is evident with E – W trending strike-slip faults along N – S cross sections. The
EPGFZ is expected to be sub-vertical in most of the western and central Southern Peninsula (section
4.2). Vertical displacement on a vertical fault, when viewed parallel to strike, does not contribute to
any horizontal shortening perpendicular to strike. The cross sections are almost perpendicular to the
trend of the EPGFZ, which limits the under- or over-estimation of shortening along these section due
to vertical displacements along the EPGFZ. The last point of concern is the conservation of line-length
during deformation. Displacement on layer-parallel decollement levels results in a reduction of line-
length during deformation, by decoupling the overlying from the underlying sequence without markers
to track its absolute displacement. Most of the deformation will however be accommodated within
the overlying sequence. Decollement levels are mainly found in Eocene cherty limestones. By restoring
the line-length of a formation above the Eocene decollement levels this problem is, at least partly,
circumvented. Even with the above uncertainties taken into account, line-length restoration provides
a good first-order estimate of the minimum total amount of shortening (Woodward et al., 1989).

An additional complication to the conservation of bedding length during deformation is erosion, which
removes the intended restoration line. The Southern Peninsula is affected by three phases of erosion;
1) latest Cretaceous – early Paleocene, 2) early Miocene and 3) latest Miocene to present-day, with
periods of relative tectonic quiescence in between. Early Miocene erosion on the Southern Peninsula

148
Chapter 4: Geology of the Southern Peninsula

The amount of shortening in the Massif de la Selle south of Port-au-Prince (Figure 4 - 4a) is calculated
using the cross section in Figure 4 - 4b. Because early Miocene deformation did not affect this area,
and to allow a comparison with shortening in the Chaîne des Matheux, the line-length of the top
Oligocene is restored. This is above the decollement interval of Eocene cherty limestones. Shortening
is calculated for the entire section, and to the north and south of the anticlinal axis separately. The
latter is done to allow an easier comparison with published cross sections in chapter 7, which are
mainly constructed on the northern flank of the Massif de la Selle – Sierra de Bahoruco anticline, and
because the northern flank appears more deformed than the southern flank (Figure 4 - 4b). The
continuation of the now eroded top Oligocene is interpreted using the average thickness of eroded
formations and folding of the Paleocene sequence, shown by the grey line in Figure 4 - 4b. The present-
day horizontal length of the top of the Oligocene between the two pin lines in Figure 4 - 4b is 33.1 km.
The restored line-length of the top Oligocene is 35.0 km. North – south shortening of the Massif de la
Selle south of Port-au-Prince is 1.9 km or 5.4%. By the same exercise, shortening of the southern Massif
de la Selle is 0.2 km or 1.1%, while the northern flank is shortened by 1.7 km or 10.2%.

Shortening in the l’Asile region (Figure 4 - 4a) is more difficult to constrain. Early Miocene erosion
removed the entire top Oligocene which can thus not be used to restore the line-length (Figure 4 - 4c).
The only viable horizon is the top of the lower Paleocene sequence, even though this is also expected
to be partly eroded by early Miocene erosion in the l’Asile Basin. In the l’Asile Basin only limited
decollement levels are observed in the Eocene formations (section 4.2, appendix 4) and the associated
displacement is probably minor. Line-length restoration is performed in the upper Paleocene,
represented by the grey line, between the two pin lines (Figure 4 - 4c). Present-day horizontal length
is 17.2 km. The restored line-length of the top lower Paleocene is 18.1 km. Post-Paleocene north –
south shortening in the central l’Asile area north of the EPGFZ is 0.9 km or 5.0%. As shown in section
4.2, the early Miocene accounts for 1.3% of shortening, while the late Miocene – recent accounts for
the remaining 3.7%. In the western Massif de la Hotte (Figure 4 - 4a) the amount of shortening is poorly
constrained due to strong erosion down to the Cretaceous basalts (Figure 4 - 4d). For this cross section
the top of the lower Paleocene is also chosen to restore the line-length, with the same arguments as
for the l’Asile area. Where a sedimentary cover is lacking the wavelength and amplitude of the folds
and morphology is used to guide the line-length restoration. Present-day length of the top lower
Paleocene along the gray line between the two pin lines is 26.3 km, while the restored length is 27.2
km, resulting in 0.9 km or 3.3% of north – south shortening, without subdivision between the Miocene
deformation phases.

150
Chapter 4: Geology of the Southern Peninsula

4.6 Displacement along the EPGFZ


Determining the displacement along a strike-slip fault is mainly based on correlative markers that cross
the fault zone, which are subsequently offset by fault activity. This technique works best if these
correlative markers; 1) have a discrete, well defined, geographic extent, 2) have a high uniqueness
within the research area, 3) are at a high angle to the fault zone, and 4) are in place prior to the onset
of strike-slip activity. These correlative markers can either be lithological or geomorphological.

Examples of lithological markers include; a) intrusions with a small geographic extent and high
geochemical uniqueness (i.e. monzogranite intrusion, San Andreas Fault (SAF): Frizzell et al., 1986;
Matti and Morton, 1993), b) distinctive fluvial deposits at a high angle to the fault zone (i.e. Mint
Canyon Formation fluvial conglomerates with unique source region, SAF: Ehlig et al., 1975; Ehlert,
2003), c) alluvial fans that cross the fault zone (i.e. Biskra Palms fan, Mission Creek fault, SAF system:
Van der Woerd et al., 2006). Examples of geomorphological markers include; d) incised canyon systems
at high angle to the fault zone (i.e. offshore SOFZ: Leroy et al., 2015) and e) fluvial streams that cross
the fault zone (i.e. Momance section, EPGFZ: Prentice et al., 2010).

Caution should be exercised when interpreting the displacement of correlative markers. Whether or
not a correlative marker records the total displacement along a strike-slip fault depends on; 1) the
onset of fault activity versus the timing of creation of the marker, and 2) the sense of slip along the
fault. If a correlative marker was created during activity of the fault the resulting offset of that marker
will only record part of the total displacement. Lithological units are three dimensional, but are used
in two-dimensional map view as markers to construct offsets. Differential erosion or burial of
lithological markers during activity of the strike-slip fault can influence their map view extent, which
may lead to under- or overestimation of the total displacement. If the slip vector of a strike-slip fault
is not parallel to the Earth’s surface, the offset of markers along the surface trace only records part of
the total displacement. Surface-offsets are therefore best viewed as recording the minimum
displacement of a strike-slip fault.

When correlative markers are used, care should be taken in determining the position of these markers
along the length of the fault, since the amount of displacement varies along the fault’s length;
displacement is zero at the tip of a fault and usually increases to a maximum value near the center of
the fault (Barnett et al., 1987). Strong variations of displacement along the length of mature faults (Kim
and Sanderson, 2005) result from fault growth by the scale-invariant linkage of smaller faults
(Tschalenko, 1970). The segmented nature of a fault system, with different fault strands absorbing

151
Chapter 4: Geology of the Southern Peninsula

some scale-invariant behavior to fault propagation and growth (Walsh and Watterson, 1988; Bonnet
et al., 2001). Larger faults result from growth and linkage of smaller faults (Peacock and Sanderson,
1991; Walsh et al., 2002; Childs et al., 2009), a process that is at least in part responsible for the
observed up to two orders of magnitude scatter in displacement – length ratios of faults (Figure 4 - 5;
Dawers and Anders, 1995; Kim and Sanderson, 2005). Walsh et al. (2002) note that in a developing
fault system, certain faults tend to increase in length compared to others due to segment linkage,
which subsequently focusses displacement on the newly created fault, with concomitant termination
of activity on smaller faults. Fault-length scaling is useful in providing upper limits for fault
displacement only, which can serve as a first-order quality control on structural interpretations and
cross sections at depth (Faulkner et al., 2010).

4.6.1 Previous work on the EPGFZ


The amount of left-lateral offset along the EPGFZ has long been the topic of debate (Calmus, 1983; Van
den Berghe, 1983a; Bourgueil et al., 1988; Mann et al., 1995; Saint-Fleur, 2014).

Calmus (1983) attempts to find both an upper and a lower limit of displacement for the EPGFZ. This
author bases an upper limit of 50 km total offset in the western Massif de la Hotte on the occurrence
of some blocks of Macaya-type limestone within the Eocene formations south of Maniche, and the
distance to a westernmost outcrop of Macaya Formation limestone along-strike north of the EPGFZ.
On the present-day geological map (Figure 4 - 6) the Macaya Formation is present to the north of the
EPGFZ, albeit locally eroded, from Tiburon until the western l’Asile region. There is thus no
westernmost limit to the Macaya Formation north of the EPGFZ, which can therefore also not be used
for constraining the maximum offset along the EPGFZ.

In the western Massif de la Hotte around Les Irois, north of the EPGFZ, Calmus (1983) notes the
presence of dolerite bodies within the Cretaceous CLIP basalts, which he also finds to the south of the
EPGFZ in the outcropping CLIP basalts east of Morne Perdu Temps (Figure 4 - 6). Intrusive dolerite sills
and dikes are however a common occurrence within the CLIP (Donnelly et al., 1973; Révillon et al.,
2000), making it difficult if not impossible to use them as a correlative markers without properly
understanding their geographic extent.

In the western Massif de la Selle, Calmus (1983) and Van den Berghe (1983) propose an offset of 25
km along the EPGFZ (Figure 4 - 6). Their mutual argument is mainly based on a few outcrops of blocks
on the northern side of the EPGFZ, between Miragoane and Petit-Goâve, that visually resemble the
‘Jacmel blocks’ series of Mercier de Lépinay et al. (1979) that are found west of Beloc (Figure 4 - 6;

153
Chapter 4: Geology of the Southern Peninsula

Calmus, 1983; Van den Berghe, 1983). If these outcrops north of the EPGFZ can be proven, they may
provide a good correlative marker since their occurrence appears otherwise limited to the Beloc region
(Mercier de Lépinay et al., 1979). However, neither Van den Berghe (1983) nor Calmus (1983, 1984)
provide any description, stratigraphic position, or a literature reference to the occurrence of these
blocks in the proposed area between Miragoane and Petit-Goâve. Even if their statement is correct,
the amount of offset could be much less, since the western limit of the ‘Jacmel Blocks’ is unknown
(Figure 4 - 6; Mercier de Lépinay et al. (1979).

South of Port-au-Prince a 30 km offset is estimated by Van den Berghe (1983), based on the occurrence
of the middle Miocene oyster Ostrea haitensis in two basins on either side of the EPGFZ. However, as
this author already notes, it cannot be excluded that middle Miocene oyster-rich formations lie buried
under the Cul-de-Sac plain in the north. This appears likely and, moreover, the lack of oyster-rich
middle Miocene on the southern side can easily be explained by erosion (Figure 4 - 6).

None of the lithological markers provided above give an unambiguous and reliable amount of offset
along the EPGFZ. The least unreliable is the maximum of 25km offset of ‘Jacmel Blocks’ on the ridge
between Miragoane and Petit-Goâve to the north of the EPGFZ, and the ‘Jacmel Blocks’ to the west of
Beloc to the south of the EPGFZ, which remains to be proven.

4.6.2 Approach
The lack of unique, geographically well-defined lithological markers hinders the determination of
EPGFZ-related offset. Differential uplift and subsidence on either side of the EPGFZ, combined with
rapid weathering and erosion in the Southern Peninsula’s climate, makes correlating lithological
boundaries across the EPGFZ difficult. Instead of using lithological markers, this section attempts to
quantify the total offset along the active trace of the EPGFZ using displaced river channels.

Offset river channels along the Rivière Momance, south of Port-au-Prince (Figure 4 - 6), are used by
Prentice et al. (2010) to determine m- and ka-scale displacements along the EPGFZ, in a study
conducted in response to the 2010 Mw 7.0 Léogâne earthquake (Bilham, 2010b; Calais et al., 2010;
Hayes et al., 2010), or km- and Myr-scale displacements by (Saint-Fleur, 2014). A similar approach is
used by Leroy et al. (2015) to determine km- and Myr-scale displacements along the active trace of the
SOFZ offshore northern Haiti, using incised valleys up to 800m wide and 600m deep.

In this study, river channel offset is determined from 30m gridded AsterDEM data at the Rivière
Momance – Rivière Froide segment, south of Port-au-Prince, and at the Rivière de Cavaillon, in the
Clonard Basin (Figure 4 - 6).
154
Chapter 4: Geology of the Southern Peninsula

4.6.3 Offset of the Rivière Momance and Rivière Froide


The Rivière Momance and Rivière Froide follow the trace of the EPGFZ on the southern side of a
narrow, elongate ridge to the south of Port-au-Prince (PAP) (Figure 4 - 7e). On the southern side of the
EPGFZ the Rivière Momance flows in a NNW-direction and is incising a valley up to 2000m wide and
700m deep. Its course rapidly changes on the southern side of the PAP-ridge to flow in a WSW-
direction, which at the western end of the ridge changes to NNW and onto the Léogâne delta plain.
The offset between the two main branches of the Rivière Momance is around 15km, with the northern
branch displaced relatively towards the west. This offset is similar to that obtained by (Saint-Fleur,
2014). At the eastern end, the Rivière Momance is separated from the Rivière Froide by a 500m wide,
100m high NW-trending ridge. On the southern side of the EPGFZ the Rivière Froide follows the trace
of the Redoute fault and flows in a NW-direction, which changes sharply to a WSW-direction as it
enters the EPGFZ-related valley. Tributaries of the Rivière Froide that feed into the river along the
EPGFZ flow in a NNW-direction, similar to the tributaries of the Rivière Momance. At the western end
the river bends sharply northwards, roughly where the Gressier fault intersects the ridge, and flows
into the Port-au-Prince Bay. The offset between the two main branches of the Rivière Froide is around
7km, with the northern branch displaced relatively towards the west.

4.6.4 Evolution of the Rivière Momance – Rivière Froide EPGFZ segment


Uplift of the anticlinal central Massif de la Selle is constrained from the late Miocene onwards, with
subaerial erosion taking place from the Messinian onwards (Bourgueil et al., 1988). During this time
the rivers likely flowed northwards on the northern side of the emerging anticline, in a direction away
from the central east-west trending axis (Figure 4 - 7a). The narrow, elongate ridge on the southern
side of Port-au-Prince probably resulted from activity on the EPGFZ, as its southern flank hosts the core
of this fault system (see section 4.2). In northern Hispaniola, the San Francisco Ridge is also postulated
to result in response to strike-slip activity on the SFZ (Winslow et al., 1991). During incipient creation
of the PAP-ridge, the north-flowing Rivière Momance was located directly west of the uplifting ridge
(Figure 4 - 7b). Directly east of the Rivière Momance a watershed diverted drainage from the Massif
de la Selle eastwards along the incipient EPGFZ (Figure 4 - 7b). With progressive offset along this
segment of the EPGFZ, the southern Rivière Momance became progressively displaced to the east
(Figure 4 - 7c). At some time during the activity of the EPGFZ, the Gressier and Redoute faults became
active (see section 4.2 for discussion). The Gressier fault created a northern drainage pathway through
the PAP-ridge for the river flowing along the trace of the EPGFZ (Figure 4 - 7c). This effectively created
and controlled the course of the northern Rivière Froide (Figure 4 - 7c). Activity on the Redoute Fault
controlled drainage in this area and created the southern Rivière Froide (Figure 4 - 7c).

156
Chapter 4: Geology of the Southern Peninsula

Increased displacement on the EPGFZ offset the course of the southern river segments eastwards,
progressively bringing them into their present-day configuration (Figure 4 - 7d and Figure 4 - 7e).

The above scenario explains the difference in offset between the Rivière Momance and the Rivière
Froide, as the thrust faults that constrain the flow of the Rivière Froide originated while the EPGFZ
segment was already active and displacing the Rivière Momance. The offset of the Rivière Momance
thus records the maximum offset along this segment of the EPGFZ, at least since the creation of the
PAP-ridge. Had the offset along this fault segment been 500m larger it would have displaced the small
ridge directly adjacent to the Rivière Momance to the east, and connected it with the northern Rivière
Froide (Figure 4 - 7e).

EPGFZ-parallel GPS velocities are rather constant throughout the Southern Peninsula at 8 mm yr-1
(Calais et al., 2016). These GPS-velocity models constrain the motion between central Haiti and the
southern part of the Southern Peninsula (Calais et al., 2016; Symithe and Calais, 2016). Under the
assumption that all strike-slip displacement is accommodated by offset along the EPGFZ and that the
modeled GPS velocities are representative for short geological timescales, a 15 km displacement for
the Rivière Momance results in an onset of activity at 1.9 Ma for this segment of the EPGFZ. However,
additional strike-slip faults are observed in the Southern Chaîne des Matheux (section 4.2), onshore
the Southern Peninsula (section 6.2), and in the offshore between the Southern Peninsula and Gonâve
Island (Bien-Aime Momplaisir, 1986; Mann et al., 1995). Strike-slip displacement is therefore probably
partitioned over multiple oblique or strike-slip faults, and not only taken up by the EPGFZ. The 1.9 Ma
age for the initiation of activity on the segment south of Port-au-Prince is therefore a minimum age.

4.6.5 Offset of the Rivière de Cavaillon


The Rivière de Cavaillon flows in an eastwards direction along the trace of the EPGFZ, on the northern
side of a narrow, elongate ridge south of Maniche (Figure 4 - 8e). To the north of the EPGFZ, the
northern Rivière de Cavaillon flows in a SSE-direction and is incising a valley up to 2500m wide and
500m deep (Figure 4 - 8e). On the eastern termination of the ridge, the Rivière de Cavaillon abruptly
bends and flows southwards (Figure 4 - 8e). Left-lateral offset of the Rivière de Cavaillon is around 12
km. A restraining bend on the western end of the Clonard Basin controls the ESE-directed flow of the
tributaries into the Rivière de Cavaillon (Figure 4 - 8e). Directly west of this restraining bend the Ravine
du Sud also makes an abrupt 90 degrees turn, from flowing eastwards to southwards (Figure 4 - 8e).
There is however no correlative river to the north of the EPGFZ to link with the Ravine du Sud, making
it impossible to determine the offset of this river.

158
Chapter 4: Geology of the Southern Peninsula

4.6.6 Evolution of the Rivière de Cavaillon segment


The Rivière de Cavaillon is bounded on the south by a ridge that is elongated in the direction of the
EPGFZ (Figure 4 - 8d). Although this ridge has much less topographic elevation, the creation is possibly
also linked with activity of this segment of the EPGFZ, similar to the Port-au-Prince and San Francisco
ridges discussed above. The Rivière de Cavaillon flows on the southern side of the Clonard Basin, which
is interpreted by Mann et al. (1983, 1995) as a Quaternary pull-apart basin related to the EPGFZ. The
evolution of the Rivière de Cavaillon segment follows a similar pattern as the Momance-Froide
segment at Port-au-Prince described above. The uplifted east-west axis of the anticlinal central Massif
de la Hotte resulted in southwards-directed drainage on its southern flank (Figure 4 - 8a). This drainage
became disturbed by the initiation of the pull-apart basin and the relative uplift of the ridge to the
south (Figure 4 - 8b). With increased displacement along this segment of the EPGFZ the Rivière de
Cavaillon became displaced towards the east (Figure 4 - 8c), until it arrived in its present-day
configuration (Figure 4 - 8d).

The Rivière de Cavaillon is offset by about 12 km. Following the same reasoning as for the Rivière
Momance – Rivière Froide segment, this results in a minimum age of initiation of 1.5 Ma.

4.6.7 Discussion

The different segments of the EPGFZ are plotted on a length versus maximum displacement chart
(Figure 4 - 5). Determining the length of a fault segment is dependent on the scale of observation (Kim
and Sanderson, 2005). The length of the entire EPGFZ system, as calculated based on the definition by
Mann et al. (1995) from the Cayman Ridge to the Dominican Republic, is around 1200 km. Taking only
the section from the Jamaican restraining bend to the Dominican Republic results in a length of 500
km. Both these estimates are plotted in Figure 4 - 5, with an averaged displacement on the EPGFZ,
combining displacement on the Momance-Froide and Cavaillon segments, of 14 km.

The EPGFZ is clearly segmented, both offshore in the Jamaica Passage (Corbeau et al., 2016b) and
onshore (Figure 4 - 6) the Southern Peninsula. Determining the exact length of each segment is difficult.
One way of characterizing the length of the Momance-Froide segment south of Port-au-Prince is by
measuring it along the relatively continuous trace from Pétionville in the east to the Tapion Ridge
restraining bend (Prentice et al., 2010) in the west, which results in around 50 km length (Figure 4 - 6).
Another way is by assuming that each small bend represents the start of a different fault segment
(Prentice et al., 2010), in which case the segment south of Port-au-Prince becomes limited from
Pétionville to the Léogâne delta plain, at around 30 km length (Figure 4 - 7). In much the same way the
length of the Rivière de Cavaillon segment can be arbitrarily defined from Dubois in the west (following

160
Chapter 4: Geology of the Southern Peninsula

Calmus, 1983) to Rousseau in the east (following Mann et al., 1995), which results in a length of around
40 km. However, since the Ravine du Sud is not displaced around Camp Perrin, which would be
expected if the EPGFZ continued further west at this location as proposed by Calmus (1983), it appears
more reasonable to limit the western extent of the Rivière de Cavaillon segment to a location just east
of Camp Perrin. This results in a total segment length of around 30 km (Figure 4 - 8).

A 1200 km length for the EPGFZ with 14 km displacements falls outside of the lower bound of global
strike-slip faults, while a 500 km length with the same displacement falls around the lower boundary
(Figure 4 - 5). With a displacement of around 14 km, and using the average trend line of the global
strike-slip dataset, results in an expected segment length in the order of 150 km. However, the strike-
slip system probably consists of a broad zone with multiple on- and offshore strike-slip faults, that
overlap and are adjacent to one-another, all of which accommodate part of the left-lateral
displacement between central Haiti and the Caribbean Plate. For the system as a whole, the total
length is thus probably easier to constrain than the bulk-displacement within the strike-slip zone.
Therefore, assuming that the segment lengths are correct, a 1200 km length gives an expected
displacement of around 150 km, while a length of 500 km results in around 65 km of expected
displacement (Figure 4 - 5).

The length versus displacement of the Rivière Momance – Froide and Rivière de Cavaillon (50 and 30
km length, 15 and 12 km displacement, respectively) fall around the upper bound of global strike-slip
faults (Figure 4 - 5). Assuming that the segment lengths are correct, and using the same global trend
line for strike-slip faults, results in a displacement of 4 and 2.5 km for the Momance-Froide and
Cavaillon segments, respectively (Figure 4 - 5). However, inversely to the strike-slip system as a whole,
the displacement on the individual fault segments is better constrained than their length, as discussed
above. Therefore, assuming a displacement of 15 and 12 km results in a length of ~160 km for the
Momance-Froide segment and ~120 km for the Cavaillon segment (Figure 4 - 5).

A ~160 km length for the Momance-Froide segment can be achieved by assuming that the EPGFZ
continues eastwards under the Cul-de-Sac plain into the Dominican Republic (Figure 4 - 6), as proposed
by Mann et al. (1995). Similarly, a ~120 km length for the Cavaillon segment is possible by assuming a
link with the EPGFZ segment south of L’Asile (Figure 4 - 6).

The relatively young age of strike-slip activity in the Southern Peninsula (around 10 – 7.5 Myrs)
combined with the strong segmented character of the system, could also explain the discrepancy
between the assumed fault system length and measured displacements. In a developing fault system,
due to the linkage of fault segments, total fault system length is often larger than would be expected
based on measured individual fault segment displacements (Kim and Sanderson, 2005).

161
Chapter 4: Geology of the Southern Peninsula

162
Chapter 5: Geology of the Chaîne des Matheux

5.1 Stratigraphy of the Chaîne des Matheux


The pre-Eocene stratigraphy of the Chaîne des Matheux based on published data is presented in
section 5.1.1 Figure 5 - 2, while the general stratigraphy of the eastern Chaîne des Matheux based on
published data is presented in section 5.1.2 and Figure 5 - 3. The correlation between published data
and stratigraphic observations made during field campaigns as part of this study are presented in
section 5.3.

5.1.1 Pre-Eocene stratigraphy


The pre-Eocene stratigraphy of the Chaîne des Matheux outcrops in the Pierre Payen anticline in the
westernmost part of the chain (Figure 5 - 1). The core of this structure exposes three stacked fault-
bounded Cretaceous series (Figure 5 - 2a, Vila et al., 1988), which are, from bottom to top: A) a volcano-
sedimentary sequence of tuffs, greywackes and weathered basaltic flows, interbedded with
biomicrites containing Globotruncana marginata and Globotruncana lapparenti (Santonian; Young et
al., 2018), B) a terrigenous sequence with a flysch-like character, consisting of claystones, calcareous
fine sandstones and m-scale beds of greywackes, the latter containing Globotruncana gansseri and
Globotruncana mayaroensis (Maastrichtian; Young et al., 2018), and C) massive white epipelagic
limestones with chert levels containing Globotruncana arca and Globotruncana calcarata (Campanian;
Young et al., 2018). The latter sequence (C) is also observed in faulted contact overlying the previous
two (A and B) (Figure 5 - 2a, Vila et al., 1988). Unconformably overlying all three Cretaceous formations
are conglomerates containing clasts of massive limestones (C) and tuffs (A), which are in turn
conformably overlain by massive platform limestones containing Ranikothalia bermudezi (Thanetian;
BouDagher-Fadel, 2008). The correct stratigraphic order for the pre-Eocene stratigraphy is shown in
Figure 5 - 2b. These series have been correlated to formations on the Southern Peninsula (Vila et al.,
1988; Pubellier et al., 1991); the Santonian volcano-sedimentary sequence A to the CLIP basalts, the
flysch-like terrigenous sequence B to the Rivière Glace/Beloc/Marigot Formations, and the epipelagic
cherty limestones (C) to the Macaya Formation. Based on this correlation these authors propose that
the basement of the Chaîne des Matheux is similar to that of the Southern Peninsula. However, the
lack of geochemical data on the tuffs and basalts prohibits a direct comparison with the CLIP basalts.
The massive epipelagic limestones also differ from the dm-bedded pelagic limestones of the Macaya
Formation, the first deposited in water depths up to 200m, the latter in a slope or basin environment.

164
Chapter 5: Geology of the Chaîne des Matheux

Alternatively, this sequence of Upper Cretaceous tuffs, greywackes and limestones can be correlated
to the northern island-arc terranes. The description of the Santonian tuffs, greywackes and basalts of
sequence A fits well with the tuffs, basalts, andesites, dacites, rhyolites and intercalated Coniacian to
early Campanian sediments exposed in the Peralta Belt, which are in turn similar to the island-arc
volcanics of the Massif du Nord (Lewis et al., 1991). In the Massif du Nord the volcano-sedimentary
units are overlain by a flysch-like turbiditic sequence containing shale-siltstones with chert levels and
limestone intercalations of Campanian to Maastrichtian age (Bourgueil et al., 1988; Lewis et al., 1991),
which could correspond to sequences C and B. In the absence of geochemical data on the nature of
the tuffs and basalts in the Pierre Payen region, there is no conclusive evidence for assigning either a
CLIP (Southern Peninsula) or island-arc (northern Haiti) origin to these sequences.

A Chaîne des Matheux basement B Chaîne des Matheux basement


(not to scale), Vila et al., 1988 (not to scale), correct stratigraphic order
Massive platform limestones Massive platform limestones
Than. Than.
Conglomerates; clasts of a and c Conglomerates; clasts of a and c
c; Folded massive epipelagic b; Fine sandstones, siltstones,
Camp. Maast.
limestones with chert levels greywackes [Flysch]

b; Fine sandstones, siltstones, c; Folded massive epipelagic


Maast. Camp.
greywackes [Flysch] limestones with chert levels

Sant. Sant.

Fault Unconformity
Figure 5 - 2: A: Present-day stratigraphic order of the pre-Eocene series in the western Chaîne des Matheux,
Pierre Payen anticline, based on the work by Vila et al. (1988). B: Lithologies in the correct stratigraphic order.

165
Chapter 5: Geology of the Chaîne des Matheux

5.1.2 General stratigraphy of the eastern Chaîne des Matheux


The stratigraphy of the northern and southern flanks of the eastern Chaîne des Matheux is similar until
middle Miocene times, after which a diachronous evolution is apparent between the northern flank
(shallower facies) and the southern flank (deeper facies) (Figure 5 - 3; Bourgueil et al., 1988).

The oldest sediments exposed in the eastern Chaîne des Matheux are lower to middle Eocene massive
platform limestones containing large benthic foraminifera and Melobesia algae measuring at least
250m in thickness (Butterlin, 1960; Desreumaux, 1987). These are overlain by 60m of dm-bedded
chalky limestones with cherts of late Eocene age (Butterlin, 1960; Desreumaux, 1987). Middle Eocene
basalts are found in the in the western Chaîne des Matheux (Saint-Marc region) (Perodin Formation,
Boisson and Pubellier, 1987) and the eastern Montagnes du Trou d’Eau (El Aguacate Formation Biju-
Duval et al., 1982; Hernaiz Huerta et al., 2007), but not in the eastern Chaîne des Matheux (Bourgueil
et al., 1988). The existence of the Oligocene in the Chaîne des Matheux is subject to debate (Bourgueil
et al., 1988). Butterlin (1960) and Desreumaux (1987) argue that the Oligocene is represented by dm-
bedded cherty that are in conformable contact with upper Eocene and lower Miocene limestones. This
is disputed by Van den Berghe (1983) who observes Aquitanian limestones containing Globorotalia
kugleri (N4; Hilgen et al., 2012) directly overlying upper Eocene limestones. The Aquitanian consists of
100m of white turbiditic cherty limestones (Van den Berghe, 1983a) that grade upwards into 200m of
Burdigalian basinal chalky limestones with minor cherts (Van den Berghe, 1983a; Desreumaux, 1987).

The diachronous evolution starts in the middle Miocene. On the northern flank, chalky limestone
sedimentation continues through the Langhian and early Serravallian. During the Serravallian
sedimentation progressively becomes less calcareous and increasingly silty. Serravallian (Bourgueil et
al., 1988) silty limestones of the Madame Joie Formation (Woodring, 1922) progressively change to
calcareous siltstones of the Thomonde Formation (Woodring, 1922) during the latest Serravallian
(Desreumaux, 1987). The Langhian and Serravallian interval is approximately 200m thick (Van den
Berghe, 1983a). The Tortonian (N16) retains the same calcareous-silty facies and has a thickness of
roughly 200m (Bourgueil et al., 1988). These facies are capped by an 80m thick Tortonian to Messinian
(N17) reefal limestone bar, which is overlain by 200m of Messinian clastics with marine influences
(Bourgueil et al., 1988). The top of the sequence consists of 900m of azoic sands and conglomerates
(Las Cahobas Formation; Woodring et al., 1924), dated Pliocene to early Pleistocene (N21) in the
western Chaîne des Matheux (Van den Bold, 1981; Maurrasse, 1982).

166
Chapter 5: Geology of the Chaîne des Matheux

On the southern flank and northern border of the Cul-de-Sac plain, chalky limestone sedimentation
continues until the latest Serravallian. The change to silty limestones of the Madame Joie Formation is
recorded some 2.5 Myr later than on the northern flank (Bourgueil et al., 1988). The base of the
overlying calcareous silts of the Thomonde Formation is dated middle Tortonian (N16). Based on
average sedimentation rates the top of this 600m thick azoic series is expected to be in the uppermost
Messinian, while the overlying limestone bar is possibly of latest Messinian or earliest Pliocene age
(Bourgueil et al., 1988). The Thomonde Formation is locally intercalated with basalts on the southern
flank of the Chaîne des Matheux (Boisson, 1987). The coarse sands and conglomerates of the Las
Cahobas Formation are missing on the southern flank, although this formation may be equivalent to
the 725m Pliocene multicolored clays, silts and fossiliferous sands that are overlying a 95m thick reefal
limestone bar in the offshore Cul-de-Sac-1 (Crux) well (Bourgueil et al., 1988).

Quaternary basalts in the Thomazeau and La Vigie (also known as Saut d’Eau) volcanic centers in the
Chaîne des Matheux (Butterlin, 1960) are 1.8 – 1.6 Ma (Thomazeau) and 2.3 Ma (La Vigie) (K-Ar; Wadge
and Wooden, 1982; Pubellier et al., 1991) mafic alkaline basalts (Kamenov et al., 2011b).

5.2 Haitian fold-and-thrust belt


The Haitian fold-and-thrust belt (HFTB) developed in western Hispaniola from the early Miocene
onwards (Pubellier et al., 2000). The HFTB consists of a series of stacked NW – SE trending southwest-
verging thrust sheets with piggy-back basins on their northeastern flanks (Figure 5 - 4). The offshore
continuation of the HFTB into the Gulf of Gonâve and possibly south of Gonâve Island is visible on
seismic profiles (Mann et al., 1995). Shallowing of depositional facies in the piggy-back basins records
the progressive activation of the thrust faults (Figure 5 - 4; Pubellier et al., 2000). Activation of a thrust
results in uplift and erosion of the associated fold above the thrust plane, which in turn provides
clastics that can be deposited in basins at the front and back of the structure. The onset of clastic
sedimentation within these basins thus records the uplift of the anticline. Due to the lack of paleo-
current directions it is not possible to establish the exact source region of the clastics. The timing of
activity on a thrust fault is therefore best interpreted as the average of the recorded onset of clastic
sedimentation in the basins associated with this thrust. The northern Plateau Central and northern
Haiti record activation of the southern Mt. Paincroix Fault Zone (SMPFZ) (N5, Pubellier et al., 2000; N5
≈ 20 Ma, Hilgen et al., 2012; Figure 5 - 4). The northern Montagnes Noires and Artibonite Valley record
the activation of the Montagnes Noires Fault Zone (MNFZ) (N11, Pubellier et al., 2000; N11 ≈ 14 Ma,
Hilgen et al., 2012), while the Chaîne des Matheux and northern Cul-de-Sac plain records Matheux
Fault Zone (MFZ) activation (N15, Pubellier et al., 2000; N15 ≈ 10 Ma, Hilgen et al., 2012; Figure 5 - 4).

168
Chapter 5: Geology of the Chaîne des Matheux

5.3 Detailed stratigraphy of the eastern Chaîne des Matheux


The stratigraphy of the eastern Chaîne des Matheux is studied north of the Cul-de-Sac plain along two
northeast – southwest trending transects that are separated by around 15 km (Figure 5 - 10). The
locations described in sections 5.3.1 and 5.3.2 below are found on the geological map (Figure 5 - 10).
Each transect is divided into a northern flank and a southern flank. The eastern section follows Route
National #3, from Trou Caïman to Mirebalais, with the corresponding stratigraphic columns presented
in Figure 5 - 5. The western section follows Route Titanyen – Saut-d’Eau, from Titanyen to Dubon, with
the corresponding stratigraphic columns presented in Figure 5 - 6. These stratigraphic columns are
used to build the cross sections in section 5.4.

5.3.1 Stratigraphy Trou Caïman – Mirebalais section


Northern flank
The northern flank of this section consists of the elevated terrain from Bois Dorme northwards (Figure
5 - 10). Around Bois Dorme, a tectonic breccia containing massive white limestones with locally intact
dm-beds of yellow to red benthic foraminiferal and algal limestones outcrop. Based on the exposure
in a quarry just northwest of Bois Dorme, this sequence measures at least 150m. Between Bois Dorme
and Fond Diable a 50m thick coral biostrome complex rich in benthic foraminifera and red algae
outcrops, which is overlain by around 200m of dm-bedded to massive white to yellow discocyclina-
bearing foraminiferal limestones. Around Terre Rouge a repetition of dm-bedded to massive yellow
limestones is observed. The combined thickness of these platform limestones is at least 400m. The
facies are the same as the lower to middle Eocene platform limestones described by Butterlin (1960)
and Desreumaux (1987). Between Terre Rouge and Saintonge the lithology changes to dm-bedded
well-stratified grey to yellow limestones with brown chert nodules interbedded with minor cm-thick
marly layers. This 300m thick calcareous sequence contains meter to 10’s of meters thick stratigraphic
intervals of limestone breccia in the lower part. These intra-formational breccia intervals have an
erosive base and consists of pebble to boulder sized angular to subrounded clasts of limestone and
coral debris, floating in a marly matrix. The breccia intervals are interpreted as mass transport deposits
(MTD’s). The age for this sequence ranges from late Eocene to early Miocene (Butterlin, 1960; Van den
Berghe, 1983a; Desreumaux, 1987). The upper 100m are of Aquitanian age, while the lower 200m are
of late Eocene or Oligocene age (Van den Berghe, 1983a). The local absence of the Oligocene (Van den
Berghe, 1983a) is possibly explained by the erosive nature of these submarine mass transport deposits.
No sub-aerial erosive contact could be observed for the Oligocene interval.

170
Chapter 5: Geology of the Chaîne des Matheux

Southern flank
The southern flank of this section consists of the terrane at low elevation south of Bois Dorme, most
notable along the Ravine Sèche (Figure 5 - 10). The lowermost stratigraphic interval is outcropping
directly west of Trou Caïman and consists of white well-stratified foraminiferal limestones. These are
similar to the lower to middle Eocene platform limestones observed on the northern flank. Around 2
km into the Ravine Sèche, a 100m thick sequence of beige micritic limestones with brown chert
horizons is found. This description fits well with the Aquitanian limestones of Desreumaux (1987),
although the age could also be late Eocene or Oligocene. Overlying is a series of chalky limestones with
minor cherts. Due to the strong deformation it is not possible to estimate the thickness. The chalks are
similar to the Burdigalian chalks described by Van den Berghe (1983). The chalks are in faulted contact
with an overlying sequence of massive conglomerates with a calcareous matrix. Bed thickness ranges
from a few to up to 10 meters and clasts are rounded to subrounded pebbles to boulders. At the base
the matrix is well cemented and calcareous, while stratigraphic upwards the matrix becomes more
marly and less consolidated. Total thickness of this conglomeratic sequence is at least 150m. These
conglomerates differ from the late Eocene to Oligocene MTD’s on the northern flank by their
roundness, thickness, color, and consolidation. Based on their stratigraphic position these
conglomerates are probably Serravallian to Tortonian. Upwards in the stratigraphy and downstream
of the Ravine Sèche the conglomerate beds become thinner and the abundance of them within the
sequence decreases. The sequence becomes dominated by poorly consolidated greenish silts and clays
containing coral fragments, interbedded with cm-beds of rusty silty limestones. This indicates a
progressively more distal source for the clastics, but it is difficult to interpret the depositional
environment. This sequence has not been observed on the northern flank. It could represent the
equivalent of the calcareous silts of the Tortonian to Messinian Thomonde Formation (Bourgueil et al.,
1988), but the lithologies also fit well with the description of Pliocene multicolored clays, silts and
fossiliferous sands in the offshore Cul-de-Sac-1 (Crux) well (Bourgueil et al., 1988). Total thickness for
this Tortonian to possibly Pliocene sequence is at least 300m. A pronounced reefal limestone interval
of around 90 m as drilled in the Cul-de-Sac-1 (Crux) well (Bourgueil et al., 1988) is not observed in the
northern Cul-de-Sac area.

172
Chapter 5: Geology of the Chaîne des Matheux

Southern flank
The southern flank of this section lies south of Beauget (Figure 5 - 10). The highest stratigraphic interval
is observed in the southern footwall of the thrust at Beauget. It consists of calcareous silts and marls
that become more calcareous downwards in the stratigraphy. North of Saint Phard this sequence of
limestones, siltstones and marls is interbedded by basalts. These basalts are found in the same region
as the Tortonian basalts described by Boisson (1987). The sequence is likely part of the Thomonde
Formation, with a total thickness of about 400m. The top is probably located in the Messinian, while
the lower part with the intercalated basalts is Tortonian. Southwest of Saint Phard grey to green silty
claystones with rusty-colored limestone beds were observed, similar to those found in the Ravine
Sèche section further east. These are interpreted to be part of the Tortonian to Messinian Thomonde
Formation (Bourgueil et al., 1988). Downwards in the stratigraphy and further south the lithology
becomes more and more calcareous and changes through well-stratified silty bioclastic limestone to
chalks. The silty limestones are the equivalent of the (late Serravallian – early Tortonian?) Madame
Joie Formation, with the chalks probably of Langhian or Serravallian age. The thickness of these
sequences is around 50m for the Madame Joie and 100m for the chalks. Directly north of Titanyen the
lithology abruptly changes from chalks to greenish silts and clays with rusty-colored limestone beds of
the Tortonian to Messinian Thomonde Formation.

174
Chapter 5: Geology of the Chaîne des Matheux

5.4 Outcrop-scale style of deformation


The intensity and style of deformation across the Chaîne des Matheux changes depending on the
lithology and the structural positionof the rocks. The style and intensity of deformation remains similar
along-strike of the thrust sheet.

The change in intensity of deformation is best observed when comparing the Burdigalian and younger
lithologies on the northern flank with their lithological equivalents on the southern flank. The division
between northern and southern flank follows the definition used in section 5.3 and shown on Figure 5
- 10. A minor change in intensity is recorded from north to south on the northern flank. On the northern
flank around Gimballe (Figure 5 - 10), the upper Miocene is undeformed and mildly tilted towards the
northeast (Figure 5 - 7a). South of Gimballe the middle Miocene is locally affected by NE-verging
thrusts, but otherwise continues to gently dip towards the northeast. The Burdigalian north of Triano
(Figure 5 - 10) is mildly folded and cut by sigmoidal fractures with a top-to-SW motion (Figure 5 - 7b).

The Burdigalian and younger lithologies on the northern flank are only mildly deformed and locally
transected by thrusts.

In contrast, the Burdigalian and younger lithologies on the southern flank of the Chaîne des Matheux
have higher intensities of deformation. At Ravine Sèche (Figure 5 - 10), a major thrust separates the
lower Eocene hanging wall from the Miocene footwall. The Miocene is folded (Figure 5 - 7c) and locally
cut by NE-verging back-thrusts (Figure 5 - 7d). Between Titanyen and Beauget (Figure 5 - 10) the
Burdigalian and younger footwall is also folded (Figure 5 - 7e) and transected by thrust faults. In this
area the Miocene is often characterized by steep dips that remain unchanged for several tens of meters
and then change orientation abruptly via tight folds.

The Burdigalian and younger lithologies in the southern hanging wall of the Chaîne des Matheux are
generally more strongly deformed by folds and thrusts compared to the northern flank.

There is a significant difference in style and intensity of deformation on the northern flank of the
Chaîne des Matheux, between Burdigalian and younger chalky limestones and calcareous sandstones
on the one hand, and middle Eocene to Aquitanian well-stratified dm-bedded cherty mudstones on
the other.

On the northern flank of the Chaîne des Matheux (Figure 5 - 10), between Beauget and Hatte Palicarpe
in the west, and Fond Diable and Triano in the east, these middle Eocene to Aquitanian cherty
mudstones are characterized by; 1) various-wavelength gentle to tight folding along mainly NW-
trending axes (Figure 5 - 8a), 2) intra-formational deformation aided by decollement levels and NE- or

176
Chapter 5: Geology of the Chaîne des Matheux

5.5 Map-scale deformation structures


The largest structure is the ~20 km half-wavelength anticlinal NW – SE trending Chaîne des Matheux
(Figure 5 - 10). This asymmetric structure consists of several stacked thrust sheets with a dominant
SW-vergence (Figure 5 - 11). The individual thrust sheets are characterized by a long northeast dipping
back limb and a short southwest dipping front limb. The thrust sheets are bounded on the southwest
by northeast dipping thrust faults (Figure 5 - 11). The most prominent of these faults is the Beauget
Thrust (BT), which juxtaposes Eocene limestones in the hanging wall against upper Miocene calcareous
clastics in the footwall (Figure 5 - 10). To provide a geometric fit to the secondary structures on the
hanging wall of the BT requires between 8 km (Figure 5 - 11a) to 10 km (Figure 5 - 11b) of overthrusting.
One of the faults bounding these secondary structures is the Ravine Sèche Thrust (RST) (Figure 5 - 10).
This fault is interpreted as being rooted on the BT at depth (Figure 5 - 11). The footwall of the BT is also
deformed by secondary thrusts with limited offset. On such example, the Titanyen Thrust (TT),
develops fault-propagation folds at the fault’s tip (Figure 5 - 11a). To the southwest of the RST, the BT
and TT faults have a slightly larger NE-dip than the overlying formations (Figure 5 - 11). Maastrichtian
flysch, which is outcropping as a fault-bounded sequence in the western Chaîne des Matheux (see
section 5.1.1), possibly provides a decollement level for the BT and TT (Figure 5 - 11). The thrust sheet
northeast of the BT fault is transported over the Tortonian calcareous silts in the footwall, which acts
as a decollement level (Figure 5 - 11). The decollement levels result in a thin-skinned tectonic style, but
it cannot be excluded that some of the thrusts are rooted on basement involved, thick-skinned thrusts.

Another feature is the Quaternary basaltic volcanism at La Vigie, whose edifices are aligned on a NW
– SE trend along the central axis of the Chaîne des Matheux (Figure 5 - 10; Traineau and Westercamp,
1980), with Morne La Vigie being the best preserved (Bourgueil et al., 1988). La Vigie volcanism is
related to a strombolian-type of eruption, which started as submarine but later evolved to subaerial
(Bourgueil et al., 1988). Eruption of the Thomazeau basalts and volcaniclastics occurred in a submarine
setting from a volcanic edifice just east off the map area (Figure 5 - 10; Bourgueil et al., 1988). This
eruption possibly resulted from activity along transtensional NE – SW trending faults (Pubellier et al.,
1991, 2000) that delineate the basalts to the west (Figure 5 - 10). At present, the Thomazeau basalts
are at elevations up to 700m above the Cul-de-Sac Plain. The submarine extrusive setting and 1.6 – 1.8
Ma (K-Ar, Wadge and Wooden, 1982; Pubellier et al., 1991) age of the basalts indicates at least 700m
of relative uplift between the Cul-de-Sac and the basalts in less than 1.6 Myr, which averages to 0.44
mm yr-1. This uplift rate is of the same order of magnitude as the <133 Ka uplift rates calculated by
Mann et al. (1995) for the NW Peninsula of Haiti (0.37 mm yr-1) and western Chaîne des Matheux (0.19
mm yr-1). The responsible mechanism for this Pleistocene uplift is unknown (Mann et al., 1995).

181
Chapter 5: Geology of the Chaîne des Matheux

5.6 Shortening of the Chaîne des Matheux


This section quantifies the amount of shortening of the Chaîne des Matheux during post-Eocene times.
It uses the same techniques as outlined in section 4.6 for shortening of the Southern Peninsula.

The Chaîne des Matheux is uplifted in the late Miocene (section 4.2), when southward propagating
deformation of the Haitian Fold-and-Thrust belt reached this area (Pubellier et al., 2000). The structural
trend of the eastern Chaîne des Matheux is predominantly NW – SE (Figure 5 - 10) and the cross
sections are constructed from NE to SW, parallel to the direction of shortening, to limit issues with out-
of-plane transport of material. Bedding-parallel decollement levels are mainly observed in the Eocene
and Oligocene cherty limestones (Figure 5 - 11). To overcome the problem of preservation of bedding
length during deformation, a similar approach is applied as for the Southern Peninsula (section 4.6).
The line-length of a formation structurally above the decollements, in this case the top Oligocene, is
chosen for restoration. This choice makes for an easier comparison in chapter 7 with other published
cross sections from Haiti and the Dominican Republic.

The amount of shortening along the Titanyen – Dubon section of the eastern Chaîne des Matheux is
calculated using the cross section of Figure 5 - 11a. The top Oligocene (black line) is used for line-length
restoration. When this formation is eroded the trace is interpreted (dashed orange line) using the
underlying Eocene formations. The amount of post Oligocene shortening of the entire section is 6.11
km or 18.4%. The majority of shortening is accommodated in the northeastern hanging wall, which is
overthrusting the footwall at the Beauget Thrust (Figure 5 - 11a). To compare the difference in
shortening, the line-length of the top Oligocene is restored for both domains. The division between
the two domains is chosen to be vertically under the Beauget Thrust, which means that shortening due
to overthrusting at this thrust is accredited to the hanging wall of the section. This division is chosen
since it; 1) allows an easier comparison with the Trou Caiman – Mirebalais section (Figure 5 - 11b),
which is constructed entirely northeast of the Beauget Thrust, and 2) it makes it easier to plot
shortening on a map, as is done in chapter 7 for a discussion with published data. Shortening of the NE
hanging wall is 5.54 km or 23.6%, while the SW footwall is shortened by 0.57 km or 5.9%.

The same exercise is performed for the Trou Caiman – Mirebalais cross section (Figure 5 - 11b). This
section is located entirely to the northeast of the Beauget Thrust, which intersects the surface at the
southwestern end of the cross section. Total shortening of the entire section is 11.96 km or 30.7%.

Additional shortening is possibly accommodated at depth under the Morne La Vigie (Figure 5 - 11a)
and Terre Rouge (Figure 5 - 11a) regions. This can for instance result from the imbricate stacking of
thrust sheets in duplexes or in fish-tail structures by associated back-thrusts.

182
Chapter 5: Geology of the Chaîne des Matheux

5.7 Paleo-stress and relative timing


Paleo-stresses in the Chaîne des Matheux are computed from outcrop measurements on shear
fractures in the same way as described in section 4.2. Paleo-stress data in the eastern Chaîne des
Matheux is limited to only 8 distinct paleo-stress states from four localities (Figure 5 - 10). The relative
timing of these paleo-stress states are described in this section, with their absolute timing discussed
in section 5.7.2.

Along the western cross section (A), southwest of Hatte Palicarpe (Figure 5 - 10), upper Eocene to
Oligocene slope limestones are pervasively fractured, faulted and folded (Figure 5 - 12). Orthogonal
and bedding perpendicular joints trend roughly ENE – WSW and NNW – SSE after restoring the bedding
to a horizontal position (Figure 5 - 13). The ENE - WSW joints (set 1a) are most systematic and
dominant, with NNW - SSE cross joints (set 1b) usually abutting against them (Figure 5 - 13), indicating
that the set 1a systematic joints were created prior to set 1b cross joints. Both joint sets are reactivated
as shear fractures by a strike-slip stress regime with NE – SW trending σ1 axis (Figure 5 - 13a, same as
Figure 5 - 10.1). The limestones hosting these reactivated joints are folded along NW-trending fold
axes (Figure 5 - 12c). Set 1 joints are in a bedding-perpendicular orientation throughout the structure,
indicating that the phase of folding post-dates creation of these joints. The structure is cut by two main
thrusts with a different orientation and vergence. One thrust is verging towards the NNE (thrust #1,
Figure 5 - 12c), while a second thrust is verging towards the SW (thrust #2, Figure 5 - 12c). The bedding
planes are locally reactivated as shear fractures. This reactivation resulted in sigmoidal shear fractures
that are slightly inclined to the bedding, but also created drag on the perpendicular joint sets close to
the bedding plane (Figure 5 - 13c). Where the bedding is not reactivated the set 1 joints are occasionally
cross-cutting the bedding plane (Figure 5 - 13a). These observations indicate that reactivation of the
bedding post-dates creation of the bedding perpendicular set 1 joints.

Irrespective of the position on the structure (Figure 5 - 13a vs Figure 5 - 13c) the reactivated bedding
planes show a top-to-SE motion. In the present-day orientation, the paleo-σ1 axis of the reverse regime
that reactivated the bedding is striking NW – SE (Figure 5 - 13c, same as Figure 5 - 10.2), and is in an
Andersonian configuration. Back-tilting the bedding to a horizontal position tilts the paleo-stress axes
out of an Andersonian configuration, indicating that these bedding planes were probably reactivated
in their present-day orientation. Reactivation of these bedding planes is not related to folding of the
limestones along the NW – SE axial trend of this structure. Flexural slip on the bedding during folding
would initiate slip in a direction perpendicular to the fold axes (i.e. Twiss and Moores, 2007), hence
top-to-SW. The direction of slip is however almost parallel to the fold axes, indicating that it results
from a different deformation event.

184
Chapter 5: Geology of the Chaîne des Matheux

The top-to-SE motion of the reactivated bedding is almost parallel to the slip on the second thrust
(thrust #2, Figure 5 - 12c). As the bedding is reactivated by a NW – SE trending reverse regime (Figure
5 - 13c), it is likely that thrust #2 also results from this stress regime, which thus post-dates top-to-SW
thrusting and folding.

In summary, the relative sequence of deformation for the structure described above is as follows; 1)
creation of ENE – WSW (set 1a) and NNW - SSE (set 1b) bedding perpendicular joints, followed by; 2)
reactivation of these joint sets as shear fractures by a strike-slip stress regime with NE – SW trending
σ1 axis; 3) folding and thrusting with a SW-vergence, and; 4) reactivation of the bedding planes with
top-to-SE motion and creation of SE-verging thrusts, resulting from a reverse regime with NW – SE
trending σ1 axis.

In the Ravine Sèche (Figure 5 - 10), Pubellier et al. (1991) also observed a fold whose bedding planes
are reactivated by slip parallel to the NW – SE trending fold axis, which is similar to the structure in
Figure 5 - 12.

Aquitanian cherty limestones outcropping at Ca Michaud (Figure 5 - 10) are also pervasively fractured
by two orthogonal bedding-perpendicular joint sets (Figure 5 - 14). These joints are trending NE – SW
and NW – SE (Figure 5 - 14) and are thus at a 30° CCW angle from the joint sets found in the structure
3 km to the northeast north (Figure 5 - 13). Abutting relations are less clear, but mainly indicate that
the NW – SE joints are abutting against the NE – SW joints. Due to the similarity with the joint sets
further north, these joints are labeled set 1a for the older NE – SW systematic joints, and set 1b for the
younger NW – SE cross joints. Both joint sets are locally displaced by minor NE – SW trending normal
shear fractures. The joints and minor normal shear fractures are in turn displaced by reactivated
bedding and top-to-SSE verging thrusts, which is similar to the structure of Figure 5 - 12 further north.

A series of N – S trending shear fractures are found at an adjacent outcrop (Figure 5 - 10.3), which
resulted from a back-tilted Andersonian strike-slip regime with ENE – WSW trending σ1 axis. This
orientation is similar to the back-tilted σ1 axis of the strike-slip regime that reactivated joint sets 1a
and 1b (Figure 5 - 10.1)

Approximately 1.5 km southwest of Ca Michaud, strike-slip and reverse shear fractures transect
Oligocene to Aquitanian cherty limestones (Figure 5 - 10.4), which result from two distinct stress states.
The first fracture set (Figure 5 - 10.4a) was initiated by a back-tilted reverse stress regime with NNW -
SSE trending σ1 axis. This stress regime is similar to the stress regime responsible for the top-to-SE
motion observed in the structure southwest of Hatte Palicarpe (Figure 5 - 10.2). The second fracture
set (Figure 5 - 10.4b) resulted from a back-tilted strike-slip regime with ENE – WSW trending σ1 axis,

186
Chapter 5: Geology of the Chaîne des Matheux

5.8 Absolute timing of paleo-stress states and deformation phases


In the eastern Chaîne des Matheux, Eocene to Serravallian limestones record the oldest deformation
phase, which resulted in two orthogonal bedding perpendicular joint sets trending ENE - WSW to NE –
SW (set 1a) and NNW – SSE to NW – SE (set 1b) (Figure 5 - 13 and Figure 5 - 14). Pervasively spaced
joints are also observed in Eocene formations in the Montagnes Noires and Terre Neuve regions by
Pubellier et al (1991). These authors propose that the joints could have formed as recent as late
Miocene times. In the Chaîne des Matheux the set 1 joint affect Serravallian and older formations,
thereby roughly limiting their creation to pre-Tortonian in the eastern Chaîne des Matheux (Figure 5 -
15). In this region the joint sets were reactivated by a strike-slip regime with NE – SW trending σ1 axis
(Figure 5 - 13a). Orthogonal joint sets that pre-date folding are also reported from the Split mountain
anticline in Utah (Silliphant et al., 2002) and anticlines in Wyoming (Laubach and Lorenz, 1992;
Hennings et al., 2000; Bergbauer and Pollard, 2004). The possible mechanisms for creating orthogonal
bedding perpendicular joints are already addressed in section 4.4. The pre-Tortonian systematic joints
trend E – W to NE – SW, roughly parallel to the direction of maximum principal stress during that time
(section 5.6). It is therefore probable that set 1a joints formed in response to layer-parallel shortening.
The pre-Tortonian timing indicates that these joints formed prior to folding and thrusting in the area,
which started during the Tortonian (section 5.2; Pubellier et al., 2000). Layer parallel shortening is often
a precursor to, or develops during the early stages of, folding and thrusting (Bellahsen et al., 2006;
Tavani et al., 2006; Amrouch et al., 2010). The set 1 joints possibly developed by layer-parallel
shortening in response to far-field stresses introduced by folding-and-thrusting in the Montagnes
Noires further north, which started at the onset of the Serravallian (section 5.2; Pubellier et al., 2000).
If the joints are indeed related to this process, it roughly constrains their initiation to the Serravallian
(Figure 5 - 15).

The timing of SW- and minor NE-verging thrusts and folds in the Chaîne des Matheux is constrained to
the Tortonian by the arrival of clastic erosional products of the uplifted chain in adjacent basins (section
5.2; Pubellier et al., 2000). This deformation is associated with a reverse stress regime with NE – SW
trending σ1 axis (Figure 5 - 10.5). If the joints indeed formed during the Serravallian, and are
subsequently reactivated by a strike-slip regime with NE – SW trending σ1 axis prior to folding and
thrusting in the Tortonian, the timing of this strike-slip regime is thus roughly constrained to the
Serravallian or early Tortonian. The minor NE – SW trending normal shear fractures at Ca Michaud are
possibly related to this strike-slip regime. The timing of this strike-slip phase fits well with observations
in the Ravine Sèche by Pubellier et al. (1991), where a N025°E trending strike-slip fault in middle (?)
Miocene sediments is subsequently folded along a NW – SE trending axis, indicating middle Miocene
or older strike-slip activity in the area (Figure 5 - 15).

188
Chapter 5: Geology of the Chaîne des Matheux

Wooden, 1982; Pubellier et al., 1991). If volcanic activity is indeed associated with this transtensional
phase, as proposed by Pubellier et al. (1991), it would indicate that the associated strike-slip regime
was at least active during the onset of extrusion at 1.8 Ma. The basalts at La Vigie have a similar
geochemistry as the basalts at Thomazeau (Kamenov et al., 2011b). Field observations indicate that
the La Vigie basalts are relatively undeformed and not affected by thrusting. The 2.3 Ma age of these
basalts (Wadge and Wooden, 1982; Pubellier et al., 1991) implies that folding-and-thrusting had mainly
ceased in the Chaîne des Matheux by Pleistocene times. This in turn limits the transtensional strike-
slip regime to the Quaternary (Figure 5 - 15).

190
Chapter 6: Paleo-fluid circulation

6 Paleo-fluid circulation

Nomenclature
A fracture is a surface along which a rock has broken. Based on relative motion across the fracture
surface during formation, two end-member types are distinguished: 1) extension fractures, where
relative motion upon fracture propagation is perpendicular to the fracture walls (mode-I propagation)
causing dilatation of the fracture, 2) shear fractures, where relative motion upon fracture propagation
is parallel to the surface (mode-II/III propagation). A fracture that experienced displacement both
parallel and perpendicular (dilatation) to the surface is an oblique extension fracture (also known as
a hybrid fracture). Extension fractures with very small displacements perpendicular to the fracture
surface and none, or very little, displacement parallel to it are called joints, which are often regularly
spaced (Twiss and Moores (2007), p37 and p38). Additionally, stylolites experience fracture-wall
perpendicular contraction, sometimes termed mode-IV (Fossen, 2010, p124)

Veins are extension fractures filled with mineral deposits (Twiss and Moores (2007), p42). A
component of dilatational motion perpendicular to the fracture walls must have occurred in order to
provide the space for the vein-forming minerals to precipitate into. Veins are thus associated with
either extension or mixed fractures or joints.

A surface of broken rock in or around a fault zone shall hereinafter be termed ‘fracture’. The modifier
‘shear fracture’ is reserved for fractures that have kinematic criteria indicating displacement parallel
to the fracture’s surface at the scale of the outcrop. The term ‘joint’ is reserved for pervasively and
regularly spaced fractures. The origin of the fractures and joints is subsequently addressed in the text
where appropriate.

The distinction between shear fractures and faults is also ambiguous and scale-dependent (Twiss
and Moores (2007), p61). Shear fractures are characterized by mm- to dm-scale displacements, while
faults have larger offsets (Fossen (2010), p121). Faults are often associated with a zone of deformation
encompassing the slip surface (Faulkner et al., 2010). These criteria; m-scale displacement and/or fault
zone, are used when distinguishing a shear fracture from a fault in the field, and is also used hereinafter
in the text.

191
Chapter 6: Paleo-fluid circulation

This chapter aims at linking the deformation phases that affected southwestern Haiti to the
corresponding paleo-fluid circulation. This is achieved by a geochemical and micro-thermometric
analysis of the different vein generations sampled in the field and associated with these deformation
phases.

It firstly provides a review of the structure, mechanisms and fluid flow properties of fault zones in
different lithologies and tectonic settings. This is followed by a review of the mechanisms responsible
for the creation of veins, their morphological characteristics, and post-precipitation deformation
mechanisms.

The results are separated for the Chaîne des Matheux and the Southern Peninsula, whose tectonic
evolution is presented in chapters 4 and 5, respectively. These results are presented in 5 subsections,
containing: 1) an overview of the different vein sets, their relation with deformation phases, and
microscopic observations on these veins including cathodoluminescence, 2) stable oxygen and carbon
isotope analysis of the veins and associated host rocks, 3) fluid inclusion microthermometry on
selected vein samples, 4) REE + Yttrium (REY) composition of selected vein and host rock samples, 5) a
discussion on the relation between the tectonic setting of the veins, their timing, and associated
geochemistry, and the possible fluid sources from which these veins precipitated.

192
Chapter 6: Paleo-fluid circulation

6.1 Structure, mechanisms, and fluid flow properties of fault zones


Despite occupying only a small volume of the earth’s crust, fault zones exert a strong control on the
crust’s mechanical and fluid flow properties. Fault zones are lithologically discontinuous,
heterogeneous and anisotropic. Brittle fault zones consist of one or multiple slip surfaces nested within
regions of high (fault core) and low (damage zone) strain, relay ramps, linked segments, contractional
and dilatational jogs and Riedel shears. They commonly display variations in complexity down dip or
along strike, and their structure, mechanical properties and permeability varies over time and space
(Faulkner et al., 2010).

6.1.1 Characteristics of fault zones, fault cores, and damage zones


In its simplest form the structure of a fault zone consists of a region of localized strain, termed the fault
core, which is surrounded by a zone of distributed fractures and faults, termed the damage zone
(Figure 6 - 1a; Chester and Logan, 1986; Caine et al., 1996). The core of a fault generally consists of
cataclasite, gouge, or a combination, while the damage zone is characterized by fractures and
secondary faults (Faulkner et al., 2010). Strain can be focused on discrete slip surfaces (i.e. Punchbowl
Fault, Chester and Logan, 1986) or can be homogenously distributed throughout the fault core (i.e.
Carboneras Fault, Rutter et al., 1986; Faulkner et al., 2003). Fault zones can also consist of multiple
cores that branch, anastomose and link, thereby encompassing lenses or blocks of fractured host rock
(Figure 6 - 1b; i.e. San Andreas Fault, SAFOD borehole, Zoback et al., 2010, 2011; Bradbury et al., 2015).
The structure of a fault zone depends on the depth of formation (Ishii et al., 2010), magnitude of
displacement, lithology, tectonic setting, and fluid flow (Faulkner et al., 2010).

Faults initiate by the amalgamation of growing dilatational (mode-I) cracks (Reches and Lockner, 1994;
Healy, 2008; Blenkinsop, 2008), which may use pre-existing planes of weakness such as tectonic joints
(Wilkins et al., 2001; Joussineau and Aydin, 2007; Moir et al., 2010). As the fault develops, progressive
damage from deformation leads to the development of a wider fault core (Scholz, 1987). Due to
complexities within the fault zones and lithological effects, there is probably not a linear increase
between fault core width and displacement (Evans, 1990), and fault cores may even show a constant
thickness with increased displacement (Shipton et al., 2006). However, fault (damage) zone width does
appear to scale with fault displacement. This is an important observation since damage zones may
provide economically significant fluid flow pathways (Faulkner et al., 2010). Fault and fault damage
zone width initially scales linearly with fault displacement (Wibberley et al., 2008), but decreases
strongly for faults with >1000m displacement (Figure 6 - 1c; Vermilye and Scholz, 1998; Micarelli et al.,
2006; Mitchell and Faulkner, 2009; Savage and Brodsky, 2011). This apparent break in damage zone
scaling is attributed to the evolutionary change from a single isolated fault towards the formation of

193
Chapter 6: Paleo-fluid circulation

secondary fault strands that, upon further displacement, result in complex multi-cored fault systems
(Savage and Brodsky, 2011).

Fault core
Lens of Strand of gouge or Host
Damage Fault Host Damage zone damaged rock (ultra)cataclasite rock
zone core rock
A B C
106

Fault zone width (m)


Secondary fault 104
& associated
localized
damage 103

102
SAF

100
fracture
density
100 102 103 104 105 106
Fault displacement (m)
permeability

Figure 6 - 1: A/B; fault zone architecture, modified from Faulkner et al. (2010). A: Simple fault zone structure,
with high-strain core surrounded by fractured damage zone. B: Multiple fault core model, with many strands of
highs-strain material enclosing lenses of fractured host rock. C: Fault displacement versus fault damage zone
width, modified from, and references within, Savage and Brodsky (2011). Thick black line is best fitting curve.
Thin black lines within graph delineate data. Light grey column are 80% confidence limits for changing point from
linear to decreasing scaling relationship between displacement and width. Abbreviation: SAF = San Andreas Fault.

In a fault zone with single-core architecture, the fault core is classically viewed as an across-fault fluid
flow barrier, while the damage zone acts as an along-fault fluid flow conduit (Figure 6 – 1a; Caine et
al., 1996). Fault zones containing multiple slip planes and fault cores are governed by high and low
permeability features that result in significant permeability anisotropy and heterogeneity, resulting in
a more complex hydrological behavior (Figure 6 – 1b; Faulkner et al., 2010). Permeability of the fault
zone is strongly controlled by the architecture of fractures and faults, and their relation to the regional
stress field (Lunn et al., 2008). Although the permeability of slip surfaces and fractures may be
substantially larger than the permeability of the unfractured host rock, their ability to transfer fluids
along and across the fault is entirely dependent upon their connectivity (Faulkner et al., 2010).

194
Chapter 6: Paleo-fluid circulation

6.1.2 Lithological control on fault zone and fluid flow characteristics


The type of lithology, being crystalline, granular or phyllosilicate-rich, together with the porosity of
that rock, exerts a strong control on fault zone characteristics and permeability (Faulkner et al., 2010).

Coarse grained, high porosity rocks commonly have fault cores consisting of amalgamated low
permeability/porosity deformation bands and high permeability slip surfaces, surrounded by a damage
zone containing compaction and cataclastic deformation bands (Antonellini et al., 1994; Fossen et al.,
2007; Rotevatn et al., 2013). In contrast, low porosity rocks, or rocks deformed under low effective
stress conditions, commonly have a fine grained fault core (Balsamo et al., 2010) surrounded by a
damage zone containing extension fractures (Evans et al., 1997; Blenkinsop, 2008).

Fault core

The hydrological properties of the fault core are influenced by the type of fault gouge, which falls into
two categories. The first type of gouge consists of granular material composed of broken, irregular
clasts, while the second type is composed of a variable proportion of phyllosilicate material.

The granular gouges develop a characteristic grain size distribution, suggesting a similar permeability
development for all granular gouges (Sammis et al., 1987; Marone and Scholz, 1989). The permeability
of granular gouges is reduced with increased amounts of shear strain (Crawford et al., 2008). Two to
three orders of magnitude reduction for shear strains up to 10, and another two to three orders of
magnitude for shear strains up to 200, including one order of magnitude permeability anisotropy, are
documented in laboratory experiments (Zhang and Tullis, 1998). This is in agreement with bore-hole
measurements that suggest a significant drop in across-fault permeability for deformation band-
dominated faults in high-porosity rocks (Shipton et al., 2002, 2005).

Because clay-content has a dominant effect on fault core permeability, phyllosilicate rich gouges tend
to have lower permeability than tectosilicate (i.e. quartz and feldspar) rich gouges (Crawford et al.,
2008). Permeability measurements on natural clay-rich fault gouge demonstrate the low permeability
of this material (Morrow and Byerlee, 1992; Faulkner et al., 2003; Wibberley and Shimamoto, 2003;
Tsutsumi et al., 2004; Mizoguchi et al., 2008), with up to three orders of magnitude permeability
anisotropy (Faulkner and Rutter, 2000). This anisotropy does not only result from a preferred
alignment of the clay minerals, but is strongly controlled by alternating microlayers of clay-rich and
porous granular material (Faulkner et al., 1998; Faulkner, 2004).

The temporal evolution of fault gouge permeability is interesting from the point of view of earthquake
cyclicity. Modeling on granular gouges at conditions representative for earthquake-depths show a

195
Chapter 6: Paleo-fluid circulation

The damage zone in high porosity rocks is often more complex, with permeability controlled by the
interplay between high permeability slip surfaces and low permeability deformation bands (Lunn et
al., 2008). This interplay can result in a significant decrease in damage zone permeability with
increasing deformation (Main et al., 2000). High porosity rocks, with deformation band dominated
damage zones, show a relatively constant microfracture density within the damage zone, and no
observable change in fracture density with distance to the fault core (Figure 6 - 2b; Anders and
Wiltschko, 1994; Shipton and Cowie, 2001).

Carbonates

Fault architecture, fault slip behavior, and deformation mechanisms for carbonate rocks can be
markedly different from brittle siliciclastic and crystalline rocks. These differences mainly result from
the ability of carbonates to deform by ductile processes, dynamic recrystallization and fluid related
pressure-solution at shallow burial conditions (<3 km), and the mechanical and lithological variability
of multi-layered carbonates with variable clay content (Delle Piane et al., 2017). At the outcrop scale
brittle, cataclastic structures dominate, while at the microscale ductile, crystal plastic features are
predominant, even at shallow crustal conditions and at low temperatures (Kennedy and Logan, 1998;
Kennedy and White, 2001; Kirschner and Kennedy, 2001; Liu et al., 2002; Molli et al., 2011).

Low porosity carbonates


Faults in low porosity carbonates, such as platform limestones, dolostones, and other massive
carbonates, have many features in common with faults in low porosity siliciclastic or crystalline rocks.
Within low porosity carbonates, faults also develop a core that is dominated by brittle processes, and
a damage zone with a highly connected network of fractures (Agosta and Kirschner, 2003; Billi et al.,
2003; Agosta and Aydin, 2006; Micarelli et al., 2006b; Bussolotto et al., 2007, 2015; Bastesen and
Braathen, 2010; Molli et al., 2011; Rustichelli et al., 2016). Fracture density in the damage zone is
similarly decreasing with distance from the core (Cello, 2000; Cello et al., 2001; Micarelli et al., 2003,
2006b; Daniel et al., 2004; Molli et al., 2011; Hausegger et al., 2010; Bullock et al., 2014; Smeraglia et
al., 2016; Cianfarra and Salvini, 2016). Well-developed polished fault surfaces characterized by an
extremely smooth slip surface (Siman-Tov et al., 2013; Fondriest et al., 2013), known as fault mirrors,
may develop in limestones (Jackson and McKenzie, 1999; Siman-Tov et al., 2013) and dolostones
(Fondriest et al., 2013). Calcite filled veins, highlighting the presence of fluids, are commonly found in
carbonate fault zones (Teixell et al., 2000; Tarasewicz et al., 2005; Bussolotto et al., 2015; Balsamo et
al., 2016). The interplay between fluids and deformation often results in pressure-solution seams

197
Chapter 6: Paleo-fluid circulation

(Salvini et al., 1999). The localization of strain on these pressure solution seams can result in shear
failure and the initiation of faults (Graham et al., 2003; Daniel et al., 2004; Wall et al., 2006; Agosta and
Aydin, 2006).

Phyllosilicate-rich carbonates
Fault zones in carbonates containing phyllosilicates (i.e. marly limestone) typically develop multiple
cores, diffuse core/damage zone boundaries, pervasive pressure solution, calcite veins, and foliated
cataclasites (Delle Piane et al., 2017). Foliation within cataclasites of the fault core results from the
alignment of phyllosilicates, which creates seams of platy minerals that favorably accommodate slip,
eventually resulting in the development of Riedel shears within the fault core (Logan et al., 1979; Rutter
et al., 1986). Within a more ductile environment the same process will result into S-C fabrics (Lister
and Snoke, 1984). Foliated cataclasites are reported in limestones with a clay content higher than 10%
(Marshak and Engelder, 1985), and from shallow (<3 km) depths under the influence of either meteoric
or deeper crustal fluids (Bussolotto et al., 2007; Meneghini et al., 2012; Clemenzi et al., 2014). At the
microscale, marly fault rocks display penetrative foliation, shear bands, and pervasive pressure
solution seams (Tesei et al., 2013; Bullock et al., 2014; Clemenzi et al., 2014). These pressure solution
seams grow thicker with increased deformation (Delle Piane et al., 2017). Microscale observations in
clay-rich fault rocks show that frictional sliding and pressure-solution are the dominant mechanisms,
while in pure massive limestones cataclasis is the dominant mechanisms (Tesei et al., 2013; Bullock et
al., 2014; Viti et al., 2014).

In multilayered limestone/marl systems, the fault core – damage zone transition is diffuse and
characterized by a progressive fading of the bedding, while massive limestones on the other hand
develop sharp, striated fault cores (Tesei et al., 2013, 2014; Clemenzi et al., 2014). With sufficient
mechanical difference between the multilayers, fault dip can change from shallow in clay-dominated
layers to steep in limestone dominated ones, and even create flat-ramp-flat fault geometries in
compressive settings (Peacock and Sanderson, 1992, 1994; Giorgetti et al., 2016; Ferrill et al., 2017a;
b). Within a multilayered sequence, deformation is preferentially taken up by the clay-rich intervals,
with the limestone beds remaining almost undeformed (Lena et al., 2015). In short, phyllosilicates
within carbonates have a strong influence on 1) the fault rocks’ mechanical behavior due to the low
friction of clay minerals (Tesei et al., 2012), 2) deformation mechanisms during faulting, since clay
minerals have a strong effect on pressure-solution processes (Renard et al., 2001), and 3) fluid transfer
across and along faults due to the oriented structure of clay-dominated seams (Viti et al., 2014).

198
Chapter 6: Paleo-fluid circulation

High-porosity carbonates
Fault zones in high porosity granular carbonates and chalks behave similarly as fault zones in high
porosity sandstones. They are similarly characterized by cataclastic deformation bands with reduced
porosity, permeability, and grain size (Delle Piane et al., 2017). The thickness of these deformation
bands increases with increased displacement (Tondi et al., 2006, 2012), often resulting in
anastomosing or Riedel-like deformation band zones (Tondi, 2007). Microscale observations show that
deformation is accommodated by grain rotation and/or sliding, fracturing, and pore collapse, which
gives way to pressure solution at higher levels of deformation (Delle Piane et al., 2017). With increased
deformation the deformation processes are: 1) shear strain localization and compaction, 2)
development of discrete pressure solution seams in the compacted shear bands and pressure solution
at grain contacts, 3) shearing of the pressure solution seams and 4) cataclasis along the sheared
pressure solution seams (Cilona et al., 2012).

Earthquake-slip indicators
Pseudotachylytes, solidified melts produced by frictional heating during an earthquake-slip event
(Cowan, 1999), are extremely rare in carbonate hosted fault zones (Viganò et al., 2011; Delle Piane et
al., 2017). This is possibly due to the de-carbonation temperature of calcite and dolomite (<700 °C;
Samtani et al., 2002), being significantly lower than the melting temperature (>1000 °C; Irving and
Wyllie, 1975). The heat released by co-seismic slip induces thermal decomposition of the carbonates,
which probably plays an important role in the frictional resistance of faults within carbonate host rock
during slip (Han et al., 2007a; b; Sulem and Famin, 2009).

199
Chapter 6: Paleo-fluid circulation

given rock can sustain before fracturing (Figure 6 - 4a). This failure envelope is, in its simplest form,
defined by three rock parameters: 1) Tensile strength (T), 2) the cohesive strength (C), and 3) the
coefficient of internal friction (μ = tan φ). Failure of a rock occurs as the Mohr circle touches the failure
envelope, with the point of intersection defining the double angle (2θ) at which the fracture forms.
Note that, due to the symmetry of the failure envelope, a conjugate fracture set can develop upon
failure of the rock, with opposing shear sense on the conjugate fractures.

A B τ Tensile failure
τ

σ1 Pf
σ3 σ1 σn
φ σ’3 σ’1
Coulomb θ Pf
tan
criterion + σ n φ σ’n = σn - Pf
C
τ=
criterion σn σ3
C C τ Shear failure
τ
2θ Pf
σn
T
σn
σ’3 σ’1 σ3 σ1

Pf
σ’n = σn - Pf

Figure 6 - 4: Mohr diagram and failure envelope. A: Mohr diagram with Coulomb-Griffith failure envelope and
modes of failure. B: Tensile failure under elevated fluid pressure Pf. C: Shear failure under elevated fluid pressure
Pf. Modified from Fossen (2010). Abbreviations; T = tensile strength of the rock, C = cohesive strength of the rock,
τ = shear stress, σn = normal stress, φ = angle of iternal friction of the rock, θ = angle between fracture -normal
and σ1, Pf = fluid pressure, σn’ = effective normal stress.

Isotropic, intact rocks


Because true tensile stresses do not exist in the Earth, two conditions must be met for tensile fractures
to develop (Twiss and Moores, 2007): 1) (pore) fluid pressure must be sufficiently large to shift the
effective minimum principal stress into the tensile domain (Figure 6 - 4b), and 2) differential stress
must be small enough for extensional fractures to form, rather than shear fractures (the Mohr circle
has to touch the failure envelope within the tensile domain as opposed to the shear domain, such as
in Figure 6 - 4c).

Pure extensional fractures develop when the Mohr circle touches the failure envelope at the point of
maximum tensile stress T. The fracture plane experiences zero shear stress (τ = 0) and an effective
tensile normal stress of T = σ3 (Figure 6 - 4a). The common occurrence of veins associated with pure
extensional fractures indicates that the specific conditions of elevated fluid pressure in combination
with small differential stresses are often met in the Earth’s crust (Bons et al., 2012).

201
Chapter 6: Paleo-fluid circulation

Hybrid fractures or oblique extension fractures develop when the Mohr circle touches the failure
envelope within the tensional domain of T < σn < 0 (Figure 6 - 4a). The plane of failure experiences a
non-zero shear stress and an effective tensile normal stress. The resulting conjugate fracture set
develops at a small angle to each other, due to the increase in the slope of the failure envelope. Note
that pure extensional fractures are effectively a special case in which the conjugate sets combine into
a single fracture. The angle between conjugate shear fractures sets gives a direct determination of the
paleo-stress orientation, since σ1 lies in the bisector of the acute angle, σ2 in the intersection, and σ3
in the bisector of the obtuse angle of the conjugate fracture sets (Angelier, 1984; Hancock, 1985).

When the Mohr circle touches the failure envelope in the compressional domain (σn > 0), the plane of
failure experiences a non-zero shear stress and an effective compressional normal stress (Figure 6 -
4a). For progressively lower angles of internal friction the Coulomb fracture criterion line becomes
progressively less inclined and approaches horizontal when the internal friction angle approaches zero.
This indicates that, according to the Coulomb fracture criterion, fractures that initiate in intact,
isotropic rocks cannot form at angles higher than 45° to σ1.

Anisotropic, fractured rocks


Isotropic, intact rocks are scarce in nature, with layered sequences, bedding contacts, cleavage,
existing fractures and veins all giving rise to anisotropic conditions. Each potential plane of failure will
have its own failure envelope, rather than a single envelope for the rock as a whole.

Layered carbonate sequences are often characterized by competent (i.e. limestone) and less
competent (i.e. clays or marls) beds. Upon burial these rocks often experience layer-parallel extension,
resulting in a smaller differential stress (and thus smaller Mohr circle) for incompetent than for
competent beds, whose layer-normal stress (σ1) is the same (Figure 6 - 5a; Bons et al., 2012). This
implies that the competent beds fail first and, depending on the fluid pressure, develop extensional,
hybrid, or shear fractures. If the fractures open in a dilatational manner, they provide precipitation
sites for vein minerals (Bons et al., 2012). Upon layer-parallel compression the incompetent beds are
still characterized by a smaller Mohr circle compared to the competent beds, with the same layer-
normal stress, which is σ3 in this case (Figure 6 - 5b). Depending on fluid pressure this will again develop
shear or mixed fractures in the competent beds. Note that due to the coalescence of σ3 for the
competent and less competent beds, pure extension fractures, which may be subsequently filled with
vein minerals, can develop parallel to the bedding in both the competent and incompetent beds (Bons
et al., 2012).

202
Chapter 6: Paleo-fluid circulation

A Layer-parallel extension B Layer-parallel shortening

τ τ

competent competent

incompetent
σn incompetent σn
σ1 σ1 σ3 σ3
Figure 6 - 5: Mohr diagram with failure in anisotropic layered rocks. A: Layer-parallel extension in a sequence of
competent and incompetent lithologies. B: Layer-parallel shortening in a sequence of competent and
incompetent lithologies. Modified from (Bons et al., 2012).

Once a rock fails, the fracture represents a plane of weakness. Whether or not such a preexisting
fracture is reactivated, depends on the fracture’s orientation and friction coefficient together with the
stress field (Fossen, 2010). A non-cemented fracture has lost its cohesion (C = 0) and has its own
frictional sliding criterion envelope (based on Amontons’ second law of friction), which is different
from the intact-rock failure envelope (Figure 6 - 6; Twiss and Moores, 2007). The preexisting fracture
is reactivated if the point on the Mohr circle that corresponds to the double (2θ) angle of the fracture-
normal stress (σn) and the applied σ1 touches the frictional sliding failure envelope. If any part of the
Mohr circle touches the intact-rock failure envelope first, then new fractures are formed. In the case
of faults with a very low friction (μ = 0.1 for some parts of the SAF; Carpenter et al., 2011, 2012) the
shallow dip of the reactivation envelope permits reactivation of fault planes at very high angles to the
applied maximum principal stress. For faults where σ1 is at a high angle to the trace of the fault, such
as the SAF in California (Zoback et al., 1987, 2011), Median Line/Rokko Awaji segment in Japan (Famin
et al., 2014), or the EPGFZ in Haiti (section 4.2), a low friction coefficient, combined with a relatively
low differential stress and/or elevated fluid pressure, could explain why these faults remain active.

A Stable condition B Failure along pre-existing fracture

σ1 σ1
n
τ rio Plane of τ
crite weakness θ Reactivation θ
re
ra ctu σn σn
f σ3 σ3
mb
ulo n τ
Co terio
din g cri
io nal sli
Frict
2θ σn 2θ σn

Figure 6 - 6: Frictional sliding along pre-existing plane of weakness. A: insufficient effective differential stress for
failure; stable condition. B; sufficient effective differential stress for failure on pre -existing fracture. Note that if
the angle between σ1 and the plane of the pre-existing fracture becomes large enough, the Mohr-circle will touch
the Coulomb failure envelope before reactivation occurs, thereby promoting the creation of new fractures.
Modified from Twiss and Moores (2007) and Fossen (2010).

203
Chapter 6: Paleo-fluid circulation

Stepped and shear veins


Displacement on fractures containing steps or jogs results in stepped and shear veins (Bons et al.,
2012). Stepped veins occur when the displacement vector is non-parallel to certain segments of the
fracture, which can lead to the opening of small pull-aparts along the fracture, resulting in diamond-
shaped veins connected by a thin vein or fracture (Bons et al., 2008). Shear veins are a special case of
stepped veins, where increased displacement and a slight dilatation along the fracture results in
slickenfibres forming on small dilatant jogs (Passchier and Trouw, 2005). Slickenfibres consist of tiled
sheets of vein-filling minerals that are separated by secondary minerals, inclusions, or wall-rock
fragments (Bons et al., 2012). Macroscopically these minerals appear composed of fibrous crystals, but
at microscopic scales they are mainly blocky, resulting from growth competition into the dilatant
fracture space (Koehn and Passchier, 2000).

6.2.2 Vein morphology


Crystal shape

The crystal shape of a mineral can be used to determine the formation history of the vein, as in the
classification below. If a crystal shape is well developed and easily recognizable the crystal is termed
euhedral, if no well-formed crystal faces exists it is termed anhedral, and the intermediate case is
termed subhedral (Passchier and Trouw, 2005). The size of a crystal can be classified as
macrocrystalline, microcrystalline or cryptocrystalline (Bates and Jackson, 1987).

Blocky crystals are equant grains, commonly associated with growth into an open fluid-filled space
(Bons et al., 2012). Elongate blocky crystals (Fisher and Brantley, 1992) are formed when crystals
undergo growth competition, for instance when growing from the vein wall into the open space. This
leads to elongate crystals with a crystallographic preferred orientation (CPO) at a high angle to the vein
wall (Bons et al., 2012). Fluid inclusion studies on blocky calcite veins document a range of
homogenization temperatures from 50 – 270 °C and salinities of almost zero to 34 wt.% equivalent
NaCl (Muchez et al., 1995; Tritlla et al., 2001; Grandia et al., 2003; Bons et al., 2009; Wiltschko et al.,
2009; Morad et al., 2010). The lack of twinning in blocky quartz makes them better suited for fluid
inclusion studies than calcite, with documented temperatures ranging from 50 – 450 °C and variable
salinities (Boullier and Robert, 1992; Cox, 1995; Munz et al., 1995). Fibrous crystals have very large
length/width ratios, normally smooth vein boundaries, and show a minor increase in fiber width in the
growth direction (Bons and Montenari, 2005). This results from growth in a very limited space without
growth competition, where the vein opening rate is lower than the growth rate of the slowest crystal
(Bons et al., 2012). This contrasts with stretched crystals, which can show broad morphological

204
Chapter 6: Paleo-fluid circulation

similarities with fibrous crystals, but result from a different growth mechanism (Bons, 2000) and are
often characterized by serrated boundaries known as radiator structures (Bons et al., 2012). Stretched
crystals grow by crack-sealing mechanisms, where the crack surface can cut through the previously
precipitated mineral (localized stretched crystals) or through the wall rock (delocalized stretched
crystals), followed by precipitation within the newly formed crack (Bons et al., 2012). Cryptocrystalline
quartz is referred to as chalcedony, which is often characterized by a feathery or flamboyant texture
(Sander and Black, 1988; Dong et al., 1995). Fluid inclusions in chalcedony are generally too small for
micro-thermometric analysis (Bodnar et al. (1985), in Sander and Black, 1988).

Syntaxial veins

In syntaxial veins the minerals grow from the wall rock inwards (Bons et al., 2012). This typically occurs
from both walls towards the center, but can also be asymmetrical, from one side to the other (Fisher
and Brantley, 1992). Syntaxial veins can form by the creation of space from a single cracking event
(Figure 6 - 7; Wilson, 1994). During the initial stages of mineral growth, growth competition can result
in the creation of elongate blocky crystals. If the crystals grown into an open fluid-filled space,
nucleation of new crystals on the growth front suppresses the elongate shape and produces a blocky
texture (Figure 6 - 7; Bons et al., 2012). Syntaxial veins tend to collapse due to overburden or tectonic
stresses when fluids leave the crack and the veins are not fully mineralized when they are not fully
mineralized, resulting from overburden pressure or tectonic stresses (Fisher and Brantley, 1992; Fisher
et al., 1995). Any imperfectly sealed crack may act as a plane of subsequent failure and mineral growth
in multiple crack-seal events (Holland and Urai, 2010). Individual mineral growth versus crack-seal
events can often be inferred from inclusion bands, containing particles included in the vein growth
front, or visualized using cathodoluminescence (Koehn and Passchier, 2000; Renard et al., 2005). If the
minerals at the vein wall lock onto asperities, this provides the lateral offset of one vein wall to the
other (Bons et al., 2012).

Stretching veins

Stretching veins form by repeated crack-seal events, where the plane of cracking is not always the
same one (Figure 6 - 7; (Ramsay, 1980; Lee and Wiltschko, 2000; Renard et al., 2005). This contrasts
with syntaxial veins, for which by definition crack-sealing must occur at the mineral growth front (Bons
et al., 2012). Stretching veins do not experience growth competition, since there is no systematic trend
in the location of the crack-and-growth front, resulting in predominantly blocky textures (Bons et al.,
2012). Due to the lack of a systematic location for newly precipitated minerals, the shape of stretched

205
Chapter 6: Paleo-fluid circulation

crystals can only be used to determine the total relative displacement of the vein walls (Bons et al.,
2012). Even though stretching and syntaxial veins are different by definition, both crystal growth styles
can occur in a single vein, often leading to highly complex local vein morphology (Cervantes and
Wiltschko, 2010).

Figure 6 - 7: Scheme linking vein opening mechanism (antitaxial, syntaxial or stretching) and crystal morphology
to the growth plane and number of crack-seal events. Crystal types; b) blocky, c) elongate blocky, d) stretched
crystals and e) fibrous. From Bons et al. (2012).

Antitaxial veins

Antitaxial veins are characterized by a median line (Durney and Ramsay, 1973) or median zone (Oliver
and Bons, 2001; Bons and Montenari, 2005) from which fibrous crystals grow outwards, resulting in
two growth planes located on the interface between wall rock and vein (Figure 6 - 7; Bons et al., 2012,
and references therein). The median zone, which is the seed of the antitaxial vein crystals, originates
as a thin crack filled by blocky, elongate blocky, or stretched crystals, and can show signs of multiple
crack-growth events (Oliver and Bons, 2001; Bons and Montenari, 2005). Once the growth mechanism
of the vein switches to antitaxial fibrous growth brittle failure no longer occurs (Bons et al., 2012).
Antitaxial veins are well suited for tracking the total opening trajectory of a vein (Bons et al., 2012).

6.2.3 Mechanisms of vein deformation


The natural samples treated in this work are limited to temperature and pressure conditions resulting
from a few km’s of sedimentary overburden. The associated low-grade metamorphic P-T conditions
limit the number of deformation mechanisms that can affect the veins and minerals. Only low-
temperature low-pressure mechanisms are listed in the section below.

206
Chapter 6: Paleo-fluid circulation

Calcite and dolomite veins

Calcite deforms by cataclastic flow and fracturing under very low-grade conditions (Kennedy and
Logan, 1998; Billi, 2010). Larger calcite fragments are strongly twinned and cut by stylolites and smaller
veins, while smaller fragments can remain twin- and strain free (Passchier and Trouw, 2005). Brittle
deformation may be assisted by twinning, solution transfer and, for finer grains, dynamic
recrystallization and dislocation glide (Wojtal and Mitra, 1986; Kennedy and Logan, 1998; Kennedy and
White, 2001; Molli et al., 2011). In the presence of water under low-grade conditions, stylolites may
develop due to pressure solution (Burkhard, 1990; Kennedy and Logan, 1997, 1998). Deformation
twinning in calcite can develop from diagenetic conditions onwards (Schmid et al., 1981) along the
three e-planes and can occur at very low shear stresses, between 2 – 12 MPa (Turner et al., 1954;
Burkhard, 1993; De Bresser and Spiers, 1997). Because the amount of strain that can be
accommodated by twinning is limited, twinning is achieved by grain boundary sliding, grain boundary
migration and pressure solution (Passchier and Trouw, 2005). Twins are used as indicators of
temperature (Burkhard, 1993; Ferrill et al., 2004), strain (Groshong, 1972), and stress (Parlangeau et
al., 2018). Dolomite is different from calcite, with deformation twinning only occurring above 300 °C
and on f-planes within the crystal. Because dolomite is stronger than calcite at low-grade conditions,
dolomite layers might develop boudinage in a calcite matrix (Passchier and Trouw, 2005).

Quartz veins

At conditions below 300 °C, pressure solution and brittle fracturing are the dominant deformation
mechanisms (Dunlap et al., 1997; Van Daalen et al., 1999; Stipp et al., 2002). The structures that
develop are typically kink bands, undulose extinction, and fractured grains (Nishikawa and Takeshita,
1999), with pressure dissolution along stressed grain faces, transfer and subsequent deposition of
quartz in adjacent veins (Passchier and Trouw, 2005). Dynamic recrystallization may occur in strongly
deformed quartz at very low-grade conditions (Wu and Groshong, 1991). The slip systems and glide
planes that are active within the quartz crystal depend on the differential stress, temperature, strain
rate and presence of water along grain boundaries or within the crystal lattice (Passchier and Trouw,
2005).

207
Chapter 6: Paleo-fluid circulation

6.3 Methodology
Sampling techniques
The orientation of veins and their kinematic criteria are recorded at the outcrop using a Brunton Geo
Pocket Transit compass, which is set at N10°W declination to account for the average magnetic
declination of southern Haiti (https://ngdc.noaa.gov/geomag-web/). From the outcrop oriented hand
samples are collected that preferably contain both vein and host rock, although in some instances only
the vein or the host rock could be sampled. The samples are cut using a diamond coated circular saw
as perpendicular to the trend of the vein as possible. For veins with mode II/III opening and observable
slip direction, samples are cut both perpendicular to the vein and parallel to the displacement. One of
the cut-halves is reserved for thin section manufacturing performed in-house at IFPEn. Standard 30 x
45 x 30μm thin sections are made by first cutting a cube from the cut-half. The mirror-half of this cube
is used to create fluid inclusion wafers, if necessary. The material left of this cube is to sample specific
veins for stable oxygen and carbon isotope or X-Ray Diffraction (XRD) analysis. This allows a direct
correlation between the isotope, XRD and fluid inclusion microthermometry results, and their
projected sample location on the associated thin section.

Analytical techniques
Petrography of the thin sections is studied using conventional transmitted light microscopy with plane-
polarized light (ppl) and cross-polarized light (xpl). Cathodoluminescence microscopy (CL) is used to
better constrain different vein generation and vein growth. The CL microscope used is a Nikon Eclipse
ME600 equipped with a CITL Mk5 optical stage, connected to a Cathodyne OPEA system operated at
16 – 20 kV at 420 – 660 μA. Exposure time and light settings of the camera are kept constant for all
samples to allow a direct visual comparison of the results.

Double polished wafers with 30 x 45 x 100μm dimensions are made by Thin Section Lab, France, and
used for fluid inclusion microthermometry studies. These wafers are bathed in acetone for around 12
hours to detach the wafer from the glass plate. A first scan for fluid inclusions is conducted next, after
which the wafer is cut into smaller ~1 cm2 chips. These chips are analyzed in a Linkam Pr 600 heating-
freezing stage mounted on a Nikon LV100 Eclipse microscope. The Linksys32 software package is used
to conduct the analysis. The accuracy of the microthermometry data are about ±1 °C for heating runs
and ±0.5 °C for cooling runs. Petrography of the inclusions is studied first to determine if they are
primary or (pseudo-)secondary. Size and filling factor of the inclusions is determined next, which is
calculated by extracting the surface area of the gas bubble (B) from the surface area of the inclusion
(I), and subsequently dividing by the inclusion area (I) according to the formula F = (I – B)/I.

208
Chapter 6: Paleo-fluid circulation

Homogenization temperature of the fluid inclusions is studied first. The sample is incrementally heated
until the vapor bubble has almost disappeared. Minor heating and cooling is applied to constrain the
homogenization temperature (Th) of the inclusion. Upon cooling the temperature at which the gas
bubble nucleates after homogenization is recorded (Tngas). A second run is performed to test the
consistency of the homogenization temperature of the inclusion. This process is repeated for every
fluid inclusion in the chip, progressing from the lowest homogenization temperature inclusion to the
highest temperature. The sample is subsequently frozen at two incremental steps; room temperature
to -25°C at 35°C/min, followed by 5°C/min from -25°C to -60°C, after which this temperature is held
for 1 minute to allow complete freezing of the fluid inclusion. The sample is subsequently heated in
five incremental steps; -60 to -30 °C at 10°C/min, -30 to -20 at 5°C/min, -20 to -5 at 2°C/min, -5 to -2 at
2°C/min, and -2 to zero at 0.3°C/min. During this heating cycle the temperature is manually held at
times to better observe the behavior of the fluid inclusion. The first appearance of liquid is
documented, which is the upper constraint for the real eutectic temperature of the fluid system
(Teapparent), as is the final ice melting temperature (Tmice) of the inclusion.

Stable carbon and oxygen isotopic analyses were performed on 30 to 40 μg of carbonate powder
extracted from veins. For each sample, veins were hand drilled using a dental milling end to avoid
mixture with host rocks. Carbonate samples were placed in glass vials and carbon dioxide was
extracted on carbonate powder after dissolution with dehydrated phosphoric acid under vacuum at
70°C necessitating a correction for dolomite samples according to Rosenbaum and Sheppard [1986].
The δ13C and δ18O analyses were performed using a Kiel IV carbonate automated preparation device
coupled to a DELTA V isotope ratio mass spectrometer at ISTeP laboratory from Sorbonne Université
(Paris, France). Isotope values are reported in delta notation relative to Vienna Peedee Belemnite
(VPDB/PDB). Repeated analyses of a marble working internal standard (calibrated against the
international standard NBS-19) indicate an accuracy and precision of 0.05‰ and 0.1‰ (1σ)
respectively. All stable oxygen and isotope values report in the following chapters are relative to the
VPDB standard, unless otherwise stated.
Vein and host-rock mineralogy of selected samples is further analyzed with X-ray diffraction (XRD),
using a D2 Phaser Bruker at the ISTeP laboratory. This requires only a few 10’s of μg of material, making
it a very suitable tool to investigate small veins.

The geochemistry of the liquid and vapor phase is examined using a Renishaw inVia confocal Raman
microscope at ENS in Paris.

Whole-rock major, minor, trace and REE geochemistry for selected samples is analyzed by Bureau
Veritas in Vancouver, Canada. The full set of results can be found in Appendix 5.

209
Chapter 6: Paleo-fluid circulation

6.4 Results from the Chaîne des Matheux


The tectonic history of the Chaîne des Matheux is summarized in section 5.8, the chronology of the
vein sets is given in Figure 6 - 8 and briefly summarized below, with the sample locations shown in
Figure 6 - 9. The Chaîne des Matheux is addressed prior to the Southern Peninsula, because 1) the
deformation is better constrained, both temporally and spatially, and 2) the sample density is higher.

Orthogonal, bedding perpendicular joints were created during middle Miocene, probably Serravallian
times (section 5.8). These joints are grouped into two sets. Set 1a consists of systematic E – W to NE –
SW joints that are occasionally mineralized, which likely resulted from layer-parallel shortening. Set 1b
consists of N – S to NW – SE cross joints that are abutting against the joints of set 1a and are thus
younger (Figure 6 - 8a).
A σ1/2
~ Serravallian σ1/2 Set 1b N Set 1a
σ3 σ3
σ3 σ3

σ1/2
B ~ Serr. - e. Tort. σ1 N σ1
Set 2
σ3 σ3
σ3 σ3

σ1
C ≤ Tortonian σ3 N Set 3
σ3

σ1 σ1 σ1 σ1

σ3
D ≤ Messinian ? σ3 N Set 4 σ3
σ1 σ1
σ1 σ1

σ3

E σ2 Set 5a
~ Quaternary Set 5b N σ2

σ1 σ1 σ3 σ3
σ1

σ2

Figure 6 - 8. Chronology of deformation and associated veins in the Chaîne des Matheux. Left; schematic
representation of deformation and veins within a multilayered volume of rock. Grey are marly / less competent
intervals, white are limestone / competent intervals. Center left; dominant orientation of structures active during
this time. Center right; schematic Mohr-diagram with Coulomb-Griffith failure criterion for the rock. Right;
fracture and vein development and associated stresses.

210
Chapter 6: Paleo-fluid circulation

Sets 3 and 4 (reactivated bedding and thrust) veins

Reactivated bedding planes and upwards ramping thrusts that originate from the reactivated bedding
are common structural features in the Chaîne des Matheux (section 5.4). Even though the outcrop-
scale thrusts are often rooted into reactivated bedding intervals, it is difficult to establish if a relative
timing exists between them, and both are attributed to the same thrust-related phase of deformation.
Associated veins are commonly observed on the reactivated bedding within dilatational sites (Figure 6
- 11a), or along dilatational jogs along the upwards ramping thrust faults. Even though set 3 and set 4
veins have a different timing, with set 3 resulting from a NE – SW contractional regime and set 4
resulting from a slightly younger NW – SE contractional regime, the similarity in tectonic setting and
vein morphology allows both sets to be treated in this section.

A Outcrop B Thin section C


counts XRD ~2cm
3000
CaCO3
2000

1000

0
10 20 30 40 50 60 70
°2Θ CuK α

Dissolution
bands

~75cm 2 mm
L. - m. Eocene

D PPL

Vein calcite growth direction

Host rock
platform
limestones
Crack-seal
events ?
2 mm

E XPL

2 mm

Figure 6 - 11: Reactivated bedding vein (set 3a). For location see Figure 6 - 9.

Set 3 and 4 veins normally range from mm- to cm-scale thickness, attaining a maximum thickness of 3
cm (set 3, Figure 6 - 11a). Noticeable from the thin section are the dark wavy bands sub-parallel to the
bedding (Figure 6 - 11b). These bands contain dark, insoluble clays and oxides, which at high
magnification show a jigsaw-pattern (Figure 6 - 11c). In the vicinity of the bands the calcite is

213
Chapter 6: Paleo-fluid circulation

6.4.2 Stable oxygen and carbon isotopes


Host rock samples (n = 17) from the Chaîne des Matheux have a narrow range of δ 18O values from
-1.99 to -4.25‰, while the range of δ13C values is wider, from +2.58 to -9.72‰. These values are mainly
located outside of the global range for Late Cretaceous and Cenozoic marine carbonates (Figure 6 -
14a; Prokoph et al., 2008). This global range of δ18O values is similar to the range of surface lake waters
in Hispaniola over the past 10000 years (+2 to -2‰ δ18O; Hodell et al., 1991; Lane et al., 2011).

A Host rock ∂18O vs ∂13C B Vein sets


4 4

2 2 C
0

∂13C (‰VPDB) calcite


∂13C (‰VPDB) calcite

0
-2
-2
-4
A

-4 -6
D
-6 -8
B
-10
-8
-12
-12 -10 -8 -6 -4 -2 0 -12 -10 -8 -6 -4 -2 0
∂18O (‰VPDB) calcite n = 17 n = 29 ∂18O (‰VPDB) calcite
C ∂13C vein vs host rock D ∂18O vein vs host rock
4

10
=-
∂18O Hos rock (‰VPDB) calcite

0
2 roc
k
C ‰
∂13C Host rock (‰VPDB) calcite

st -5 C
ho -2 ck=
0 ei n- ‰ s t ro
CV 5 o
13 =- -h
Δ
oc
k -4 ein A B
-2 o st r 18 OV D
-h Δ
B ein -6 0‰
-4 CV k= ‰
+5
13
Δ r oc
‰ -8 ost ck=
-6 =0 A 5‰ -h
os t ro
roc
k =+ ein -h
st rock
-10 18 OV in
-8 - ho ost Δ
O Ve
ein -h 18
Δ
CV ein
-12
CV
13
-10 Δ 13
-12 -10 -8 -6 -4 -2 0
D Δ
n = 26 ∂18O Vein (‰VPDB) calcite
-12
-12 -10 -8 -6 -4 -2 0 2 4 E Age of host rock vein
n = 26 ∂13C Vein (‰VPDB) calcite 4

2 C
Age host rock; Fluid inclusion I.D.
Miocene A = set 3 0
∂13C (‰VPDB) calcite

Eoc. - Oligo. B = set 4


C = set 1a -2
Vein types; D = set 3 A
-4
Set 1a/b (joints) Set 4 (thrust-related)
-6
Set 2 (normal) Set 5a/b (strike-slip) D
Set 3 (thrust-related) Karst Unknown -8

-10
B
∂18O and ∂13C of Late Cretaceous and
Cenozoic carbonates (Prokoph et al., 2008)
-12
Bulk mantle (-5.0‰ ∂13C; Hoefs, 2009) -12 -10 -8 -6 -4 -2 0
∂13C mixing trend of vein samples ∂18O (‰VPDB) calcite
n = 29
Figure 6 - 14: Stable oxygen and carbon isotope data for veins and host rocks. A: Host rock δ 18O vs δ13C. B: Vein
sets δ18O vs δ13C. C: δ13C vein vs host rock. D: δ18O vein vs host rock. E: Age of the host rock vein δ 18O vs δ13C.
Vein stable isotope data labelled A, B, C and D correspond to the fluid inclusion examples A, B, C and D in Figure
6 - 15 and section 6.4.3.

216
Chapter 6: Paleo-fluid circulation

A strong depletion towards negative δ13C is especially pronounced in the Eocene host rocks (Figure 6 -
14a). Host rock samples often show strong dissolution and/or recrystallization (Figure 6 - 11) or
brecciation and re-cementation (Figure 6 - 12). This may significantly reset the original host rock
isotope signal with that of the fluids responsible for the crystallization/re-cementation. This can in turn
strongly influence the vein versus host rock values for both stable carbon and oxygen. Rather than
trying to quantify the amount of resetting of the host rock, a stronger emphasis is placed on the isotope
signal of the calcite veins when trying to understand the paleo-fluid circulation.

Vein calcite (n = 29) from the Chaîne des Matheux has a narrow range of δ18O values from -0.96 to
-5.79‰, while the range of δ13C values is wider, from +3.37 to -12.63‰ (Figure 6 - 14b). Vein calcite
from set 1a/b bedding perpendicular joints have positive δ13C values that are typical for Late
Cretaceous and Cenozoic carbonates (Figure 6 - 14b; Prokoph et al., 2008). The same holds for δ13C
values of set 2 calcite veins along normal-shear fractures, although these display more negative δ18O
at around -6‰ (Figure 6 - 14b). The δ13C values of the veins plot on a mixing line towards negative
values (Figure 6 - 14b). Set 3 thrust-related veins show a decreasing trend from -2 to -6.5‰ δ13C over
a narrow range of δ18O (Figure 6 - 14b). Set 4 thrust-related veins plot between -9 and -12‰ δ13C, and
also display a narrow range of δ18O values (Figure 6 - 14b). The sample obtained from a set 5b vein
along a strike-slip fault is also characterized by a low -10‰ δ13C value. This sample is also slightly more
depleted at -5.5‰ δ18O (Figure 6 - 14b). The lowest δ13C values are reported for veins related to karst
processes, with values as low as -12.6‰ (Figure 6 - 14b).

Calcite that precipitated in set 1a and 1b joints is slightly enriched up to +3‰ Δ13C with respect to the
host rock (i.e. sample C; Figure 6 - 14c). Calcite from set 2 normal-shear fractures display a similar
behavior, albeit with an enrichment up to +5‰ Δ13C (Figure 6 - 14c). Thrust-related veins of set 3 range
between +2 to +7 Δ13C disequilibrium with the host rock, while thrust-related veins of set 4 have
disequilibria ranging from -4 to -7‰ Δ13C (Figure 6 - 14c). One sample of a set 5b strike-slip related
vein on the southern side of the Chaîne des Matheux is strongly depleted with respect to the host rock,
up to -8‰ Δ13C (Figure 6 - 14c). The largest negative carbon isotope disequilibrium value is found for
veins related to karstification processes, which is -10‰ Δ13C with respect to the host rock (Figure 6 -
14c). Stable oxygen isotope disequilibrium between vein calcite and host rock is limited between +2
and -4‰ Δ18O (Figure 6 - 14d).

The maximum age of the veins is constrained by the age of the host rock, which is shown in Figure 6 -
14e. There is no correlation between age of the host rock (and thus maximum age of the vein) and the
δ18O or δ13C values of the vein calcite.

217
Chapter 6: Paleo-fluid circulation

All the inclusions presented in Figure 6 - 15a are filled with clear liquids, without any signs of
hydrocarbon within the inclusions. The inclusion in Figure 6 - 15a were tested using Raman
spectroscopy to try and identify the composition of the vapor phase. No signal other than calcite
(CaCO3) and water (H2O) could be identified.

The filling factor (F) for the inclusions, calculated prior to any heating or cooling cycles, ranges from
0.87 to 0.98, with an asymmetric distribution towards the higher filling factor (Figure 6 - 15b).

Only one poorly defined eutectic temperature is found for sample B, which is constrained at around
-25 °C (Figure 6 - 15a). Such a temperature is indicative for the presence of NaCl within the system,
and possibly minor other salts (Goldstein and Reynolds, 1994). Two inclusions from this sample B gave
a final melting temperature of -0.3 °C (Figure 6 - 15a). Assuming that the system only contains H2O and
NaCl and using the correlation between final ice melting temperature and weight percentage NaCl
from Bodnar (1993), results in a very low salinity of 0.5 wt. % NaCl for sample B.

The homogenization temperatures for the inclusions from veins in the Chaîne des Matheux range from
138.4 to 82.5 °C, with the majority binned between 80 and 120 °C (Figure 6 - 15c). Sample C from a set
1a calcite vein (Figure 6 - 9) has a homogenization temperature of 128 °C. Sample A from a set 3 calcite
vein located along the southern front of the Chaîne des Matheux (Figure 6 - 9) has a homogenization
temperature of 138 °C. Sample B from a set 4 vein, collected at the same outcrop as sample A (Figure
6 - 9), and has homogenization temperatures ranging from 98 to 118 °C. Sample D from a set 3 calcite
vein, collected 2 km northwest of samples A and B (Figure 6 - 9), has homogenization temperatures
ranging from 82 to 104 °C. It is interesting to note the ~45 °C difference in homogenization
temperatures for set 3 vein samples A and D collected at locations just 2 km apart in the same
formation and located along the same thrust fault (Figure 6 - 9). The spread in temperatures possibly
signals a difference in the timing of vein precipitation, for instance during progressive uplift of the
Chaîne des Matheux, or could signal a difference in fluid source.

A plot of calcite δ18O (VPDB) values on oxygen isotope fractionation curves calculated after Kim and
O’Neil (1997) versus homogenization temperature (Figure 6 - 15d) shows that the δ18O (SMOW) of the
parent fluids from which the calcite precipitated ranges from 8 to 18‰. These values mainly fall into
the domain of formational waters for samples A, B and C, with a slight overlap towards meteoric waters
for sample D (meteoric and formational water range from Hoefs, 2009).

219
Chapter 6: Paleo-fluid circulation

6.4.4 Whole-rock geochemistry


A total of 5 veins and associated host rock samples from the Chaîne des Matheux were analyzed for
their whole-rock composition, of which the REE and Yttrium (REY) data is plotted in Figure 6 - 16. The
small number of samples is due to the large quantity of material (>5 grams) needed for analysis, which
limited the size of the veins that could be targeted. No data indicates that the REY concentration is
below the detection limit.
A Chaîne des Matheux: REY (PAAS normalized)
1
Host rock types
REY/PAAS

0,1

0,01

0,001
La Ce Pr Nd Sm Eu Gd Tb Dy Y Ho Er Tm Yb Lu
LREE MREE HREE
Miocene marly Miocene cherty Eocene platform
limestones limestones limestones

B 1
Veins
REY/PAAS

0,1

0,01

0,001
La Ce Pr Nd Sm Eu Gd Tb Dy Y Ho Er Tm Yb Lu
LREE MREE HREE
Set 1a/b Set 2 Set 3 (thrust- Set 4 (thrust- Set 5 Unknown
(joint) (normal) related) related) (strike-slip)
Vein
C 1
Veins and host rocks
REY/PAAS

0,1

0,01

0,001
La Ce Pr Nd Sm Eu Gd Tb Dy Y Ho Er Tm Yb Lu
LREE MREE HREE
Set 1a/b Set 2 Set 3 (thrust- Set 4 (thrust- Set 5
(joint) (normal) related) related) (strike-slip) Unknown
Vein
Host Rock

Figure 6 - 16: Rare earth elements + Yttrium (REY) data for samples from the Chaîne des Matheux, PAAS
normalized after Taylor and McLennan (1985). A: REY for host rocks. B: REY for veins. C: REY for veins and host
rocks.

220
Chapter 6: Paleo-fluid circulation

Host rock limestones (Figure 6 - 16a) have negative Ce, mildly positive Y anomalies, and a mild
enrichment in heavy REE compared to light REE, which is typical for carbonates that formed in a marine
environment (Bau et al., 1997; Bolhar et al., 2004; Bau et al., 2010; Piper and Bau, 2013). The Yttrium
anomaly for the Eocene platform limestones is however less pronounced than for the Miocene marly
and cherty limestones, the latter characterized by a strong positive Y anomaly (Figure 6 - 16a).

REY values for the vein calcite is lower than the host rock value. For all the veins a number of elements
fall below the detection limit, thereby limiting the usefulness of the data (Figure 6 - 16b).

Calcite veins of set 1a joints are more strongly depleted in REY compared to the host rock and have a
less pronounced Ce anomaly (Figure 6 - 16c), which could indicate mixing with allochthonous fluids in
a more open system. Alternatively, the more depleted REY signal of the veins could suggest more
limited fluid-rock interaction in a closed system, with REY elements preferentially retained in the host
rock over the fluid. This is probable since the host rock is a marly limestone that has significant
quantities of clay minerals, which can preferentially contain REY elements (Piper and Bau, 2013). This
relatively closed system hypothesis is preferred over the more open system, since 1) the set 1a calcite
veins are almost in isotopic equilibrium with their host rock (Figure 6 - 14), and 2) set 1a veins are
found along joints that are generally bed-bounded within the formation, resulting in predominantly
intra-formational pathways for paleo-fluid circulation.

Set 2 veins are only slightly depleted compared to the host rock and have a similar REY pattern,
characterized by a negative Ce and positive Y anomaly (Figure 6 - 16c). Veins characterized by a similar
REY pattern and only mildly depleted REY concentrations with respect to the host rock indicate a high
fluid-rock interaction (Nuriel et al., 2012) and an almost closed system.

For set 3 and set 4 thrust-related veins the slope from La to Ce is similar for the veins and for the host
rock, with the veins mildly depleted in the limited REY elements available (Figure 6 - 16c). Since these
veins are hosted by Eocene platform limestones with a very limited clay fraction, it is most likely that
the strong depletion in REY of the veins compared to the host rock signals paleo-fluid circulation in a
relatively open system. This is supported by the large disequilibrium in stable isotope values between
the set 3 and 4 veins and the host rock (Figure 6 - 14).

221
Chapter 6: Paleo-fluid circulation

Calcite veins related to karstification processes

Stable carbon isotopes from vein calcite display the widest spread of isotope values from +4 to -12.6‰.
All intermediate values are located on a mixing line between these two extremes (Figure 6 - 17). At the
lowest end are calcite veins related to karst processes close to the earth’s surface (section 6.4.1; Figure
6 - 17f). Low δ13C values can result from a number of processes.

One of these is contamination of the parental fluid, to various degrees, by hydrocarbons (Hoefs, 2009).
Calcite that precipitated from fluids contaminated by hydrocarbons are usually on a mixing line
towards very depleted δ13C values (Hudson, 1977; Budai et al., 1983), which can be as low as -32‰
δ13C (Ferket et al., 2003; Beaudoin et al., 2011), although positive δ13C values are also reported (Poros
et al., 2012). These strongly depleted values are significantly lower than δ13C values from samples in
the Chaîne des Matheux (Figure 6 - 17). Additionally, optical and cathodoluminescence microscopic
observations, Raman analysis on fluid inclusions, and the geochemistry of the fluids within these
inclusions, all lack evidence of hydrocarbons. Based on the δ13C isotope signal alone it can however
not be fully excluded that a minor fluid contamination by hydrocarbons occurred. Even though there
is no working hydrocarbon system in the Chaîne des Matheux, some minor traces of oil and gas are
reported from wells in the Cul-de-Sac Basin (Bourgueil et al., 1988).

Calcite veins related to karstification processes, which have the lowest δ13C values of all samples in this
study (Figure 6 - 17), point to another possible fluid origin. These veins formed proximal to the earth’s
surface (section 6.4.1). Other calcite precipitates that result from karst processes proximal to the
earth’s surface are speleothems (i.e. stalagmites and stalactites; Banks and Jones, 2012), related to
dissolution/precipitation processes between limestones and cold, less than 30 °C meteoric waters
(Lohmann, 1988). Carbon isotope values for stalagmites in limestone caves in Israel plot on a mixing
line, but mainly group between -10 and -13‰ δ13C (Bar-Matthews et al., 1997; Nuriel et al., 2012). In
the case of calcite precipitates in caves, such δ13C values are typically associated with precipitation
from meteoric fluids that carry dissolved CO2, which is derived from the respiration of C3-type
vegetation that supplies CO2 to waters infiltrating the soil zone (Lohmann, 1988; Bar-Matthews et al.,
1996, 1997, 2003; Baker et al., 1997; Nuriel et al., 2011, 2012; Ünal-Ïmer et al., 2016).

Another type of calcite that precipitates proximal to the earth’s surface are travertine deposits
resulting from hydrothermal activity (Pentecost, 2005; Banks and Jones, 2012). Travertine calcite from
hydrothermal systems generally have positive δ13C values (Shi et al., 2014; Frery et al., 2017; Capezzuoli
et al., 2018). Positive δ13C values do not match the -12.6‰ δ13C from the karst-related calcite veins.

Based on the geochemical, vein morphological, and structural evidence, it is most likely that the calcite
veins related to karstification processes precipitated from meteoric fluids carrying soil-derived CO2. If

223
Chapter 6: Paleo-fluid circulation

strongly depleted δ13C vein calcite is related to soil-derived CO2, it puts a constraint on the timing of
these veins. Prior to uplift of the Chaîne des Matheux in the Tortonian, this region was below sea level.
This means that prior to the Tortonian, there cannot be any input of soil derived CO2, because there
was no soil for C3 or C4 plants to grow on.

Sets 1a and 1b (joints)

These joints and veins result from tensile fracturing of the rock, which requires a horizontal σ3 axis
perpendicular to the joint or vein. The systematic set 1a joints probably resulted from bedding-parallel
shortening during the Serravallian, with the set 1b cross joints developing in response to strain
relaxation (section 6.4.1). The orthogonal bedding perpendicular set 1a and set 1b veins have the
highest δ13C values of all the calcite samples (Figure 6 - 17a). These veins are roughly in δ18O equilibrium
with the host rock and slightly elevated in δ13C (section 6.4.2), have a negative Ce anomaly in REY
(section 6.4.4), and display a non-luminescence similar to the host rock (6.4.1). This is most likely
related to significant fluid-rock interaction in a relatively closed fluid-circulation system, resulting in an
equilibration between the fluids and the host rock (section 6.4.4). The fluid temperature was roughly
130 °C and δ18O of the parental fluid is typical for formational waters (section 6.4.3).

This indicates that these calcite veins precipitated from intra-formation fluids that were not
contaminated by a meteoric or deeper high-temperature component.

Set 2 (normal-shear) veins

These set 1 joints are offset by younger normal faults and associated veins, which have a similar δ13C
value but are slightly depleted in δ18O compared to set 1 veins (Figure 6 - 17b). The normal faults and
fractures are possible related to the minor strike-slip episode that reactivated the joint sets during the
Serravallian to early Tortonian (section 6.4.1), prior to onset of thrusting. Set 2 calcite has similar
mineralogical properties as the joints, but are less depleted in REY, indicative for significant fluid-rock
interaction and equilibration between the fluids and the host rock (section 6.4.4). The δ18O (VPDB) of
the calcite is slightly depleted with respect to the host rock and also more depleted compared to set 1
veins. Such depletion could reflect slightly higher temperatures for the fluids compared to the host
rock (Ünal-Ïmer et al., 2016) and a slightly deeper source. The normal faults have a larger height than
the set 1 joints and occasionally offset multiple beds. This could create a pathway for warmer fluids to
ascend upwards through the stratigraphy.

224
Chapter 6: Paleo-fluid circulation

The similarity in geochemistry with the host rock indicates that these fluids are sourced from a deeper,
warmer reservoir, but with similar lithology as the host rock in which the veins are situated.

Sets 3 (reactivated bedding and thrust) veins

Reactivation of the bedding and upwards ramping thrusts with associated set 3 veins developed during
the phase of folding-and-thrusting in the Chaîne des Matheux from the Tortonian to the Pliocene
(section 5.8; Figure 6 - 17c). These structures with predominantly top-to-SW vergence offset set 1 joints
and set 2 minor normal faults. These veins are characterized by multiple crack-seal events indicating
that they formed during activity on the faults. The δ13C values for set 3 veins range from -3 to -7‰
(Figure 6 - 17c), they are enriched in δ13C with respect to the host rock (Figure 6 - 14c), but have a
similar cathodoluminescence intensity (section 6.4.1). The samples that are least depleted in δ13C have
the highest homogenization temperature (-3‰, 140 °C), while the most depleted samples have the
lowest temperatures (-7‰, 80 – 105 °C) (section 6.4.3). The -3‰ δ13C value of the high temperature
samples is higher than the -5‰ of the bulk mantle (Hoefs, 2009), which could thus result from
contamination of warm fluids with a mantle component. The low temperature samples with -7‰ δ13C
cannot result from deep fluids with a mantle- δ13C signature along. The low temperatures and low δ13C
values can be explained by the mixing of formational waters with meteoric, cold, low δ 13C surface
waters. This is supported by the mixed meteoric/formational δ18O signal of the parental fluids (section
6.4.3). The minor depletion in δ13C for the high temperature vein samples could also be explained by
a minor mixing of meteoric waters with deeper, warm, formational waters, without significantly
affecting the δ18O signal.

The decreasing temperature of the fluids and increase in the meteoric component is probably related
to uplift of the Chaîne des Matheux during the folding phase. Increased folding-and-thrusting increased
the fault and fracture density, which in turn resulted in increased pathways for cold, low δ13C meteoric
fluids to percolate into the system, thereby overprinting any signal from warmer, deeper fluids.

Sets 4 (reactivated bedding and thrust) veins

Set 4 thrust-related veins developed along thrust faults and reactivated bedding planes with
predominantly top-to-SE vergence, possibly from the Messinian to the Pleistocene, although their
timing is poorly constrained (section 5.8; Figure 6 - 17d). These veins are probably syn-thrusting due
to the crack-seal mechanisms controlling their creation (Figure 6 - 11), although the lack of post-
mineralization deformation could indicate that they precipitated towards the end of fault activity. The

225
Chapter 6: Paleo-fluid circulation

δ18O values for set 4 veins are similar to the set 3 veins, but the δ13C values are markedly lower at
around -9 to -12 ‰ (Figure 6 - 17d). These set 4 veins are in δ18O equilibrium, but in disequilibrium for
δ13C with the host rock (Figure 6 - 14). This indicates a fluid component that is different from the host
rock, but depleted in δ13C, which could be related to the infiltration of meteoric waters.
Homogenization temperatures of 100 – 120 °C are intermediate between the set 3 veins, with δ18O of
the parent fluids indicative for predominantly formational waters (section 6.4.3). This indicates that
the volume of meteoric water carrying soil-derived CO2 is relatively small compared to the total fluid
volume, but nonetheless large enough to lower the δ13C signal to around -11 ‰.

These meteoric waters likely percolated downwards along the fault zones along the southern front of
the Chaîne des Matheux (Figure 6 - 9).

Sets 5a and 5b (strike-slip) veins

The set 4 thrusts and associated veins are offset by transtensional strike-slip faults of set 5 along the
southern border of the Chaîne des Matheux (Figure 6 - 9 and Figure 6 - 17e). The -10‰ δ13C value of
calcite that precipitated during activity of these strike-slip faults is similar to set 4 veins, while they are
slightly more depleted in δ18O (Figure 6 - 17e). Activity of these transtensional strike-slip faults is
roughly constrained to the Quaternary (section 5.8). Calcite veins related to the Quaternary strike-slip
faults that affect the southern side of the Chaîne des Matheux, are in cathodoluminescence
equilibrium with the host rock, being non-luminescent. Similar to the thrust-related veins, this calcite
is roughly in δ18O equilibrium with the host rock, but strongly depleted in stable carbon at -9‰ Δ13C.
This indicates that set 5 veins precipitated from fluids that are not in equilibrium with the host rock in
a relatively open system.

A similar fluid source is inferred as for set 3 and set 4 veins, where formational fluids are strongly
contaminated by low δ13C cold meteoric surface fluids carrying soil-derived CO2 that percolated
downwards through the fault zone.

226
Chapter 6: Paleo-fluid circulation

6.5 Paleo-fluid circulation and evolution of the Chaîne des Matheux


The fluid source for the bedding perpendicular set 1 veins associated with the Serravallian set 1 joints
are probably intra-formational waters that circulated in a relatively closed system (section 6.4.5; Figure
6 - 18a). These fluids precipitated in the joints following tensile fracturing related to layer-parallel
shortening induced by far-field stresses from folding-and-thrusting in the Montagnes Noires further
north (section 6.4.5).

A ~Serravallian (set 1) B ~Serravallian - early Tortonian (set 2)


SW NE SW NE

C ~Tortonian - Pliocene (set 3) D ~Messinian (?) - Pliocene (set 4)


SW CdM Low T MN NE NW CdM SE
?

High T

E
~ Quaternary (set 5)
SW CdM MN NE

nice

low-T; low δ13C)


(medium-T; high δ13C)
(high-T; medium δ13C)

Figure 6 - 18: Schematic evolution of paleo-fluid circulation in the Chaîne des Matheux. All schematic sections
are oriented NE – SW, apart from section D which is oriented NW – SE and thus along-strike of the decollement.

A Serravallian to early Tortonian strike-slip regime reactivated the set 1 joints and probably created NE
– SW trending normal faults and associated set 2 veins (section 6.4.5). Due to the larger height and
connectivity of these normal faults compared to the set 1 joints (section 5.8), this enhanced paleo-fluid
circulation and allowed migration of slightly warmer formational fluids from deeper stratigraphic levels
(section 6.4.5; Figure 6 - 18b).

Tortonian to Pliocene folding-and-thrusting in the Chaîne des Matheux resulted in progressive uplift
and subaerial erosion (section 5.8). Set 3 veins related to top-to-SW verging folding-and-thrusting are
characterized by either high temperature, relatively less depleted carbon isotope values, or by low
temperature, relatively more depleted carbon isotope values (section 6.4.5). The high temperature set
3 veins possibly precipitated from warm fluids derived from a deeper source, in a relatively open
system (Figure 6 - 18c). The geochemistry of the low temperature set 3 veins shows evidence for a
mixing with cold, meteoric surface waters carrying soil-derived CO2, which infiltrated the fault zone
and percolated downwards (Figure 6 - 18c).

227
Chapter 6: Paleo-fluid circulation

A younger phase of minor folding-and-thrusting with top-to-SE vergence is poorly constrained from
roughly the Messinian to the Pliocene (section 5.8). Set 4 calcite veins associated with this deformation
phase have intermediate homogenization temperatures compared to the set 3 veins (section 6.4.5).
The strongly depleted carbon isotope signal is interpreted to reflect the infiltration of cold, meteoric
surface waters carrying soil-derived CO2 along the faults (Figure 6 - 18d). The stable oxygen of the
parental fluid indicates a predominantly formational nature of the fluids, implying that the quantity of
surface waters was large enough to affect the stable carbon signal, but not significant enough to
change the stable oxygen signal.

During the Quaternary the southern front of the Chaîne des Matheux, proximal to the Cul-de-Sac plain,
became affected by strike-slip faults (section 5.8). Meteoric fluids percolated through these strike-slip
faults from which calcite precipitated along the fault planes (Figure 6 - 18e). The geochemical data
(section 6.4) is unfortunately insufficient to assess if an important component of deeper, warmer fluids
contaminated these meteoric waters and the calcite that mineralized from it.

228
Chapter 6: Paleo-fluid circulation

6.6.1 Tectonic setting, vein types, and microscopic observations


Calcite-silica veins

Samples containing veins of both calcite and silica are only found in Cretaceous limestones, either as
part of the Macaya Formation or intercalated between the CLIP basalts. There is no apparent
correlation between the deformation phase and the occurrence of silica within the veins, since they
are found in thrust-related and strike-slip settings, and along bedding-perpendicular joints. All sample
locations are however proximal to the CLIP basalts.

Three types of silica-calcite veins occur. The first one (Figure 6 - 21a) contains micritic quartz that is the
main vein-filling phase, which may be overprinted by younger calcite veins. The second type (Figure 6
- 21b) is an initial calcite vein that is being replaced by silica. This silica contains a very fine banding
that is parallel to the edge of the dissolved calcite crystals. This cryptocrystalline silica is probably a
chalcedony variety. The blocky extinction in the lower central part of the xpl image of Figure 6 - 21b
does not reflect the crystal structure of the silica, but results from a thin veneer of partly dissolved
calcite. The third type (Figure 6 - 21c) consists of a calcite vein that is was not completely cemented,
with subsequent precipitation of silica in the open voids in the center of the vein. This silica has a
flamboyant texture typical for chalcedony (Sander and Black, 1988; Dong et al., 1995). These veins do
not show a textural equilibrium between calcite and silica that would otherwise indicate coeval
precipitation of the two minerals. This precludes using stable oxygen isotopes from both calcite and
silica as an indicator of the temperature of precipitation, since this requires both phases to precipitate
simultaneously (Sharp and Kirschner, 1994).
PPL XPL CL XRD vein Notes
A cc
counts

CaCO3
1500 - silica is primary precipitation (oldest)
1000
SiO2 - silica is micritic (dominantly) or
HR chalcedony (minority)
si 500 - calcite is secondary (youngest)
0.5 mm 0
10 20 30 40 50 60 70
°2Θ CuK α

B counts
1500 CaCO3 - calcite is primary precipitation (oldest)
1000
- silica is secondary (youngest)
- silica is chalcedony variety (dominantly)
si HR SiO2 or micritic (minority)
500
cc - dissolution of calcite followed by
0.5 mm 0
precipitation of silica
10 20 30 40 50 60 70
°2Θ CuK α

C counts
CaCO3 - calcite is primary precipitation (oldest)
HR
1500

- silica is secondary (youngest)


cc si 1000
- silica is chalcedony variety, no micritic
500
SiO2 silica observed
HR - silica precipitated in voids within
0.5 mm 0
10 20 30 40 50 60 70 calcite vein
°2Θ CuK α

Figure 6 - 21: Examples of samples containing both silica and calcite veins. Host rocks are all Cretaceous slope
limestones. Abbreviations; HR = host rock, cc = calcite, si = silica.

231
Chapter 6: Paleo-fluid circulation

6.6.2 Stable oxygen and carbon isotopes


Host rock samples (n = 32) from the Southern Peninsula have a range of δ18O (VPDB) values from +0.26
to -5.95‰. The range of δ13C (VPDB) values is wider, from +3.01 to -6.33‰. The majority of the host
rock samples is depleted in both δ18O and δ13C with respect to typical Late Cretaceous and Cenozoic
marine carbonates (Figure 6 - 29a; Prokoph et al., 2008). This global range of δ18O values is similar to
the range of surface lake waters in Hispaniola over the past 10000 years (+2 to -2‰ δ18O; Hodell et al.,
1991; Lane et al., 2011).

There is no apparent correlation between the age of the host rock and the amount of depletion. Similar
to the Chaîne des Matheux, host rock samples from the Southern Peninsula often show strong
brecciation and re-cementation (i.e. Figure 6 - 25 and Figure 6 - 27), which may significantly reset the
original host rock isotope signal with that of the fluids responsible for the crystallization/re-
cementation. A stronger emphasis is thus placed on the isotope signal of the calcite veins rather than
the vein versus host rock signal when trying to understand the paleo-fluid circulation.

Vein calcite (n = 42) from the Southern Peninsula plots along two distinct trends on the δ18O vs δ13C
plot (Figure 6 - 29b). The first trend is along δ18O, with a range from -1.61 to -11.02‰, for veins with
positive δ13C values. This trend is mainly observed for the set 1, set 2 and set 3 veins (i.e. A, B, E and G
in Figure 6 - 29b). The second trend is towards more depleted δ13C, with a range from +3.15 to
-12.75‰, for veins with a limited range of δ18O from -0.5 to -5.5‰. This is mainly observed for set 4,
set 5 and set 6 veins (i.e. D, F, and H in Figure 6 - 29b). The samples from the core of the EPGFZ (set 4)
have an intermediate position along this δ13C mixing line. Both the set 5 thrust-related veins (H in
Figure 6 - 29b) and set 6a strike-slip and set 6b normal fault related veins from the west of the
Miragoane Basin (D and F in Figure 6 - 29b) plot on the lower end of this mixing line.

Set 1 calcite is in δ13C isotopic equilibrium with the host rock (Figure 6 - 29c), but strongly depleted in
oxygen at -6‰ Δ18O (Figure 6 - 29d).

Set 2 calcite from bedding perpendicular joints and set 3 from veins in a strike-slip setting are in δ13C
isotopic equilibrium with the host rock (Figure 6 - 29c) and are mildly depleted in oxygen between 0
and -4‰ Δ18O, which is the range for many samples (Figure 6 - 29d).

Set 4 calcite veins from the core of the EPGFZ diverge from this trend, being depleted in stable carbon
(-6 to -8‰ to Δ13C; Figure 6 - 29c) and enriched in stable oxygen (+3 to +4‰ to Δ18O; Figure 6 - 29d)
compared to the host rock.

238
Chapter 6: Paleo-fluid circulation

Set 5 thrust-related veins vary between -5 and -9‰ Δ13C compared to the host rock (Figure 6 - 29c).
Host rock of the sample depleted at -6‰ δ13C from set 5 is transected by small calcite veins that may
contaminate the host rock signal.

Set 6a and 6b calcite show a similar trend, with both depleted at -9‰ Δ13C compared to the host rock
(D and F in Figure 6 - 29c) and only mildly depleted in oxygen at -2‰ Δ18O (D and F in Figure 6 - 29d).
A Host rock ∂18O vs ∂13C B Vein sets
4
4 G
2
2
∂13C (‰VPDB) calcite

0
A
0 E B

∂13C (‰VPDB) calcite


-2 C
-2
-4
-4
-6
-6
-8 H
-8 F
-10
-12 -10 -8 -6 -4 -2 0
D
-12
∂18O (‰VPDB) calcite n = 32
-12 -10 -8 -6 -4 -2 0
C ∂13C vein vs host rock n = 42 ∂18O (‰VPDB) calcite
4
-10
‰ D ∂18O vein vs host rock
2 ck
= G
ro
∂13C Host rock (‰VPDB) calcite

ost B
-h C A
∂18O Hos rock (‰VPDB) calcite

0 ein 0
V E ‰ F D
Δ
13 C
H - 5‰
k=
-5 H E B
-2 D F roc
k= -2
t ro
c
A
st os
- ho G -h
-4 ein -4 ein
13 CV 18 OV C
Δ ‰ Δ
0 -6
-6 k= 5‰ 0‰
oc +
k=
st r oc
k=
roc +5

- ho r -8 st
-8 ein ost ho k=
-h n- oc
13 C
V
ein ei o st r
-10
Δ CV -10 18 OV -h
ein
13
Δ Δ
18 OV
-12 Δ
-12 -12 -10 -8 -6 -4 -2 0
-12 -10 -8 -6 -4 -2 0 2 4 n = 34 ∂18O Veins (‰VPDB) calcite
n = 34 ∂13C Veins (‰VPDB) calcite E Age of host rock vein
4
Age host rock; Fluid inclusion I.D. G
Miocene A = set 3 2
Paleocene B = set 3
Eoc. - Oligo. Cretaceous C = unknown
0
A
D = set 6b
Vein types; E B
Set 1 (thrust- Set 4 Set 6b -2 C
related) (EPGFZ core) (normal)
Set 2 Set 5 (thrust-
-4
(joints) related) Unknown
Set 3 -6
Set 6a
(strike-slip) (strike-slip) E to H see text
-8 H
F
∂18O and ∂13C of Late Cretaceous and
Cenozoic carbonates (Prokoph et al., 2008) -10
D
Bulk mantle (-5.0‰ ∂13C; Hoefs, 2009) -12
Mixing trend ∂13C -12 -10 -8 -6 -4 -2 0
Mixing trend ∂ O 18
n = 42 ∂18O (‰VPDB) calcite

Figure 6 - 29: Stable oxygen and carbon isotope data for veins and host rocks. A: Host rock δ 18O vs δ13C. B: Vein
sets δ18O vs δ13C. C: δ13C vein vs host rock. D: δ18O vein vs host rock. E: Age of the host rock vein δ 18O vs δ13C.
Vein stable isotope data labelled A, B, C and D correspond to the fluid inclusion examples A, B, C and D in Figure
6 - 32 and section 6.6.3. Samples labelled E, F, G and H; see text.

239
Chapter 6: Paleo-fluid circulation

The minimum age of the veins is constrained by the age of the host rock, which is shown in Figure 6 -
29e. There is no apparent correlation between minimum age of the vein and the δ18O or δ13C values.

Vein deformation and cathodoluminescence

Observations on samples with cross-cutting calcite veins shows that; 1) the oldest veins are more
deformed than the youngest ones (Figure 6 - 30), 2) the youngest veins have lower
cathodoluminescence intensity compared to the older veins, irrespective of the luminescence of the
host rock (Figure 6 - 30), and 3) the least deformed veins have lower δ13C values compared to the most
deformed veins (Figure 6 - 31a).

A Vein deformation and cathodoluminescence


PPL XPL CL Notes
I I = micritic silica
2 mm I
II = calcite 1 (oldest)
II II III = deformation zone
III III
IV IV IV = calcite 2 (youngest)
III III Vein; ∂13C ∂18O Defo CL
Calcite 1 (II) -0.10 -6.46 5 3
II II
Calcite 2 (IV) -6.01 -4.30 2 1
B
PPL CL

II I = brecciated limestone
II = calcite 1 (oldest)
I III III = calcite 2 (intermediate)
2 mm
IV = calcite 3 (youngest)
Vein; ∂13C ∂18O Defo CL
Calcite 1 (II) -4.67 -2.97 4 2
IV Calcite 3 (IV) -7.37 -1.98 3 0

III

Figure 6 - 30: Vein deformation and cathodoluminescence intensity. A; Chert layer from upper Eocene limestones
west of Miragoane Basin. B; Core of EPGFZ, same location as Figure 6 - 25. Relative deformation (Defo) and
cathodoluminescence (CL) intensities; see text for explanation.

To understand the mechanism behind the observations above, all veins from the Southern Peninsula
are assigned a relative value of deformation intensity (0 being undeformed, 5 being strongly deformed)
and a relative value for cathodoluminescence intensity (0 being non-luminescent, 5 being brightly
luminescent). On a plot of δ18O versus δ13C, overlain with the relative deformation intensity (Figure 6
- 31a), veins with the lowest δ13C values are generally the least deformed. Note that few samples show
an inverse relation to this trend. There is no apparent correlation between δ 18O and deformation
intensity (Figure 6 - 31a).

On a plot of deformation versus cathodoluminescence intensity, overlain with the δ 13C value of the
vein calcite, a similar trend is observed (Figure 6 - 31b). The least deformed veins have the lowest

240
Chapter 6: Paleo-fluid circulation

that Mn2+ levels are not too high (Boggs and Krinsley, 2006). To test if the non- to dull luminescent
veins result from precipitation of calcite from oxygenated waters, or if this is due to insufficient
concentrations of divalent Mn, the ppm count of Fe and Mn is plotted in Figure 6 - 31c. Note that from
the whole-rock data there is no control on the valence of Mn and Fe, and that only the total
concentrations are plotted. The samples that display medium to bright luminescence contain between
80 and 1250 ppm of Mn (Figure 6 - 31c). The fact that these samples are luminescent indicates that
enough Mn is in a divalent state within these veins. The veins that are non- to dull luminescent
unfortunately all have Mn concentrations below the detection limit of 77 ppm, which is higher than
the ppm limit for dull luminescence activation (Figure 6 - 31c; Machel (2000). It can therefore not be
concluded if the lack of luminescence is due to the presence of Mn4+ and thus related to oxygenated
conditions, or due to a general absence of Mn in the samples, irrespective of the valence.

The deformation intensity of the veins is overlain on the Mn vs Fe plot in Figure 6 - 31d. This shows
that the less-deformed, non- to dull luminescent veins generally have low concentrations of Mn,
whereas strongly deformed veins with dull to bright luminescence have higher Mn concentrations.

6.6.3 Fluid inclusion microthermometry


A total of 41 chips from 7 samples were analyzed, of which the vast majority could not be used for fluid
inclusion microthermometry studies. The limitations are similar to those encountered with the Chaîne
des Matheux samples (section 6.4.3), which is mainly related to twinning and deformation of the
calcite on the one hand, and the irreversible loss of the vapor bubble upon analysis, for instance due
to inclusion deformation and/or leaking, on the other. Based on all these limitations and the low
number of measured samples, these results should be viewed and interpreted with caution. Chips from
3 quartz veins were also analyzed for fluid inclusions. The silica in these samples was either micritic or
a chalcedony variety. No inclusions could be identified, probably due to the small crystal size of the
silica.

Only 11 reliable homogenization temperatures were measured from 5 chips, taken from 4 samples.
Because these inclusions appear isolated within the calcite crystal, with no clear relation to the crystal
structure or growth zones, and not on inclusion trails nor on twin planes, these inclusions are
interpreted as primary fluid inclusions.

All the fluid inclusions presented in Figure 6 - 32a are filled with clear liquids, without any signs of
hydrocarbon within the inclusions. The inclusion in Figure 6 - 32a were tested using Raman
spectroscopy to try and identify the composition of the vapor phase. Unfortunately, no signal other
than calcite (CaCO3) and water (H2O) could be identified.

242
Chapter 6: Paleo-fluid circulation

Fluid inclusions from set 3 veins, samples A and B (Figure 6 - 32c), have homogenization temperatures
of 140 to 170 °C. Calcites from set 6b that precipitated along normal faults to the west of the Miragoane
Basins display lower temperatures ranging from 65 to 105 °C (Figure 6 - 32c).

A plot of δ18O (VPDB) calcite versus fluid inclusion homogenization temperature, using oxygen isotope
fractionation curves calculated after Kim and O’Neil (1997) (Figure 6 - 32d), shows that the δ18O
(SMOW) of the parent fluids from which the calcite precipitated ranges from 6 to 20‰. Samples A and
B from set 3 veins are located within the domain of formational waters (Figure 6 - 32d; Prokoph et al.,
2008). Set 6b calcite veins precipitated from fluids with 6 and 14‰ δ18O (SMOW) (Figure 6 - 32d),
indicative for formation waters mixed with meteoric fluids (Prokoph et al., 2008).

6.6.4 Whole-rock geochemistry


A total of 10 veins and associated host rock samples from the Southern Peninsula were analyzed for
their whole-rock composition, of which the REE and Yttrium (REY) data is plotted in Figure 6 - 33a and
Figure 6 - 34. This small number is due to the large quantity of material (>5 grams) of material needed
for analysis, which limited the size of the veins that could be targeted. An additional 8 Cretaceous and
3 Paleocene basalt samples were analyzed for their whole-rock composition, of which the REE and
Yttrium (REY) data is plotted in Figure 6 - 33b.

Host rock limestones (Figure 6 - 34a) have negative Ce, mildly positive Y anomalies, and a mild
enrichment in heavy REE compared to light REE, which is typical for carbonates that formed in a marine
environment (Bau et al., 1997; Bolhar et al., 2004; Bau et al., 2010; Piper and Bau, 2013). The
Cretaceous and Paleocene basalts have similar REY patterns (Figure 6 - 33b), which are markedly
different from the limestones. These basalts do not have a Ce or Y anomaly, but show a strong positive
Eu anomaly (Figure 6 - 33b). The Paleocene volcaniclastics have REY patterns intermediate between
the basalts and limestones (Figure 6 - 33a). They are characterized by a positive Eu anomaly, lack a Y
anomaly, and may or may not show a negative Ce anomaly.

Large positive Eu anomalies are reported for fluids expelled from, and limestones that formed in the
vicinity of, hydrothermal vent systems at mid-oceanic ridges (Michard et al., 1983; Bau and Dulski,
1999). This positive Eu anomaly results from fluid-rock interactions with basalts and requires fluid
temperatures in excess of 200 °C in order to liberate the Europium, while upon cooling the fluids retain
their positive Eu signal, which is subsequently incorporated into calcite and limestones that form
proximal to these systems (Bau et al., 2010).

244
Chapter 6: Paleo-fluid circulation

A Host rock types Southern Peninsula: REY (PAAS normalized)


1
Sediments

0,1
REY/PAAS

0,01

0,001
La Ce Pr Nd Sm Eu Gd Tb Dy Y Ho Er Tm Yb Lu
LREE MREE HREE
B 10
Basalts

1
REY/PAAS

0,1

0,01
La Ce Pr Nd Sm Eu Gd Tb Dy Y Ho Er Tm Yb Lu
LREE MREE HREE
Sediments; Basalts;
Eocene Cretaceous Paleocene Paleocene Paleocene Cretaceous
limestones limestones limestones (volcani)clastics basalts basalts

Figure 6 - 33: Rare earth elements + Yttrium (REY) data for samples from the Southern Peninsula, PAAS
normalized after Taylor and McLennan (1985). A: REY for host rock sediments. B: REY for basalts.

A positive Eu anomaly therefore signals high-temperature fluid-rock interactions with basalts, although
it has no implications for the temperature of the fluids at the time of calcite precipitation.

REY values for set 2 calcite veins are slightly depleted compared to the limestone host rock and
generally show a similar pattern, with negative Ce and positive Y anomalies similar to the host rock
(Figure 6 - 34a), indicating significant fluid-host rock interaction in a relatively closed system (Nuriel et
al., 2012). Two of these veins however also display a mild to pronounced Eu anomaly. Both of these
samples are calcite veins from upper Paleocene limestones that directly overly lower Paleocene
volcaniclastics and basalts. This positive Eu anomaly possibly results from a mixture of intra-
formational fluids with fluids from the underlying lower Paleocene basalts and volcaniclastics, which
subsequently migrated into the overlying upper Paleocene formations along the set 2 joints.

Set 3 calcite veins have very similar REY concentrations as the limestone host rock, with pronounced
negative Ce and positive Y anomalies (Figure 6 - 34b), indicating strong fluid-host rock interaction in a
relatively closed system (Nuriel et al., 2012). One vein sample has a very pronounced positive Eu
anomaly. This vein is located in upper Paleocene limestones directly overlying lower Paleocene basalts.
The positive anomaly could again be the result of a mixture of intra-formational fluids with fluids from
the underlying basalts that migrated along the associated strike-slip faults.

245
Chapter 6: Paleo-fluid circulation

A Veins Southern Peninsula: REY (PAAS normalized)


1
Set 2 veins

0,1
REY/PAAS

0,01

0,001
La Ce Pr Nd Sm Eu Gd Tb Dy Y Ho Er Tm Yb Lu
LREE MREE HREE
B 1
Set 3 veins

0,1
REY/PAAS

0,01

0,001
La Ce Pr Nd Sm Eu Gd Tb Dy Y Ho Er Tm Yb Lu
LREE MREE HREE
C 1
Set 6b veins

0,1
REY/PAAS

0,01

0,001
La Ce Pr Nd Sm Eu Gd Tb Dy Y Ho Er Tm Yb Lu
LREE MREE HREE
D 1
Silica veins & chert layers

0,1
REY/PAAS

0,01

0,001
La Ce Pr Nd Sm Eu Gd Tb Dy Y Ho Er Tm Yb Lu
LREE MREE HREE
Set 2 Set 3 Set 6b Chert Silica
(joint) (strike-slip) (normal)
Vein
Host Rock
Figure 6 - 34: Rare earth elements + Yttrium (REY) data for samples from the Southern Peninsula, PAAS
normalized after Taylor and McLennan (1985). Host rock and veins with the same line pattern are from the same
sample. A: REY for set 2 calcite veins. B: REY for set 3 calcite veins. C; REY for set 6b calcite veins. D; REY for silica
veins and chert layers.

246
Chapter 6: Paleo-fluid circulation

Set 6b veins from normal faults located at the western end of the Miragoane Basin are strongly
depleted in REY concentrations and have a different REY pattern compared to the limestone host rock
(Figure 6 - 34c). This probably indicates limited fluid-host rock interaction or the infiltration of a fluid
that is not in geochemical equilibrium with the host rock (Nuriel et al., 2012).

Chert layers within Eocene limestones have a similar REY pattern as the host rock (Figure 6 - 34d).
These are depleted in Ce but do not show a Y anomaly. The similar REY concentration of chert and host
rock suggests a close genetic relation. The silica veins are strongly depleted in REY compared to their
host rock, with several REE’s below the detection limit (Figure 6 - 34d). The host rock for one of these
veins (solid line) is a lower Paleocene volcaniclastic, which has a similar REY pattern as the Paleocene
basalts (Figure 6 - 33b).

247
Chapter 6: Paleo-fluid circulation

Set 1: Thrust-related calcite-silica veins

Set 1 veins are tentatively related to the latest Cretaceous to early Paleocene deformation event
(section 6.6.1). Calcite from these thrust-related veins have the lowest δ18O values of all the samples,
and high δ13C values (Figure 6 - 35a). The calcite is in δ13C equilibrium with the host rock, but strongly
depleted in oxygen at -6‰ Δ18O (section 6.6.2). Although not constrained by fluid inclusion
microthermometry, the low δ18O could suggest higher temperatures for the fluids compared to the
host rock (Ünal-Ïmer et al., 2016). The very bright luminescence of the calcite (section 6.6.1) indicates
that it precipitated from reducing fluids (Boggs and Krinsley, 2006). The similarity in luminesce intensity
with the host rock (section 6.6.1) suggests a relatively closed system with significant fluid-host rock
interaction on the one hand, while the very negative Δ18O on the other hand suggests a disequilibrium
between the fluids and the host rock.

It is thus unclear if the system was relatively open, with infiltration of deeper, warmer fluids, or
relatively closed and characterized by intra-formational paleo-fluid circulation.

Set 2: Orthogonal, bedding perpendicular veins

The orthogonal bedding perpendicular veins of set 2 resulted from layer-parallel shortening during the
late Miocene (section 4.4). The deformed vein calcite indicates post-precipitation deformation, which
is locally reflected by strike-slip reactivation of the joints. Set 2 veins have high δ13C and mildly negative
δ18O values (Figure 6 - 35b). These veins are in δ13C equilibrium with the host rock, but mildly depleted
in δ18O (section 6.6.2). The veins are mildly depleted in REY compared to the host rock and have a
similar pattern (section 6.6.4).

The above is indicative for intra-formational paleo-fluid circulation in a relatively closed system. Some
veins show a minor positive Eu anomaly that may reflect a contamination with fluids from the basalts
that are directly underlying the limestone host rock at that location. The fluids probably migrated into
the overlying formations by pathways created by the set 2 joints.

Set 3: Strike-slip shear veins at distance from the EPGFZ

Set 3 strike-slip related calcite veins collected at distance from the active trace of the EPGFZ are related
to the initial stage of strike-slip activity on the Southern Peninsula, during which deformation was more
distributed (section 4.2). These veins have high δ13C and a wide spread in δ18O values (Figure 6 - 35c).
The veins are in carbon isotope equilibrium with the host rock and mildly depleted in oxygen (section

249
Chapter 6: Paleo-fluid circulation

6.6.2). Veins and host rock generally have similarly bright luminescence (section 6.6.1). The veins
precipitated from formational fluids at temperatures between 140 – 170 °C (section 6.6.3). A
pronounced positive Eu anomaly in one of the veins is probably related to contamination of fluids
derived from the underlying Paleocene basalts at that location. Based on the stratigraphy as
constrained in section 4.2, around 3 – 4 km of overburden has eroded since the precipitation of these
veins.

The homogenization temperatures either indicate anomalously high geothermal gradients in the area,
or that the fluids originated at a deeper, hotter source. Due to the geochemical and
cathodoluminescence equilibrium between the veins and the host rock, an elevated geothermal
gradients appears most plausible, although a deep fluid source cannot be ruled out.

Set 4: Veins from the core of the EPGFZ

The set 4 calcite veins within the core of the EPGFZ are related to activity on this fault, which is active
from the late Miocene or Pliocene onwards (section 4.2). The brecciated core of this fault acts as a
fractured medium, whose deformation is probably not controlled by the Coulomb failure envelope but
by the frictional sliding criterion (Figure 6 - 35d; section 6.2.1). Set 4 calcite veins are characterized by
negative δ13C values (Figure 6 - 35d). Compared to the host rock these veins are enriched in δ18O but
depleted in δ13C. The luminescence is also in disequilibrium with the host rock. The above suggests
that the fluids from which this calcite precipitated were not in equilibrium with the host rock. A
possible explanation for the low δ13C values (-5 to -7.5‰) could be a mixing with meteoric waters
carrying soil derived CO2 that percolated downwards along the EPGFZ. Such oxidizing fluids could
explain the observed non-luminescent calcite, assuming that enough Mn is present in the veins. Some
of the calcite within these veins is however luminescent, which possibly precipitated from reducing
fluids. Because the EPGFZ is expected to join the Moho discontinuity and could have a connection with
the mantle (Benford et al., 2012c), a lower crustal or upper mantle derived component of the fluids
could also be responsible for the low δ13C values. Because the mantle has a δ13C of -5 ‰ (Hoefs, 2009),
a mantle component alone is not enough to explain the observed δ13C values of these set 4 veins.

It is likely that the calcite precipitated from fluids that are a mixture of deeper lower crustal or upper
mantle derived fluids on the one hand, and meteoric surface waters on the other. Unfortunately, due
to the small sizes of the calcite veins, no whole-rock samples could be obtained and thus no REY data
is available to test this hypothesis.

250
Chapter 6: Paleo-fluid circulation

Set 5: Thrust-related veins

South-verging thrusts locally overprint strike-slip activity in the Southern Peninsula, whose timing is
constrained from Pliocene to present-day (section 4.2). The δ13C values for set 5 calcite veins related
to these thrusts ranges from -9 to -12 ‰ (Figure 6 - 35e), which is similar to the thrust-related calcite
veins in the Chaîne des Matheux. The veins are strongly depleted in stable carbon compared to the
host rock (section 6.6.2). Set 5 calcite is non-luminescent, which differs from the generally luminescent
host rock (section 6.6.1). The above suggests that these veins precipitated from fluids that were not in
equilibrium with the host rock.

The low δ13C values possibly indicate that the veins precipitated from meteoric surface waters carrying
soil-derived CO2 that percolated down along the thrust faults, similar to what is proposed in the Chaîne
des Matheux for low δ13C thrust-related calcite veins (section 6.4.5).

Set 6: Strike-slip and normal faults west of Miragoane Basin

Set 6a and set 6b calcite veins along strike-slip and normal faults on the western end of the Miragoane
pull-apart basin share the same mineralogical, isotopic, and cathodoluminescence properties (section
6.6.2 and 6.6.3), and are likely related to opening of this basin during the Quaternary or possibly
Pliocene (section 4.2). Set 6 vein calcite has very negative δ13C values (Figure 6 - 35f) and is in carbon
isotopic disequilibrium with the host rock (section 6.6.2). The calcite is non-luminescent and thereby
differs from the brightly luminescent host rock (section 6.6.1). The temperature of the fluids ranges
from 65 – 105 °C, with δ18O of the parental fluid indicating a mixture of formational and meteoric
waters (section 6.6.3). Set 6b veins are strongly depleted in REY concentrations and have a different
REY pattern compared to the limestone host rock (section 6.6.4). Based on the difference in
luminescence and the disequilibrium in stable carbon isotope with the host rock, the REY data is
interpreted to reflect that the calcite precipitated from external fluids not in equilibrium with the host
rock.

The source for these low δ13C vein calcites could be meteoric waters carrying soil-derived CO2, which
infiltrated the subsurface by percolating along the normal and strike-slip faults.

251
Chapter 6: Paleo-fluid circulation

6.7 Paleo-fluid circulation and evolution of the Southern Peninsula


The latest Cretaceous to early Paleocene phase of deformation is associated with reactivation of the
bedding planes and thrusts with a predominantly SW-vergence (section 4.2). The geochemistry of
associated set 1 vein calcite mainly indicates paleo-fluid circulation in a relatively closed system (Figure
6 - 36a), although a minor contamination with relatively warm fluids cannot be excluded.

A S N B S N
latest Cretaceous - early Paleocene (set 1) early Miocene

No samples

C S N DS N
late Miocene (set 2) late Miocene (set 3)

?
?
E S N F S N
≤ late Miocene (set 4) Pliocene (set 5)
EPGFZ EPGFZ

? ?
G S N
(Plio?) - Quaternary (set 6)
Miragoane
EPGFZ basin

low-T; low δ13C)


(medium-T; high δ13C)
(high-T; medium δ13C)
?

Figure 6 - 36: Schematic evolution of paleo-fluid circulation in the Southern Peninsula.

No veins related to the early Miocene deformation phase have been found (Figure 6 - 36b)

The orthogonal bedding perpendicular veins of set 2 resulted from layer-parallel shortening during the
late Miocene (section 4.4). The geochemistry of these veins is mainly indicative for intra-formational
paleo-fluid circulation in a relatively closed system, with locally some contamination of fluids derived
from underlying basalts (section 6.6.5). These fluids probably migrated into the overlying formations
by pathways created by the set 2 joints (Figure 6 - 36c).

Set 3 strike-slip related calcite veins are related to the initial stage of strike-slip activity on the Southern
Peninsula, during which deformation was more distributed (section 4.2). The geochemistry of these

252
Chapter 6: Paleo-fluid circulation

veins is again mainly indicative for intra-formational paleo-fluid circulation in a relatively closed
system, with locally some contamination of fluids derived from underlying basalts (section 6.6.5; Figure
6 - 36d). The 140 – 170 °C homogenization temperatures of the fluid inclusions (section 6.6.3) coupled
with only 3 – 4 km of overburden probably indicates that the geothermal gradients was anomalously
elevated, although a deep fluid source cannot be ruled out.

The set 4 calcite veins within the core of the EPGFZ are related to activity on this fault, which is active
from the late Miocene or Pliocene onwards (section 4.2). The geochemistry of these calcites indicates
that they precipitated from fluids in disequilibrium with the host rock, probably a mixture of meteoric
waters and a component of fluids with a deeper, lower crustal or upper mantle origin (section 6.6.5;
Figure 6 - 36e). It is likely that the increase of deformation along the EPGFZ resulted in increased
pathways for deeper fluids, but the hard evidence for such a paleo-fluid circulation is presently lacking.

South-verging thrusts locally overprint strike-slip activity in the Southern Peninsula, whose timing is
constrained from Pliocene to present-day (section 4.2). The geochemistry and microscopic
observations on associated set 5 calcite veins suggests that they precipitated from fluids that were not
in equilibrium with the host rock. It is most likely that the veins, characterized by low δ13C values
(section 6.6.2), precipitated from meteoric surface waters carrying soil-derived CO2 that percolated
down along the thrust faults (Figure 6 - 36f). No component of deeper crustal fluids is observed.

Following the focus of strike-slip deformation along the EPGFZ a number of pull-apart basins, such as
the Miragoane basin, were created during the Quaternary or Pliocene (section 4.2). Faults on the
western termination of this basin acted as pathways for the downwards percolation of relatively cold,
meteoric fluids with soil-derived CO2 (Figure 6 - 36g), from which set 6 calcite veins with low δ13C values
precipitated (section 6.6.2). There is no evidence for mixing with deeper crustal or upper mantle fluids.

253
Chapter 6: Paleo-fluid circulation

254
Chapter 7: Discussion and regional implications

7 Discussion and regional implications

This last chapter is aimed towards three goals. The first is to better understand how deformation is
distributed over southwestern Hispaniola, how it changed over geological time scales, what the drivers
of deformation are, assess the impact of the EPGFZ on deformation on the Southern Peninsula, and to
compare the above with present-day strain rates. The second aim is to use the deformation history of
the Southern Peninsula as constrained in this study, combined with the increased knowledge on the
stratigraphy, to build paleogeographic reconstruction maps of the Southern Peninsula in Haiti. The
third direction is to construct a crustal-scale cross section linking the Southern Peninsula and Chaîne
des Matheux, use this to identify possible fluid pathways and sources, and assess the implications for
the regional geodynamic framework.

7.1 Shortening and shortening rates in southwestern Hispaniola


The purpose of this section is to quantify the amount of shortening to the southwest of the Massif de
Nord – Cordillera Central in Haiti and the Dominican Republic, which helps understand the geodynamic
evolution of this region. Based on the structural and morphological map (section 3.1, Figure 3 - 1) it
appears that NE – SW directed shortening is increasing from the northwest to the southeast, because;
1) the northwest is at lower elevations (mainly offshore) compared to the southeast (entirely onshore),
2) the wavelength and amplitude of the structures are smaller in the northwest compared to the
southeast (based on bathymetry and DEM), and 3) faults are more closely spaced in the southeast
compared to the northwest. To check if shortening really changes with location and if so, by how much,
shortening is constrained using cross sections from this study (chapter 4 and 5) and from the literature
(Appendix 1). The implications that the amount and timing of shortening has on the Southern
Peninsula, Haitian Fold-and-Thrust Belt, and the northern Caribbean region, is subsequently discussed.

The data are combined with the structural framework of the morphotectonic zones in southwestern
Hispaniola to create strain distribution maps. Using the timing and duration of deformation within
these morphotectonic zones, as constrained in the previous chapters, these strain maps are converted
into average strain rate maps, which are compared with present-day strain rate maps derived from
GPS velocity modeling studies.

255
Chapter 7: Discussion and regional implications

7.1.1 Comparison of shortening with published cross sections


Cross sections from this study used to calculate shortening are found in section 4.6 and 5.5. The
published cross sections are found in Appendix 1, while the resulting shortening is summarized in Table
7 - 1 and Figure 7 - 1.

Cross sections from this study are between 15 and 35 km in length and constructed perpendicular to
the main structural trends. Selected cross sections from literature are chosen with lengths between 10
and 50 km, and with strikes that are perpendicular to the local structural trend, or within a ±20° error
margin. Note that the large onshore sections F, G, H and I (Figure 7 - 1) are constructed from multiple
small, 10-km length-scale cross sections (Appendix 1). The selected cross sections are broken down to
constrain the amount of shortening over discrete morphotectonic zones. These are, from north to
south; 1) the Peralta Belt, 2) the Montagnes Noires – Northern Sierra de Neiba (including the Plateau
Central and San Juan Basin), 3) Chaîne des Matheux – Southern Sierra de Neiba, 4) Sierra de Martín
García, 4) Cul-de-Sac – Enriquillo Basin, 5) Northern Massif de la Selle – Sierra de Bahoruco, 6) Southern
Massif de la Selle – Sierra de Bahoruco, and 7) Massif de la Hotte.

To constrain the total amount of shortening, a horizon is selected that was deposited prior to the onset
of deformation. In the Peralta Belt, folding and thrusting started during the late Eocene (Witschard
and Dolan, 1990; Hernaiz Huerta and Pérez-Estaún, 2002), and the base middle Eocene is used (Table
7 - 1). In the regions Plateau Central – San Juan Basin – Montagnes Noires – Sierra de Neiba – Chaîne
des Matheux – Cul-de-Sac – Enriquillo Basin, diachronous folding-and-thrusting started during the
Aquitanian or more recent (section 5.2; Pubellier et al., 2000), and either the base of the Miocene or
the base of the middle Eocene is selected (Table 7 - 1). Three phases of deformation affected the
Southern Peninsula during the Cenozoic (section 4.2). To allow a comparison with Miocene to recent
deformation in the regions north of the Southern Peninsula, and without incorporating the latest
Cretaceous to early Paleocene deformation phase (section 4.2), the base of the upper Paleocene is
selected in the Massif de la Selle and l’Asile basin areas (Table 7 - 1). This incorporates shortening
resulting from two deformation phases, which is discussed accordingly. The Massif de la Selle region is
unaffected by the early Miocene phase of deformation and thus only enjoyed post-early Paleocene
shortening related to the late Miocene to recent phase (section 4.2). When possible the base of the
Miocene is selected, and otherwise the base of the middle Eocene is chosen, which both record the
same deformation phase (Table 7 - 1).

256
Chapter 7: Discussion and regional implications

Length Line-length Shortening Heading Heading QC heading Reference Shortening


Label Domain horizon (base)
[Km] [Km] km % section [°] structures [°] [Δ°] cross section
A Western la Hotte 26.30 27.20 0.9 3.3 0 90 0.0 This study U. Paleocene
B L'Asile Basin 17.20 18.10 0.9 5.0 15 90 15.0 This study U. Paleocene
C Selle South Flank 18.30 18.50 0.2 1.1 0 100 10.0 This study Miocene
C Selle North Flank 14.83 16.51 1.7 10.2 0 95 5.0 This study Miocene
C Selle total 33.13 35.01 1.9 5.4 0 95 5.0 This study Miocene
D&E Selle South Flank 12.98 13.33 0.4 2.6 30 100 20.0 VdB-83 M. Eocene
F Selle North Flank 6.68 7.57 0.9 11.8 15 110 5.0 HH-07 Miocene
F Enriquillo 11.58 11.99 0.4 3.4 15 105 0.0 HH-07 Miocene
F Neiba - Matheux eq. 12.90 15.41 2.5 16.3 15 105 0.0 HH-07 M. Eocene
G Bahoruco North Flank 12.87 16.15 3.3 20.3 30 110 10.0 HH-07 Miocene
G Enriquillo 15.64 16.39 0.8 4.6 30 100 20.0 HH-07 Miocene
G Neiba - Matheux eq. 10.81 14.03 3.2 23.0 30 115 5.0 HH-07 Miocene
G Neiba - Noires eq. 33.91 43.65 9.7 22.3 30 115 5.0 HH-07 Miocene
H Bahoruco North Flank 4.67 5.57 0.9 16.2 30 110 10.0 HH-07 Miocene
H Azua 17.08 17.51 0.4 2.5 30 110 10.0 HH-07 Miocene
I Peralta Belt 32.87 71.13 38.3 53.8 50 135 5.0 HH-02 M. Eocene
H&I Martin Garcia 61.34 69.95 8.6 12.3 50 120 20.0 HH-02&07 Miocene
J Matheux 26.96 38.92 12.0 30.7 45 130 5.0 This study Miocene
K Cul-de-Sac 9.12 9.69 0.6 5.9 45 115 20.0 This study Miocene
K Matheux 17.90 23.44 5.5 23.6 45 130 5.0 This study Miocene
L Montagnes Noires 36.32 43.17 6.9 15.9 30 125 5.0 Pub-00 M. Eocene
M Montagnes Noires 45.06 53.18 8.1 15.3 30 125 5.0 Pub-00 M. Eocene
N Cul-de-Sac eq. 44.19 44.29 0.1 0.2 30 110 10.0 Corb-16 Miocene
N Matheux eq. 37.17 37.50 0.3 0.9 30 120 0.0 Corb-16 Miocene
N Montagnes Noires eq. 8.64 8.71 0.1 0.7 30 110 10.0 Corb-16 Miocene
Table 7 - 1: Shortening in southwestern Hispaniola. Abbreviations; VdB-83 = Van den Berghe (1983a), HH-07 = Hernaiz Huerta et al. (2007), HH-02 = Hernaiz Huerta and Pérez-
Estaún (2002), Pub-00 = Pubellier et al. (2000), Corb-16 = Corbeau et al. (2016).

258
Chapter 7: Discussion and regional implications

All the cross sections in Appendix 1 are 1:1 but their respective scales differ. Cross sections D and E
from Van den Berghe (1983a) are less than 10 km in length, and combined to provide a reasonable
estimate of shortening of the southern flank of the eastern Massif de la Selle (Figure 7 - 1). These cross
sections are rescaled to 1:1 to remove vertical exaggeration (Appendix 1). Shortening is calculated
using the base of the middle Eocene. Cross section F (Figure 7 - 1) is a compilation of two cross sections
constructed by Hernaiz Huerta et al. (2007). Cross sections G and H (Figure 7 - 1) are from the same
authors, which are compilations of three and two smaller cross sections, respectively. Shortening on
all three is calculated using the base of the Miocene (Appendix 1). In cross section F the line of
calculation jumps from the base Miocene to the base middle Eocene along a sub-vertical fault. Because
there is no shortening over a vertical fault (Woodward et al., 1989), this will have a minimal impact on
the calculated shortening. Cross section G only covers the southern part of the Sierra de Neiba – San
Juan Basin region. To provide an estimate of total shortening of this thrust sheet, the northern limb is
extended until the LPSJFZ (Figure 7 - 1), and the average percentage of shortening for this part of the
structure is projected from cross section L.

To determine shortening in the Sierra de Martín García, the northern part of cross section H is
combined and rescaled to fit with the southern part of cross section I (Figure 7 - 1). Cross section I is
from Hernaiz Huerta and Pérez-Estaún (2002), which is compiled from 4 smaller cross sections.
Shortening along this cross section is calculated using the base of the Miocene (Appendix 1). Over the
LPSJFZ towards the Peralta Belt the line of calculation jumps to the base middle Eocene, since neither
the thickness of the Eocene under the Azua Basin nor the Miocene in the Peralta Belt are constrained.
This results in an underestimation of the total shortening, which is thus only a minimum. Cross sections
L and M are from Pubellier et al. (2000). Both cross sections are rescaled to 1:1 to remove vertical
exaggeration. Shortening is calculated using the base of the middle Eocene (Appendix 1). Along cross
section M the line of calculation changes to base Miocene on the back-limb of the structure along a
steeply dipping normal fault. This results in an underestimation of the total shortening, which is thus
only a minimum. Cross section N is an offshore seismic line (Figure 7 - 1) from Corbeau et al. (2016).
The top image in Appendix 1 is at two times vertical exaggeration for the sea floor using an average
1.5 km/s seismic velocity in sea water (Bourbié et al., 1987). The bottom image shows the base of the
Miocene unit, as interpreted by Corbeau et al. (2016), rescaled to be roughly 1:1. This is achieved using
an averaged seismic velocity of 3.0 km/s for mildly consolidated marls/limestones (Bourbié et al.,
1987), which is the likely lithology for this Miocene package based on the onshore Miocene
stratigraphy (section 5.1.2). The lack of seismic velocities or well constraints for this seismic dataset
makes any attempted depth conversion very speculative. The resultant geometries and shortening
should therefore be treated with caution.

259
Chapter 7: Discussion and regional implications

7.1.2 Distribution of shortening


The averaged percentage of shortening per morphotectonic zone, as constrained using published cross
sections and the cross section from this study, is shown in Figure 7 - 1.

Shortening in the Massif de la Hotte varies between 3 and 5% (Figure 7 - 1). On the southern flank of
the Massif de la Selle and Sierra de Bahoruco shortening is only between 1 and 3% (Figure 7 - 1). There
is a sharp contrast between the amount of shortening on the southern flank and along the northern
flank in this region (Figure 7 - 1). In the Massif de la Selle shortening on the northern flank varies
between 10 and 12% (Table 7 - 1), while further east on the northern flank of the Sierra de Bahoruco
shortening varies between 16 and 20% (Table 7 - 1). Shortening in the Cul-de-Sac – Enriquillo – Azua
Basins varies from 3 to 6% (Table 7 - 1), with average values of 4% (Figure 7 - 1). This is significantly
less than for the regions north and south of these basins. Shortening in the Chaîne des Matheux –
southern Sierra de Neiba varies between 16 and 30% (Table 7 - 1). There is no clear trend in the
geographical distribution, and the average value of 23% is taken for the entire chain (Figure 7 - 1).
Shortening in the Montagnes Noires is around 15.5%, while the northern Sierra de Neiba is shortened
by around 22% (Table 7 - 1). Although this variation could be due to differences in construction of the
cross sections of Pubellier et al. (2000) versus Hernaiz Huerta et al. (2007), a trend of increased
shortening from northwest to southeast is tentatively attributed to the Montagnes Noires – northern
Sierra de Neiba. An oblique strike-slip fault possibly separates the Sierra de Neiba from the Sierra de
Martín García region (Hernaiz Huerta et al., 2004, 2007). Shortening in the Martín García region is
averaged at 12.3% (Figure 7 - 1). The eastern Peralta Belt is shortened by 53.8%, which is the highest
value of all regions (Table 7 - 1). The offshore domain of the Golf of Gonâve on the other hand has the
lowest shortening values of less than 1% (Figure 7 - 1). Shortening in this region is probably
underestimated by the resolution of the seismic data and the scale of interpretation.

7.1.3 Implications for displacement along the EPGFZ


The amount of Miocene to present-day shortening in the Massif de la Selle and Sierra de Bahoruco as
calculated in section 7.1.1 can serve as a constraint for quantifying the minimum amount of strike-slip
displacement along the EPGFZ during this time.

Figure 7 - 1 shows a spatial difference in the amounts of shortening accommodated in the Massif de la
Selle and Sierra de Bahoruco. This difference correlates with a change in the structural trend of faults
bounding the Massif de la Selle – Sierra de Bahoruco (Figure 7 - 1). In the northern Massif de la Selle,
the north-verging thrusts strike N095°E, while in the northern Sierra de Bahoruco these faults strike
N110°E. In the west these thrust faults originate in the area where the EPGFZ intersects the Cul-de-Sac

260
Chapter 7: Discussion and regional implications

The amount of shortening in the three different areas (northern Massif de la Selle, northern Sierra de
Bahoruco, and eastern Sierra de Bahoruco) is projected onto the GPS velocity vector for the
corresponding region (green arrow, Figure 7 - 2). This amount of shortening parallel to the GPS velocity
vector is subsequently projected onto the EPGFZ (red arrow, Figure 7 - 2) to construct the minimum
displacement along the EPGFZ.

The Massif de la Selle – Sierra de Bahoruco is subdivided into three parts based on shortening
constrained from cross sections in this study combined with published cross sections (Figure 7 - 2). The
first region, the northern Massif de la Selle, experienced 12% shortening over a distance of 12 km
towards N005°E (perpendicular to the N095°E structural trend). The 12% value is taken from cross
section F because it is constructed more perpendicular to the structural trend and therefore more
representative than the 10% deduced from cross section C. This results in 1.6 km of shortening that,
when projected onto the N085°E slip vector of the EPGFZ, results in 0.9 km of displacement. In the
second region, the northern Sierra de Bahoruco, an average of 20% shortening (cross section G) over
16 km towards N020°E results in 4.0 km of shortening. When projected onto the N085°E slip vector of
the EPGFZ this results in 2.5 km of displacement. In the third region, the northeastern Sierra de
Bahoruco (cross section H), despite having the same structural trend and area of deformation,
shortening is only 16%. When projected on the slip vector of the EPGFZ this results in 2.2 km of
displacement. This region however also experienced shortening in an WNW – ESE direction during the
Miocene, as shown on small-scale cross sections by Van den Berghe (1983). The length of these cross
section is unfortunately too small to allow a direct comparison with the cross sections of Hernaiz
Huerta et al., (2007). The large difference in displacement calculated using the northern Massif de la
Selle and northern Sierra de Bahoruco is possibly explained by additional shortening taken up by
folding-and-thrusting in the Cul-de-Sac Basin, as identified by Saint Fleur et al. (2015).

These calculated displacement values, using shortening from cross sections together with the GPS
velocity vector, are remarkably smaller than the 15 km of maximum displacement as calculated based
on geomorphological constraints in section 4.6. Additional shortening during activity of the EPGFZ is
probably taken up in the southern offshore, in the onshore parts of the Massif de la Selle – Sierra de
Bahoruco, in the Cul-de-Sac – Enriquillo Basin, or in the Chaîne des Matheux – Sierra de Neiba. This will
in turn result in an underestimation of the displacement on the EPGFZ when solemnly looking at the
northern Massif de la Selle – Sierra de Neiba. Alternatively, the amount of shortening calculated using
the cross section in these regions could underestimate the true value due to, for instance, layer-parallel
slip or increased structural complexities that are not captured in these sections. Given these
uncertainties and assumptions, the displacement on the EPGFZ as calculated using shortening in the
northern Massif de la Selle – Sierra de Bahoruco is probably best regarded as a minimum value.

262
Chapter 7: Discussion and regional implications

7.1.4 Average geological shortening rates and implications for SW Hispaniola


Average geological shortening rates for the morphotectonic zones described above are estimated
using the inferred duration of deformation, combined with the distribution of shortening as
constrained in section 7.1.3. These rates are listed in Table 7 - 2, while their geographic distribution is
shown in Figure 7 - 3. Average geological shortening rates are hereinafter abbreviated to ‘shortening
rates’ for brevity.

The timing of activity of the thrust faults bounding the Montagnes Noires and Chaîne des Matheux is
constrained by Pubellier et al. (2000), with the biostratigraphic zonation re-positioned in section 5.2.

Duration of deformation of a thrust sheet in a forward propagating fold-and-thrust belt is interpreted


to last from the onset of activity of the bounding fault associated with the first thrust sheet, until
forward propagation activates the bounding fault of the next thrust sheet, after which the first thrust
sheet is only passively transported. Using the timing of activation of the thrusts in section 5.2 results
in a duration of deformation between 14 and 10 Ma, or 4 Myr, for the Montagnes Noires, and 10 to
2.5 Ma, or 7.5 Myr, for the Chaîne des Matheux (Table 7 - 2). The 2.5 Ma cessation of shortening is
constrained by the <2.3 Ma ages of undeformed Quaternary basalts on the back-limb of the Chaîne
des Matheux at La Vigie (Wadge and Wooden, 1982; Pubellier et al., 1991). The onset of thrusting in
the southern Sierra de Neiba is of late Miocene age (Hernaiz Huerta et al., 2004) and thus similar to
the Chaîne des Matheux. The duration of deformation for the Montagnes Noires and Chaîne des
Matheux extends eastwards in the Dominican Republic (Table 7 - 2). Deformation of the eastern
Peralta Belt is constrained from the late Eocene to the Plio-Pleistocene (Hernaiz Huerta and Pérez-
Estaún, 2002), resulting in a duration of around 32.5 Myr (Table 7 - 2), although this may have occurred
in several pulses (Witschard and Dolan, 1990; Hernaiz Huerta and Pérez-Estaún, 2002). The last phase
of deformation on the Southern Peninsula started in the Tortonian (section 4.2) and continues at
present, with a duration of about 10 Myr (Table 7 - 2). The l’Asile Basin experiences 3.7% shortening
during this time, as shown in section 4.2 and 4.6, and this value is used to calculate the 10 Myr
shortening rate (Table 7 - 2). As discussed in section 7.1.2, deformation in the northern Massif de la
Selle – Sierra de Bahoruco is mainly related to strike-slip activity along the EPGFZ. The timing of focused
strike-slip activity on the EPGFZ is constrained from the late Messinian, or around 6 Ma onwards
(section 4.2), resulting in a duration of deformation of 6 Myr. Within the Cul-de-Sac – Enriquillo Basins,
deformation affects the thickness of Pliocene and younger deposits (cross sections F and G, Appendix
1; Hernaiz Huerta et al. (2007). This roughly limits duration of deformation between 5 Ma and present,
hence at 5 Myr. Using the duration of deformation results in a shortening rate that can be compared
with present-day strain rates derived from GPS velocity models (Figure 7 - 4, Calais et al., 2016).

264
Chapter 7: Discussion and regional implications

Shorte Start Referenc


End Duration Strain rate Referenc
Label Domain ning def. e cross
(Ma) (Myr) (ppb yr-1) e timing
(%) (Ma) section
A Western la Hotte 3.3 10.0 0.0 10.0 3.3 This study Estimated
B L'Asile Basin 3.7 10.0 0.0 10.0 3.7 This study Estimated
C Selle South Flank 1.1 10.0 0.0 10.0 1.1 This study B-88
C Selle North Flank 10.2 6.0 0.0 6.0 17.0 This study B-88
C Selle total 5.4 10.0 0.0 10.0 5.4 This study B-88
D&E Selle South Flank 2.6 10.0 0.0 10.0 2.6 VdB-83 B-88
F Selle North Flank 11.8 6.0 0.0 6.0 19.6 HH-07 B-88
F Enriquillo 3.4 5.0 0.0 5.0 6.8 HH-07 Estimated
Neiba - Matheux
F eq. 16.3 10.0 2.5 7.5 21.7 HH-07 Estimated
Bahoruco North
G Flank 20.3 6.0 0.0 6.0 33.8 HH-07 Estimated
G Enriquillo 4.6 5.0 0.0 5.0 9.2 HH-07 Estimated
Neiba - Matheux
G eq. 23.0 10.0 2.5 7.5 30.6 HH-07 Estimated
G Neiba - Noires eq. 22.3 14.0 10.0 4.0 55.8 HH-07 Estimated
Bahoruco North
H Flank 16.2 6.0 0.0 6.0 26.9 HH-07 Estimated
H Azua 2.5 5.0 0.0 5.0 4.9 HH-07 Estimated
I Peralta Belt 53.8 35.0 2.5 32.5 16.6 HH-02 HH-02
HH-
H&I Martin Garcia 12.3 14.0 10.0 4.0 30.8 02&07 Estimated
J Matheux 30.7 10.0 2.5 7.5 41.0 This study Pub-00
K Cul-de-Sac 5.9 5.0 0.0 5.0 11.8 This study Pub-00
Pub-
K Matheux 23.6 10.0 2.5 7.5 31.5 This study 91&00
Montagnes
L Noires 15.9 14.0 10.0 4.0 39.7 Pub-00 Pub-00
Montagnes
M Noires 15.3 14.0 10.0 4.0 38.2 Pub-00 Pub-00
N Cul-de-Sac eq. 0.2 5.0 0.0 5.0 0.5 Corb-16 Estimated
N Matheux eq. 0.9 10.0 2.5 7.5 1.2 Corb-16 Estimated
Montagnes
N Noires eq. 0.7 14.0 10.0 4.0 1.9 Corb-16 Estimated
Table 7 - 2: Shortening rates in southwestern Hispaniola. Abbreviations are the same as in Table 7 - 1. Additional
abbreviations; B-88 = Bourgueil et al. (1988), Pub-91 = Pubellier et al. (1991).

Shortening rates are calculated by expressing shortening as a dimensionless quantity in parts per billion
(ppb), divided by the duration of shortening. An example calculation is provided next; if a region
experienced 10% of shortening during a 10 Myr period, this can be expressed as a shortening of 0.1
from an initial length of 1.0. This is equivalent to 0.1*109 ppb of shortening. Divided by a duration of
10 Myr results in; 0.1*109 (ppb) / 10*106 (years) = 10 ppb yr-1 shortening rate. This expression allows
for an easy comparison with present-day shortening rates, which are also expressed in ppb yr-1.

265
Chapter 7: Discussion and regional implications

Shortening rates are higher in the Montagnes Noires – northern Sierra de Neiba (45 ppb yr-1) compared
to the Chaîne des Matheux – southern Sierra de Neiba (31 ppb yr-1) (Figure 7 - 3). In the Cul-de-Sac –
Enriquillo – Azua Basins shortening rates are smaller, averaging at 8 ppb yr-1. The Massif de la Hotte
has shortening rates of around 4 ppb yr-1, while the southern Massif de la Selle – Sierra de Bahoruco
has even lower rates of 2 ppb yr-1 (Figure 7 - 3). This is in marked contrast with the northern regions,
with average rates of 18 ppb yr-1 in the west and 30 ppb yr-1 in the east. The shortening rates for the
northern Sierra de Bahoruco are similar to the southern Haitian Fold-and-Thrust Belt (Figure 7 - 3). The
total range of shortening rates ranges from 0.5 to 56 ppb yr-1 (Table 7 - 2). This range of values is similar
to the range of present-day strain rate in southwestern Hispaniola (Figure 7 - 4, Calais et al., 2016),
which ranges from almost zero to around 80 ppb yr-1, although their distribution is markedly different.
At present, the highest values in southwestern Hispaniola are found in the Massif de la Selle region
(Symithe and Calais, 2016). If the Gonâve Island is taken into account, a WNW – ESE trend of high strain
rate is observed (Figure 7 - 4). To the north of this, the Haitian Fold-and-Thrust Belt is mostly situated
within a region of low present-day strain rate. As discussed by Calais et al. (2016), the bulls eye of high
strain rate in the Montagnes Noires (Figure 7 - 4) is probably an artifact due to the distribution of GPS
stations. The difference in geological versus present-day rates for the Haitian Fold-and-Thrust Belt
probably records the southwest-ward propagation of deformation through time. Noticeable are the
high present-day strain rates in the Cul-de-Sac – Enriquillo Basins. Values of around 50 ppb yr-1 in this
area are significantly higher than the average 8 ppb yr-1 shortening rates over the past 5 Myr. This
indicates that over the past 5 Myr, deformation may have progressively increased within these basins,
probably as a response to the southwest-ward migration of deformation in the HFTB on the one hand,
and deformation associated with displacement on the EPGFZ on the other. Calais et al. (2016) interpret
the distribution of present-day strain rates on the Southern Peninsula in the light of deformation
associated with the EPGFZ. However, the apparent WNW – ESE trend of the zone of high strain rates
could also indicate a response to deformation associated with the migration of the active front of the
HFTB. The high present-day strain rates on Gonâve Island, which is the offshore continuation of the
fold-and-thrust belt (Mann et al., 1995) and affected by post-folding-and-thrusting basement inversion
(Ellouz-Zimmermann et al., 2013, 2014), are then related to this progressive southwest-ward migrating
deformation. The area of high present-day strain rates thus had little impact over geological timescales
on the total shortening in these regions, due to the southwest-ward migrating character of associated
deformation.

267
Chapter 7: Discussion and regional implications

7.2 Miocene to present-day paleogeographic reconstruction of the Southern


Peninsula in Haiti

The paleogeographic evolution of the Southern Peninsula of Haiti from the onset of the Miocene to
present-day, as constrained by this study and published data, is presented in this section by 8
paleogeographic maps. These maps focus on the onshore part of the Southern Peninsula and take into
account; 1) north – south shortening and 2) left-lateral displacement along the EPGFZ and associated
strike-slip faults. The only control on rotations in Haiti is in the Ennery region in northwest Haiti and
the Beloc region on the Southern Peninsula (Fossen and Channell, 1988). It is however unclear what
the geographical limits to these rotations are, and it is difficult to verify the quality of these data. For
the paleogeographic reconstructions below, rotations are therefore not taken into account. The
present-day front of the Chaîne des Matheux is kept fixed through time. The northern and southern
coastlines of the Southern Peninsula in Haiti are translated with respect to the front of the Matheux
for every time step. The structural constraints are discussed below and summarized in Table 7 - 3. The
distribution of facies at the different paleogeographic maps is extrapolated from their present-day
distribution as shown on the geological compilation map of the Southern Peninsula in Figure 7 - 5,
combined with information on the timing of deposition and associated facies from the work of Calmus
(1983), Van den Berghe (1983a), Bien-Aime Momplaisir (1986), Bourgueil et al. (1988), and Amilcar
(1997).

7.2.1 Structural constraints

Shortening
The Cul-de-Sac Basin south of the Chaîne des Matheux enjoyed 5.9% of NE – SW shortening between
5 Ma and the present-day (section 7.1). The width of the basin parallel to the direction of shortening
is around 20 km (Figure 7 - 1), resulting in 1.2 km of total shortening. In the Massif de la Selle, 5.4% of
total north - south shortening (section 4.6) over 35 km width (Figure 7 - 1) corresponds to 1.9 km of
total shortening since 10 Ma (Table 7 - 3a). For the l’Asile Basin, 3.7% north – south shortening since
10 Ma (section 4.2) over a width of 33 km (Figure 7 - 1) corresponds to 1.2 km of shortening (Table 7 -
3a). Similarly, in the Massif de la Hotte, 3.3% shortening since 10 Ma (section 7.1) over a width of 43
km (Figure 7 - 1) corresponds to 1.4 km of shortening (Table 7 - 3a). The total shortening in the Cul-de-
Sac Basin and Southern Peninsula is about 2 kilometers or less for the past 10 Ma. On the scale of the
Southern Peninsula, such small values fall within the cartographic margin of error and are thus not
used in the reconstructions.

268
Chapter 7: Discussion and regional implications

Strike-slip displacement
Strike-slip displacement related to the presently active trace of the EPGFZ is estimated in the Massif
de la Hotte and Massif de la Selle at a maximum of 12 and 15 km, respectively (section 4.4).
Displacement in the Massif de la Selle in the east is better constrained than in the Massif de la Hotte
in the west, and the value of 15 km is therefore used in the reconstructions. Strike-slip activity in the
Southern Peninsula became focused along the EPGFZ from around 6 Ma onwards (section 4.2). The
average displacement rate during these 6 Myr is thus 2.5 km Myr-1, or 2.5 mm yr-1. Present-day EPGFZ-
parallel left-lateral displacement rates (section 4.2), as constrained by GPS velocity modeling, averages
around 8 mm yr-1 for the Southern Peninsula (Benford et al., 2012a; Calais et al., 2016). This is
significantly larger than the 2.5 mm yr-1 rates deduced for the EPGFZ.

These studies however use discrete blocks to model the GPS velocities, and their block boundaries
coincide with the EPGFZ on the Southern Peninsula and the LPSJFZ in the Cordillera Central. The EPGFZ-
parallel displacement rates thus not only take into account active displacement on the EPGFZ itself,
but accommodate all EPGFZ-parallel strike-slip displacements between the southern part of the
Southern Peninsula and the region southwest of the LPSJFZ. Several inactive or mildly active strike-slip
fault strands are observed onshore, on the Southern Peninsula (section 4.2). In the offshore domain
between Gonâve Island and the Southern Peninsula, strike-slip related structures are identified on
seismic data (Bien-Aime Momplaisir, 1986; Mann et al., 1995). Pliocene and younger strike-slip activity
has been evidenced along the southern flank of the Chaîne des Matheux (chapter 5) and on the
southern flank of the Sierra de Neiba (Hernaiz Huerta et al., 2007). Assuming that the timing of onset
of focused strike-slip activity on the EPGFZ at 6 Ma is correct (section 4.2), which results in 2.5 mm
yr-1 slip rate, around 5.5 mm yr-1 of slip rate is missing when only taking the active trace of the EPGFZ
into account. This 5.5 mm yr-1 is probably accommodated on the structures identified above.

The next question is how representative the present-day 8 mm yr-1 displacement rate is over geological
time scales. Relative motion between tectonic plates is the first-order control for slip rate on faults
(Mueller, 2017). This implies that as long as the relative velocity vector between blocks on either side
of a fault remains the same, the slip rate on the fault will not change significantly over geological time.
Walker et al. (2010) show that Holocene slip rates on a strike-slip fault in eastern Iran are similar to
present-day geodetic (GPS-derived) slip rates. For more complicated settings, Walls et al. (1998) show
that only with an increased understanding of fault architecture and interaction, the combined
Holocene slip rates on the San Andreas Fault system match present-day geodetic measurements. For
the Septentrional Fault Zone in northern Hispaniola, Prentice et al. (2003) find an average late
Holocene slip rate of 9 – 11 mm yr-1, which is in close agreement to present-day fault-parallel GPS
velocities of 9 – 10 mm yr-1 (Calais et al., 2016). Over larger geological time scales for the San Andreas

270
Chapter 7: Discussion and regional implications

Fault (SAF), Ehlert (2003) arrive at 240 km displacement over the past 5 Myr (48 mm yr-1), which is
similar to the slip rate found by Darin and Dorsey (2013) of 40 ± 3 mm yr-1 (200 ± 14 km since 5 Ma).
These values are in close agreement with both long term (3 Myr) 50 mm yr-1 average SAF slip rates as
constrained by sea floor spreading rates (DeMets and Dixon, 1999) and present-day SAF slip rates of
48 ± 5 mm yr-1 (Bennett et al., 1996). It appears that, as long as the geodynamic characteristics of a
system remain unchanged, present-day geodetic slip rates are representative for longer geological
time scales. In the case of the EPGFZ, extrapolating the 8 mm yr-1 geodetic slip rate to the onset of
focused strike-slip activity on the EPGFZ at around 6 Ma, results in a total displacement on the system
of 48 km. Of this displacement, a maximum of 15 km is accommodated on the presently active trace
of the EPGFZ (section 4.6), while the remaining 33 km is possibly accommodated by slip on other strike-
slip or oblique-slip faults (Table 7 - 3b).

Based on geological constraints outlined in section 4.2, it is however expected that strike-slip activity
commenced during the late Miocene, at around 10 Ma. There appears to be a period of 4 Myr before
the EPGFZ became well established. During this 4 Myr period the slip rate of the entire strike-slip
system possibly increased from zero to 8 mm yr-1. Lacking any control on how this slip rate evolved
during the 4 Myr period, it is here assumed that slip rates increased linearly. For a period of 4 Myr this
results in an additional 16 km of displacement (Table 7 - 3b). Therefore, a total of approximately 64 km
of displacement were accommodated along the strike-slip system since its initiation at around 10 Ma.
It is interesting to note that this total displacement is a close match to the 65 km of displacement
predicted for the EPGFZ, assuming that the system stretches from Jamaica to the Dominican Republic
and has a length of around 500 km, by using the global average displacement versus fault-length on
strike-slip systems (section 4.4, Figure 4 - 5; Kim and Sanderson, 2005).

To simplify the paleogeographic reconstruction, all of the strike-slip displacements are assumed to be
accommodated between the block south of the EPGFZ and the southern front of the Chaîne des
Matheux. For the interval between 0 and 6 Ma this is achieved by calculating the displacement
between the block south of the EPGFZ and the front of the Matheux, and subsequently translating the
southern coastline of the Southern Peninsula to the calculated position. Slip on the EPGFZ is computed
next, assuming a fixed slip rate of 2.5 mm yr-1, and the western coastline of the Southern Peninsula is
displaced accordingly. The northern coastline of the Southern Peninsula is then assigned an
intermediate position. Between 6 and 10 Ma the process is the same, but a decreasing slip rate from
8 to 0 mm yr-1 over this period is used. The resultant displacements are shown in Table 7 - 4 for every
time step.

271
Chapter 7: Discussion and regional implications

A: SHORTENING
Start Shortening
Shortening Width Shortening End Duration
Domain Trend def. rate (mm yr-1
(%) (Km) (Km) (Ma) (Myr)
(Ma) = km Myr-1)
Cul-de- NE -
5.9 20 1.2 5.0 0.0 5.0 0.2
Sac SW
Massif de
la Selle
5.4 N - S 35 1.9 10 0 10.0 0.2
(entire
section)
L'Asile 3.7 N - S 33 1.2 10 0 10.0 0.1
Basin 1.3 N - S 33 0.4 23 20 3.0 0.1
Massif de
3.3 N - S 43 1.4 10 0 10.0 0.1
la Hotte

B: STRIKE-SLIP
Start Displacement
Displaceme End Duration
Domain def. rate (mm yr-1
nt (Km) (Ma) (Myr)
(Ma) = km Myr-1)
EPGFZ
Massif de 15 6 0 6.0 2.5
la Selle
Additiona 33 6 0 6.0 5.5
l strike-
slip 16 10 6 4.0 4.0
Table 7 - 3: Shortening and strike-slip displacement on the Southern Peninsula.

Displacement
Displacement rate
Domain rate (km/Myr), 2.5 Ma 5 Ma 7.5 Ma 10 Ma
(km/Myr), 6 - 10 Ma
0 - 6 Ma
EPGFZ Massif de
2.5 6 13
la Selle
8.0 to 0.0 58 64
Additional
5.5 14 28
strike-slip
Table 7 - 4: Displacement along the EPGFZ and associated strike-slip faults.

272
Chapter 7: Discussion and regional implications

7.2.2 Paleogeographic reconstruction maps


Present-day
The present-day setting is shown in Figure 7 - 6a. Continental clastics, such as conglomerates and
sands, are fed onto the Cul-de-Sac by various river systems. Lagoonal facies are restricted to Lac Azuei
and Trou Caïman. On the Southern Peninsula, continental clastics are deposited in the Clonard and
Miragoane pull-apart basins. Additional depocenters for continental clastics along the EPGFZ are the
l’Asile basin and an unnamed basin further east. In the Léogâne, Les Cayes, Jacmel and Cavaillon
regions, the input of continental clastics is controlled by major river systems.

Onset of Pleistocene, ~2.5 Ma


The paleogeography at 2.5 Ma (Figure 7 - 6b), at the onset of the Quaternary, is similar to the present-
day, but with less topographic elevation. Displacement along the EPGFZ with respect to the present –
day position is around 6 km (Table 7 - 4), with an additional 14 km to the north of the EPGFZ (Table 7 -
4). According to Mann et al. (1983, 1995) the sedimentary fill of the Miragoane and Clonard pull-apart
basins is Quaternary in age, although this is not constrained by biostratigraphy. Sedimentation in the
Cul-de-Sac basin is characterized by shallow lagoonal facies (Bourgueil et al., 1988).

Onset of Pliocene, ~5 Ma
At 5 Ma (Figure 7 - 6c), at the onset of the Pliocene, erosion reached the Cretaceous basalts in the
Massif de la Selle and presumably in the Massif de la Hotte (Bourgueil et al., 1988). Displacement along
the EPGFZ is around 13 km, with an additional 28 km accommodated further north (Table 7 - 4).
Sedimentation in the Cul-de-Sac basin is characterized by bathyal marine facies (Bourgueil et al., 1988).
Marine incursions along the southern coast are controlled by NW – SE trending structural highs and
lows. Total emerged and exposed surface area of the Southern Peninsula is smaller compared to 2.5
Ma.

Onset of Messinian, ~7.5 Ma


At 7.5 Ma (Figure 7 - 6d), much of the Southern Peninsula was characterized by marine marl and
limestones sedimentation (Calmus, 1983; Van den Berghe, 1983b; Bourgueil et al., 1988). The onset of
the Messinian roughly coincides with the onset of compressional deformation in the Southern
Peninsula (section 4.2). The peninsula is probably undergoing uplift during this time (Bourgueil et al.,
1988), although it is unclear if subaerial erosion is already taking place. In this map the western Massif
de la Hotte and Massif de la Selle are projected as being subject to erosion along their central axes.
Since 6 Ma strike-slip deformation is not focused along the EPGFZ, but distributed over the Southern
Peninsula (section 4.2). Cumulative left-lateral slip is around 58 km (Table 7 - 4).

273
Chapter 7: Discussion and regional implications

Tortonian, ~10 Ma
At 10 Ma (Figure 7 - 6e), during the Tortonian, the Southern Peninsula is dominated by marine facies
(Calmus, 1983; Van den Berghe, 1983b; Bien-Aime Momplaisir, 1986; Bourgueil et al., 1988). This
timing roughly coincides with the onset of distributed strike-slip activity on the Southern Peninsula
(section 4.2). Total cumulative left-lateral displacement at 10 Ma is around 64 km (Table 7 - 4).

Langhian, ~15 Ma
At 15 Ma (Figure 7 - 6f), during Langhian times, much of the island is the site of marine sedimentation
(Calmus, 1983; Van den Berghe, 1983b; Bourgueil et al., 1988). The southwestern region is still
subaerially exposed as a result of the early Miocene deformation phase, with marine sedimentation in
this area only commencing during the Langhian or Serravallian (section 4.2; Amilcar, 1997). The Camp
Perrin region is dominated by lacustrine and lagoonal sedimentation (Calmus, 1983). According to
Calmus (1983) these clastics are sourced from an elevated region to the north of the basin. This is
unlikely since no erosion-derived sediments of similar age are found further north or east (Maurrasse,
1982; Calmus, 1983; Bourgueil et al., 1988; Amilcar, 1997). It is therefore more likely that the source
region of the Camp Perrin clastics is located towards the west, as shown on the reconstruction.

Burdigalian, ~18 Ma
At 18 Ma (Figure 7 - 6g), during the Burdigalian, the Camp Perrin (Calmus, 1983) and l’Asile (Bien-Aime
Momplaisir, 1986) regions are subject to continental or lagoonal deposition, while the remainder of
the island is subject to marine sedimentation (Calmus, 1983; Van den Berghe, 1983b; Bourgueil et al.,
1988). The continental conglomerates around Camp Perrin contain clasts of Cretaceous CLIP basalts
and Cretaceous Macaya Formation limestones (Calmus, 1983), and are probably sourced by the
exposed reliefs to the west.

Aquitanian, ~22 Ma
At 22 Ma (Figure 7 - 6h), during the Aquitanian, an erosional unconformity develops over the western
part of the Southern Peninsula, while the northwestern Jeremie region and eastern domains remain
the site of continued marine sedimentation (section 4.2; Calmus, 1983; Van den Berghe, 1983b; Bien-
Aime Momplaisir, 1986; Bourgueil et al., 1988; Amilcar, 1997). Uplift and shortening occurred over
folds with NW – SE trending axes (section 4.2). The amount of shortening during this period is around
1.3%, or 0.4 km (Table 7 - 3), which is too small to be incorporated in the paleogeographic
reconstruction.

275
Chapter 7: Discussion and regional implications

7.3 Crustal-scale cross-section


The aim of this section is to compile a crustal-scale cross section from the Massif de la Selle to the
Chaîne des Matheux, using the cross sections constructed in 4.2 and 5.5, and identify possible fluid
pathways and source regions.

7.3.1 Crustal structure


Based on a teleseismic receiver function study, Corbeau et al. (2017) determined the depth of the
Moho in Haiti (Figure 7 - 7). Moho depth is relative to the seismic station and thus to the earth’s
surface. These authors identify three distinct crustal domains. The first domain is the Southern
Peninsula, characterized by an average Moho depth of 22 ± 5 km and Vp/Vs of 1.80 ± 0.03 km s -1, the
latter indicating intermediate to mafic crust (Figure 7 - 7, Corbeau et al., 2017). The second domain
comprises the Cretaceous island-arc and Peralta Belt, characterized by an average Moho depth of 23
± 3 km and Vp/Vs of 1.75 ± 0.10 km s-1, the latter indicating felsic to intermediate crust (Figure 7 - 7,
Corbeau et al., 2017). The third domain encompasses the Haitian Fold-and-Thrust Belt and the Cul-de-
Sac Basin, characterized by an average Moho depth of 41 ± 4 km and Vp/Vs of 1.80 ± 0.05 km s -1, the
latter indicating intermediate to mafic crust (Figure 7 - 7, Corbeau et al., 2017).

Massif de la Selle
The Southern Peninsula is considered as an analogue of the thickened oceanic crust of the interior
Caribbean plate and the Beata Ridge (Mauffret and Leroy, 1997). The simplified crustal stratigraphy of
the Caribbean Large Igneous Province, based on the work by Mauffret and Leroy (1997), Diebold et al.
(1999), Révillon et al. (2000), Mauffret et al. (2001) is shown in Figure 7 - 8. Thickness of deep marine
sediments that accumulated on the original oceanic crust prior to flood basalt volcanism in the Late
Cretaceous is calculated at 500m, using a 10m Myr-1 accumulation rate typical for deep sea sediments
(Einsele, 2000) over a 50 Myr period (Figure 7 - 8). This thickness is integrated into the original crustal
thickness on the cross section. The original oceanic crustal section, between reflectors B” and R, is
considered here to have a constant thickness for the southern Massif de la Selle and the Beata Ridge.
The thickness of ultramafic underplated material is varied to match the present-day Moho depth from
Corbeau et al. (2017).

276
Chapter 7: Discussion and regional implications

Cul-de-Sac – Chaîne des Matheux


The crust under the Cul-de-Sac and Chaîne des Matheux is not outcropping and therefore only
characterized using geophysical methods. According to Corbeau et al. (2017), this crust is of an island-
arc type underplated by ultramafic material. It is however unlikely that this crust is related to the
Cretaceous island arc, as explained below. The thickness of the extinct arc in the Massif du Nord is 23
± 3 km (Corbeau et al., 2017). In palinspastic reconstructions (Pindell and Kennan, 2009; Boschman et
al., 2014), the region with the anomalously thick crust is always in a back-arc position located to the
southwest of the Caribbean island arc. The average thickness of the underplated material in the Massif
de la Selle, assuming a constant original crustal thickness of around 9 km (Figure 7 - 8), is 13 ± 5 km. If
the thickness of underplated material is similar under central Haiti, this leaves a thickness of 28 ± 5 for
the unknown crust. This is significantly thicker than the 23 ± 3 km for the island-arc crust under the
Massif du Nord. It is unlikely that the back-arc region, which is less intruded compared to the island-
arc, would have a larger crustal thickness than the central part of the island-arc. Another possibility,
also noted by Corbeau et al. (2017), is that this thickened crust is continental. Kamenov et al. (2011)
concludes that the Quaternary basalts in the Chaîne des Matheux are sourced from sub-continental
lithospheric mantle, which these authors relate to Grenvillian continental terranes in Central America.
Palinspastic reconstructions (Pindell and Kennan, 2009; Boschman et al., 2014) show that the leading
edge of the Caribbean island-arc swept past the Yucatan Peninsula during the Late Cretaceous (section
2.6; Stanek et al., 2009), with the Siuna and Chortis terranes trailing behind. It is possible that during
this process, a fragment of continental lithosphere became detached from the Siuna /Chortis terranes
or the Yucatan Peninsula. This continental material became juxtaposed against the Belize margin, and
subsequently thinned, rifted, and drifted away due to spreading in the Cayman Trough. Seismic
refraction studies of the eastern Cayman Trough (section 2.3; Mauffret and Leroy, 1997; Leroy et al.,
2000) show an eastward increase in crustal thickness into the Golf of Gonâve. This trend possibly
continues towards the Haitian Fold-and-Thrust Belt. Continental material alone however cannot
explain the Vp/Vs ratios of 1.80 for this region (Corbeau et al., 2017). It is therefore probable that
(ultra)mafic high density, high Vp velocity material is present underneath the continental crust. This
material could be emplaced during CLIP volcanism and therefore of magmatic origin, as is inferred for
the underplated ultramafic picrites and cumulates under the Beata Ridge (Mauffret and Leroy, 1997).
Another possibility is emplacement of oceanic-type crust by underthrusting or subduction, as indicated
by the dashed black line in Figure 7 - 10. The region south of the Yucatan Peninsula experienced
convergence and underthrusting during the Late Cretaceous (Pindell and Kennan, 2009), making the
underthrusting hypothesis more plausible.

278
Chapter 7: Discussion and regional implications

locus of flood basalt extrusion was located to south in the Late Cretaceous (Pindell and Kennan, 2009).
Overlying the basalts in the Chaîne des Matheux are Upper Cretaceous marine sediments, separated
from the Paleocene by an unconformity (section 5.1.1). The thickness of this pre-Eocene sequence is
unknown. For lack of a better constraint, the thickness of the Paleocene is taken from the northern
Massif de la Selle. The Maastrichtian flysch is a fault-bounded package in the western part of the
Chaîne des Matheux (section 5.1.1, Vila et al., 1988). The thickness of the Upper Cretaceous sediments
is taken such that the basal decollement level, southwest of the Beauger Thrust (BT), is located within
them. The Beauger Thrust in turn is possibly rooted in sediments that underlie the CLIP basalts in this
area, although there is no control on their thickness or lithology. For simplicity this sedimentary
sequence is therefore incorporated within the continental crustal sequence.

The Moho depths are taken from Corbeau et al. (2017) and projected onto the cross section by a line
perpendicular to the trend of the section (Figure 7 - 9). Noticeable is the large difference in Moho
depth between the FURC, FERM and PAPH stations (Figure 7 - 10). In this interpretation the EPGFZ is
partly responsible for the difference in Moho depth between FERM and PAPH. The difference between
FURC and FERM is resolved by a shallow oblique thrust, which connects with the thrust in the southern
offshore (section 4.2). The resulting geometry is an asymmetric pop-up structure under the Massif de
la Selle, resembling the geometry of the asymmetric Massif de la Selle at the surface (Figure 7 - 10).
The apparent vertical offset along the EPGFZ is much larger at depth than at the surface. This can at
least partly be explained by horizontal displacement along this fault, in combination with the large
lateral variation in Moho depth (i.e. JACM versus GRBO, 9 km Moho depth difference over 16 km
distance, Figure 7 - 9). Although the EPGFZ at the location of the cross section has a large control on
the difference in Moho depth, along its length the EPGFZ does not act as a boundary for the different
crustal domains. On Figure 7 - 7 a relatively shallow Moho depth is reported northwest of the EPGFZ
by Corbeau et al. (2017), indicating that the EPGFZ does not follow the boundary between the different
crustal domains.

The thickness of (ultra)mafic material is decreasing towards the northeast of this section. This either
reflects the decreasing thickness of the CLIP basalts, which are taken as a proxy for the thickness of the
magmatic underplated material, or reflects the limit of underthrusted oceanic crust, depending on the
favored hypothesis. To counter the decrease in thickness of (ultra)mafic material, thickness of the
continental crust must increase to the northeast. This is achieved by introducing a number of normal
faults, which reflect an ancient transitional domain between continental and oceanic crust.

Projected onto the cross section is the location of the Quaternary Morne La Vigie and Thomazeau
volcanic center (Figure 7 - 10), which lie roughly on the same structural trend (Figure 7 - 9). This

280
Chapter 7: Discussion and regional implications

volcanism is possibly related to activity on transtensional NE – SW trending faults (Pubellier et al., 1991,
2000). The responsible faults are in the same plane as the cross section and thus not shown. Based on
the surface geology the horizontal offset along these faults is limited, and it is unknown if these minor
faults can coalesce to create pathways for the magma to rise to the surface. In either case, a pathway
for these basalts must have existed, since they are extruded at the surface.

7.3.3 Fluid flow pathways


The geochemistry of calcite veins sampled from the core of the EPGFZ in the Massif de la Selle possibly
indicates a contamination of formational waters with fluids derived from a warmer source, which are
transported from depth along the EPGFZ (chapter 6). Helium isotope ratios of water samples from a
spring in the western Massif de la Selle, in the Fortin area (Figure 7 - 9), indicate a pronounced mantle
component (Battani et al., 2016).

Although it is unclear if this source is related to the EPGFZ, the NW – SE trending fault in the Jacmel
valley, or both, it shows evidence for a fluid pathway between the mantle and the surface. Based on
shear-wave splitting Benford et al. (2012b) demonstrate that the EPGFZ extends below the Moho into
the lithospheric mantle. The EPGFZ is thus a probable pathway for fluids derived from the upper mantle
or deep crustal levels. This pathway possibly developed during the Pliocene (section 4.2, chapter 6)
and is still active, at least locally, today (Battani et al., 2016). In the western part of the Southern
Peninsula, Battani et al. (2016) obtained similar helium ratios from spring fluids along strike of the
EPGFZ, further testifying to the deep fluid pathways.

The geochemistry of the limited number of samples taken from faults in the Chaîne des Matheux do
not show evidence of fluids with a deep origin, but rather indicate infiltration of the fault zones by
meteoric waters. Water samples collected from the Source Puantes springs along the front of the
Chaîne des Matheux (Figure 7 - 9) indicate a very strong component of mantle-derived helium (Battani
et al., 2016). When projected onto the cross sections this source is located around the southwestern
tip of the fold-and-thrust belt. Similar faults along the front of the Sierra de Neiba in the Enriquillo
Basin deform sediments as young as the Quaternary (Hernaiz Huerta et al., 2007), indicating that they
are possibly still active today.

281
Chapter 7: Discussion and regional implications

There are a number of possible pathways and mechanisms to explain the mantle component of the
fluids in the Sources Puantes region. The first mechanism (Figure 7 - 10 - pathway A) is in direct relation
to the basaltic volcanism in the area. K-Ar ages as young as 1.6 Ma are reported for these basalts
(Wadge and Wooden, 1982; Pubellier et al., 1991). A possible pathway for fluids to ascend from the
mantle is along the feeder system for this volcanism. These fluids could then possibly leak into the
detachment fault and are subsequently transported to the active thrust front, which roughly coincides
with the spring’s projected location at the surface (Figure 7 - 10). A second possible pathway is located
along the entire length of the decollement level, assuming that it descends through the entire crust
and into the upper mantle further northeast (Figure 7 - 10, pathway B). A third possibility is a blind,
steeply dipping, deeply rooted thrust fault under the Chaîne des Matheux (Figure 7 - 10, pathway C).

The lack of a deep fluid component from vein samples along the faults in the Chaîne des Matheux
favors the first hypothesis. If activity of thrust faults in the central part of the Chaîne des Matheux pre-
dates extrusion of the Quaternary basalts, fluids related to this volcanism cannot have contributed to
the fluid composition from which the mineralizations along these thrusts precipitated. Volcanism only
took place once the active deformation front had already migrated into the Cul-de-Sac area, with the
active thrusts acting as a pathway for fluids transported from the mantle along the feeder systems of
the basalts. Based on the available data it is not possibly to exclude any deeply-rooted thrusts, if these
exist under the Chaîne des Matheux, as possible fluid pathways.

7.3.4 Geodynamic implications


The nature of the crust under the Chaîne des Matheux has important geodynamic implications,
depending on the favored hypothesis, which are outlined below.

If the crust is predominantly continental and underplated by ultramafic material related to CLIP
volcanism, this crustal block must have been, at least partly, located inside the domain of flood basalt
volcanism in the Late Cretaceous. Palinspastic reconstruction (Pindell and Kennan, 2009) appear to
exclude the Yucatan Peninsula as the parent of this continental block, since it was located north of the
Cretaceous island-arc during the time of CLIP volcanism. Based on the same reconstructions (section
2.4; Pindell and Kennan, 2009), the Siuna terrane is the most likely origin for this block, in spite of the
high density, high velocity material under the continental block.

In either case, this continental block was subsequently transported and ended up adjacent to the Belize
margin at the onset of rifting in the Cayman Trough, since it is presently located east of this spreading
system. As outlined in section 2.6, an estimated 600 km of crustal material is consumed between the
Montagnes Noires – Sierra de Neiba and the Cordillera Central, for which Eocene to Oligocene oblique

283
Chapter 7: Discussion and regional implications

strike-slip activity along the southwestern flank is the most likely mechanism. This missing crustal
material must have been attached to the continental block, since this thick crustal block is currently
directly adjacent to the Cordillera Central. It is proposed here that this missing crust had a different
nature than the continental block. In this hypothesis, during convergence between the continental
block and the island-arc crust of the Cordillera Central, the unknown crustal material became removed
during the Eocene and Oligocene. At the onset of the Miocene the crustal block progressively collided
and sutured with the island-arc terrane of the Cordillera Central. The increased collision detached
Hispaniola from Cuba along the Oriente Fault Zone and resulted in a change of fault dynamics, from
predominantly strike-slip to folding-and-thrusting. During the late Miocene, related to increased
collision of northern Hispaniola with the Bahamas platform, the relative velocity between the
Caribbean and North American Plate became too large to be solemnly resolved via the Oriente Fault
Zone and the Haitian Fold-and-Thrust Belt, and the remainder of relative motion became progressively
accommodated on the EPGFZ. This scenario also explains why the EPGFZ became active, which is not
to be expected from a plate tectonics point of view (section 4.2), since the Motagua Fault Zone – Swan
Island Fault Zone – Mid-Cayman Spreading Center – Septentrional-Oriente Fault Zone – North
Hispaniola Deformed Belt system (section 2.1, Figure 2 - 1) is sufficient to accommodate relative
motion between the Caribbean and the North American plates.

284
Chapter 8: Conclusions and perspectives

8 Conclusions and perspectives

8.1 Conclusions
The main conclusions from this thesis are separated for the Southern Peninsula and Chaîne des
Matheux.

8.1.1 Southern Peninsula


The Cenozoic tectonic history of the Southern Peninsula in Haiti is polyphase and characterized by
three major tectonic events.

1. Deformation and uplift during the Maastrichtian to early Paleocene by NNE-directed compression
resulted in around 5% of shortening aided by crustal-scale folding and SSW-verging thrusts. These
deeply rooted thrusts possibly acted as pathways for hot formational fluids from deeper
stratigraphic levels. Associated uplift caused partial erosion of the Upper Cretaceous sedimentary
cover and CLIP basalts in large parts of the Southern Peninsula. During this deformation phase the
Southern Peninsula was likely situated south of the Siuna deformed belt (south of Yucatan).
2. Early Miocene NE-directed compression mainly affected the western and central parts of the
Southern Peninsula. In the central Southern Peninsula only 1.3% of shortening is related to this
phase. Joints that developed by layer-parallel shortening resulted in local intra-formational fluid
circulation. Uplift and erosion associated with this phase of deformation was most pronounced in
the southwestern part of the Southern Peninsula, with erosion decreasing eastwards and
continuing sedimentation in the Massif de la Selle.
3. Transpressive deformation from late Miocene to recent. Strike-slip activity commenced in late
Miocene times, possibly concomitant with uplift of the Southern Peninsula. Shortening from the
late Miocene to recent averages around 5% over the entire Southern Peninsula. Strike-slip activity
became progressively more focused along the present-day active trace of the EPGFZ from the late
Messinian onwards, with the development of pull-apart basins along its trace during the
Quaternary, possibly already during the Pliocene. The tectonic style along the EPGFZ changes from
predominantly partitioned strike-slip in the west to oblique-slip in the east. Oblique thrust in the
northern Massif de la Selle are related to the EPGFZ and rooted on this fault at depth. A maximum
total of 15 km of left-lateral displacement is recorded along the EPGFZ, which is mainly
accommodated by shortening related to oblique thrusts in the northern Massif de la Selle – Sierra
de Bahoruco. An additional 50 km of left-lateral slip is possibly taken up by other strike-slip faults
on the Southern Peninsula and in the northern offshore. Maximum principal paleo-stress axes
oriented at high angles to the trace of the EPGFZ indicate that the fault zone is mechanically weak.

285
Chapter 8: Conclusions and perspectives

The EPGFZ is rooted in the upper mantle and acted as a pathway for fluids throughout much of its
active history. Thrust faults that are rooted on the EPGFZ and minor strike-slip faults related to the
Miragoane pull-apart basin do not show evidence of deep fluid transport, but mainly acted as
pathways for meteoric surface waters to percolate downwards.

8.1.2 Chaîne des Matheux


The evolution of the Chaîne des Matheux as part of the Haitian Fold-and-Thrust belt from the Miocene
onwards can be broken down into 4 stages.

1. Layer-parallel shortening. This affected the Chaîne des Matheux area prior to late Miocene times,
possibly in response to far-field stresses coeval with folding-and-thrusting in the Montagnes Noires
further north. Joints that resulted from this NE-directed compression provided pathways for the
intra-formational circulation of fluids with anomalously high temperatures.
2. A short-lived NE – SW directed strike-slip regime reactivated these joints during the Serravallian.
Minor NE – SW trending normal faults are possibly related to this phase. These normal faults
created pathways for the upwards migration of slightly warmer formational fluids.
3. Folding-and-thrusting in the Chaîne des Matheux from the Tortonian onwards was predominantly
thin-skinned and characterized by shallow thrusts and decollement levels that ramp upwards
through the stratigraphy. This deformation was possibly aided by deeply rooted thrusts and thick-
skinned deformation, although there is no conclusive evidence to support this. Shortening during
the fold-and-thrust phase varies between 24 and 30%, depending on the along-strike location.
There is no evidence for the transport of deep basement-derived fluids along the thrusts, which
rather acted as pathways for the downward percolation of meteoric surface waters that mixed
with formational waters. During the Pliocene the locus of folding-and-thrusting migrated
southwest-wards, and activity on the thrusts in the Chaîne des Matheux had terminated at the
onset of the Pleistocene.
4. Quaternary transtensional strike-slip activity along the southern front of the Chaîne des Matheux
possibly aided the Quaternary volcanism in the region, which created pathways for deep, possibly
mantle-derived fluids to ascent to the surface.

The crust under the Chaîne des Matheux is possibly continental in origin and derived from the Siuna
terrane south of the Yucatan Peninsula. Increased collision at the onset of the Miocene between this
continental block and the Cordillera Central resulted in detachment of Hispaniola from Cuba along the
OFZ and the initiation of the Haitian Fold-and-Thrust Belt. During the late Miocene this system could
not accommodate all the relative motion between the Caribbean and North American plates, resulting
in the creation of the EPGFZ and the establishment of a new plate boundary.

286
Chapter 8: Conclusions and perspectives

8.2 Perspectives
One of the difficulties encountered during the field campaigns was to correlate the geological maps
and stratigraphic information with outcrops in the field. This is partly related to the scale and precision
of the available 1:250.000 geological maps for Haiti (Bien-Aime Momplaisir et al., 1988; Boisson and
Bien-Aime Momplaisir, 1987; Boisson and Pubellier, 1987, 1988) and the uncertainties of spatial and
temporal distribution of the Cretaceous and Cenozoic sediments. Another limiting factor is the outcrop
quality and distribution. This is mainly due to intense weathering and vegetation cover, especially for
the Southern Peninsula, where it is often difficult to correlate individual outcrops. While fully
appreciating the efforts, resources, and time spend in the mid-80’s to create the presently available
geological map sheets of Haiti, which are a vast improvement over the pre-1980 geological map,
additional work is necessary. In order to make further advances in understanding the geology of Haiti,
locate natural resources such as potable water, ores, and hydrocarbons, and to identify landslide and
earthquake risk, it is of paramount importance to perform a country-wide, smaller-scale geological
mapping exercise. This could be done with a similar construction as used for the Dominican Republic,
where a consortium of universities and geological surveys, with funding from the European Union,
created 1:50.000 and 1:100.000 geological maps and significantly improved the stratigraphy. High
quality, small-scale geological maps for the geology in Haiti will also make it easier to compare and
correlate the geology with the Dominican Republic, since most of the geological structures are
continuous and found in both countries.

Even without improved geological maps, advances can be made in other areas. One of these is related
to the paleogeographic position of the different fault-bounded morphotectonic zones in Haiti and their
palinspastic evolution. As pointed out in section 2.6, there is still great uncertainty about the
palinspastic position of the Southern Peninsula through time with respect to central Hispaniola, and
by what means these two terranes converged. At present, the only paleomagnetic study available in
Haiti is from Fossen and Channell (1988), which provides very limited information on the
paleogeographic position of the Beloc and Ennery regions in the Paleocene and Eocene, respectively.
To understand the palinspastic positions through time of the Massif du Nord, Montagnes Noires,
Chaîne des Matheux, Massif de la Selle and Massif du Hotte, additional paleomagnetic studies are
necessary. Ideally this would involve sampling per region of multiple stratigraphic intervals (and thus
multiple ages), which should be constrained by biostratigraphy or radiometric dating where possible.

These results can then be used, together with the outlines of the morphotectonic terranes and their
shortening over time as proposed in chapter 7, as input for palinspastic reconstruction software, such
as GPlates. A large-scale, northern Caribbean GPlates model already exist due to the work by
Boschman et al. (2014), which can provide regional constraints to fit the smaller-scale model into. The

287
Chapter 8: Conclusions and perspectives

resulting spatial evolution of the different blocks through time can provide useful insights for
paleogeographic maps and might serve to identify areas with higher earthquake risk.

Another area where advances can be made is in better characterizing the seismogenic fault zones in
southern Haiti, to the obvious end of better constraining earthquake risk associated with them.
Although chapter 5 and 6 address the characteristics of the EPGFZ, our understanding of its
characteristics and evolution is limited by; 1) the number of geochemical samples taken from the fault
core and damage zone, and 2) the limited spatial resolution of observations and measurements from
it. There are two ways of addressing this. The first involves studying the exhumed parts of the EPGFZ
and its older fault strands in outcrops. This would benefit by creating small cross sections across the
fault zone to determine the width of the fault core and fault damage zone in different lithologies and
along-strike. Supplementary sampling of veins and host rock can shed additional light on whether or
not this fault zone was connected with deeper fluid reservoirs during its evolution, or if this is a recent
development (see chapter 7 for discussion). A limiting factor, as already encountered during this study,
is the scarcity of continuous outcrops that expose the EPGFZ. This problem can be circumvented by
means of drilling the active, in-situ parts of the EPGFZ. Such a project has successfully been executed
along the San Andreas Fault (i.e. Zoback et al., 2011), resulting in a better understanding of the
geometry of the fault, slip rates, seismic recurrence times, stress states, and fluid flow characteristics.
Ideally this would be supplemented by high resolution onshore seismic data, preferentially acquired
over the Cul-de-Sac basin. This will visualize the deformation front of the Chaîne des Matheux, the
structures under the Cul-de-Sac Basin proper, and the faults in the southern part of this basin that are
associated with the EPGFZ. A scientific drilling and onshore seismic campaign targeting the EPGFZ
would provide invaluable information to help characterize the fault zone and increase our
understanding of the seismic risk it proposes. This is proposed in the Haiti-DRILL project (Ellouz-
Zimmermann and Pubellier, 2015), which has been submitted to the IODP/ICDP program and is
currently being evaluated.

288
List of figures

List of figures
Figure 2 - 1: Modified from Leroy (1995) and (Mauffret and Leroy, 1999). GPS velocities are indicated
by green arrows (for references see text). Red lines are locations of cross sections in Figure 2 - 2. Drill
rig symbols are locations of DSDP and ODP drill sites. Abbreviations; MAT = Middle America Trench;
MFZ = Motagua Fault Zone; CCT = Central Chortis terrane; GFS = Guayape fault system; ECT = East
Chortis terrane, SIU = Siuna terrane; SIRB; Swan Island Restraining Bend; SIFZ = Swan Island Fault Zone;
MCSC = Mid-Cayman Spreading Center; WFZ = Walton Fault Zone; HB = Hendrix pull-apart basin; OFZ
= Oriente Fault Zone; SDB = Santiago Deformed Belt; EPGFZ = Enriquillo-Plantain Garden Fault Zone;
SFZ = Septentrional Fault Zone; NHDB = North Hispaniola Deformed Belt; PRT = Puerto Rico Trench; MP
= Mona Passage; MT = Muertos Trough; LAT = Lesser Antilles Trench; SCDB – South Caribbean Deformed
Belt......................................................................................................................................................... 15
Figure 2 - 2: Geometry of the northern and eastern Caribbean subduction systems from Symithe et al.
(2015). Black line; plate interface between the North American and Caribbean plates. Dashed black
line; underplating at the Muertos Trough. Grey lines represent bathymetry at significant vertical
exaggeration compared to the earthquake depths, blue line is sea level. For locations and abbreviations
see Figure 2 - 1. ..................................................................................................................................... 16
Figure 2 - 3. A and B; in-situ model after Meschede and Frisch (1998). C and D; subduction polarity
reversal models after Pindell et al. (2006) (C) and Kerr et al. (2003) (D). E and F; continuous west-
dipping subduction after Mann et al. (2007a). G and H; continuous west-dipping subduction after
Pindell et al. (2012)................................................................................................................................ 18
Figure 2 - 4: From Mauffret and Leroy (1997). Schematic W – E cross section through the Caribbean
domain. See Figure 2 - 5 for location. ................................................................................................... 20
Figure 2 - 5: Crustal thickness in the Caribbean realm, modified from Leroy (1995). Data from Leroy
(1995) and Mauffret and Leroy (1997, 1999). Location of cross section of Figure 2 - 4 in light grey. .. 21
Figure 2 - 6: Compilation of ages and settings for intrusive and extrusive rocks, modified from
Boschman et al. (2014). For <65Ma only island-arc related or MORB units shown. Dredge samples from
Mauffret et al. (2001). HP mélange formation from Boschman et al. (2014). Intrusives are from Kerr et
al. (2003). Other references in text. ...................................................................................................... 22
Figure 2 - 7: Major onshore and offshore occurrences of N-MORB and CLIP or plume-influenced crust
in the offshore and northern Caribbean realm (Leroy, 1995; Mauffret and Leroy, 1997; Driscoll and
Diebold, 1999; Kerr et al., 2003; and references in text). Background image same as Figure 2 - 1. .... 23
Figure 2 - 8: Major onshore and offshore occurrences of IAT and CA- series crust in the northern
Caribbean. For references see text. Background image same as Figure 2 - 1. ..................................... 24

289
List of figures

Figure 2 - 9: Palinspastic reconstruction of the Caribbean realm at 125, 100, 84 and 71 Ma, from (Pindell
and Kennan, 2009). Blue lines are location of cross sections in Figure 2 - 10a to d. ............................ 25
Figure 2 - 10: Paleotectonic cross sections through the Caribbean, from (García-Casco et al., 2008).
Location of cross sections in Figure 2 - 9............................................................................................... 27
Figure 2 - 11: Slabs in the northeastern Caribbean realm derived from seismic tomography, from van
Benthem et al. (2013). ........................................................................................................................... 28

Figure 3 - 1: Morphotectonic zones in Hispaniola. Faults from Mann et al. (1991b), Escuder-Viruete et
al. (2002), Hernaiz Huerta et al. (2007), and Leroy et al. (2015). Onshore relief from AsterDEM, offshore
shaded relief from GEBCO. Locations of cities/towns from Open Street Maps (OSM). Abbreviations;
NHF = North Hispaniola Fault, CFZ = Camú Fault Zone; RGFZ = Rio Grande FZ; SFZ = Septentrional FZ;
OFZ = Oriente FZ; HT = Hatillo Thrust; HFZ = Hispaniola FZ; HVFZ = Hato Viejo FZ; BGFZ = Bonao –
Guácara FZ; LMSZ = La Meseta Shear Zone; SJRFZ = San José – Restauración FZ; LPSJFZ = Los Pozos –
San Juan FZ; MNFZ = Montagnes Noires – northern Sierra de Neiba FZ; MFZ = Chaîne des Matheux –
southern Sierra de Neiba FZ; TBF = Trois Baies Fault; EPGFZ = Enriquillo – Plantain Garden FZ........... 52
Figure 3 - 2: Stratigraphy of the main igneous units in Hispaniola. The morphotectonic zones are found
in Figure 3 - 1. HP-LT metamorphic paths after Escuder-Viruete et al. (2011b; c). Solid circles are
constrained P-T-t conditions, open circles are estimated P-T-t conditions by these authors.
Abbreviations for the metamorphic units; PBM = Punta Balandra Mélange, PB = Punta Balandra
sediments, SBS = Santa Bárbara Schists, RM = Rincón Marbles. ........................................................... 54
Figure 3 - 3: Geographic distribution of the main igneous units in Hispaniola. Faults are the same as in
Figure 3 - 1. For references see text. ..................................................................................................... 56
Figure 3 - 4: Evolution of geodynamic setting for the Cretaceous igneous units in Hispaniola. Modified
from Escuder-Viruete et al. (2006a, 2007b, 2011a). ............................................................................. 58
Figure 3 - 5: Geodynamic setting, timing, and P-T paths for metamorphic units in the Samaná and Rio
San Juan complexes, Cordillera Septentrional, Dominican Republic. Modified from Escuder-Viruete et
al. (2011c). ............................................................................................................................................. 62
Figure 3 - 6: Tectonostratigraphic chart of western Hispaniola, from roughly southwest (left) to
northeast (right). Stratigraphy grouped per morphotectonic terrane as indicated in Figure 3 - 1.
Abbreviations; EPGFZ = Enriquillo Plantain Garden Fault Zone; SP = Southern Peninsula; CC = Cordillera
Central; HFTB = Haitian fold-and-thrust belt; SFZ = Septentrional Fault Zone; CLIP = Caribbean Large
Igneous Province. Other fault zone abbreviations are listed in the legend of Figure 3 - 1. ................... 66
Figure 3 - 7: Simplified geological map of Hispaniola. Geological map of Haiti modified from Boisson
and Pubellier (1987, 1988), Boisson et al. (1989), and Bien-Aime Momplaisir et al. (1988). Geological

290
List of figures

map of the Dominican Republic modified after Hernaiz Huerta et al. (2007). Faults same as Figure 3 -
1. ............................................................................................................................................................ 69

Figure 4 - 1: Morphological map of the Southern Peninsula of Haiti with the EPGFZ. ......................... 75
Figure 4 - 2a: Palinspastic reconstruction of the northern Caribbean at 71 Ma. Black; Reconstruction by
(Pindell and Kennan (2009). Red; proposed palinspastic position of the Southern Peninsula. Figure 4 -
2b: Same as figure a, but at 56 Ma. Abbreviations: FA = Farallon; CA = Caribbean; NA = North American;
SA = South American; UNR = Upper Nicaragua Rise; LNR = Lower Nicaragua Rise; JAM = Jamaica; YUC
= Yucatan Peninsula; GOM = Golf of Mexico; SHBR = Southern Haiti - Beata Ridge; HFTB = Future Haitian
Fold-and-thrust Belt; CAY = Cayman Trough; HIS = Hispaniola; PR = Puerto Rico. ............................. 145
Figure 4 - 3: Orthogonal bedding perpendicular joints on the Southern Peninsula. Color of stereonet
indicates age of the lithology. Legend of the geological map same as in Figure 4 - 6. ....................... 146
Figure 4 - 4: Line-length restoration of cross sections on the Southern Peninsula. A; Overview map with
cross sections. B; Massif de la Selle cross section A. C; l’Asile Basin cross section D. D; western Massif
de la Hotte cross section B. ................................................................................................................. 149
Figure 4 - 5: Length vs displacement from global strike-slip faults, modified from Kim and Sanderson
(2005). Red boxes are length/displacement data for EPGFZ (500 and 1200 km length) and Rivière
Momance – Rivière Froide and Rivière de Cavaillon segments as discussed in the text. Horizontal lines
are displacement values for the associated segment lengths using the trend line of global strike-slip
faults (brown line). Vertical lines are segment length values for associated displacements. ............ 152
Figure 4 - 6: Geological map (top) and topographic map (bottom) of the Southern Peninsula. ........ 155
Figure 4 - 7: Rivière Momance – Froide EPGFZ segment displacement evolution. Abbreviations; EPGFZ
= Enriquillo-Plantain Garden Fault Zone, M = Rivière Momance, F = Rivière Froide, PAP = Port-au-Prince.
............................................................................................................................................................. 157
Figure 4 - 8: Rivière de Cavaillon EPGFZ segment displacement evolution. Abbreviations; EPGFZ =
Enriquillo-Plantain Garden Fault Zone, C = Rivière de Cavaillon, RdS = Ravine du Sud. ...................... 159

Figure 5 - 1: Structural map of the Chaîne des Matheux – Montagnes du Trou d’Eau region. Black box
is outline of the small scale geological map (Figure 5 - 10). Elevation from AsterDEM. Faults and folds
modified from Boisson and Pubellier (1987), Mann et al. (1995) and Hernaiz Huerta et al. (2007).
Abbreviations; MNFZ = Montagnes Noires Fault Zone, MFZ = Matheux Fault Zone .......................... 163
Figure 5 - 2: A: Observed pre-Eocene series in the western Chaîne des Matheux, Pierre Payen anticline,
based on the work by Vila et al. (1988). B: Lithologies in the correct stratigraphic order. ................ 165
Figure 5 - 3: Synthetic stratigraphic columns for the northern and southern flank of the eastern Chaîne
des Matheux, north of the Cul-de-Sac plain........................................................................................ 167

291
List of figures

Figure 5 - 4: Haitian fold-and-thrust belt. Top image: Diachronous clastic infill recorded in basins on the
flanks of the thrust sheets. Vertical segments represent error bars. Wavy lines represent unconformity
following basin infill completion. Grey area is the duration of terrigenous sedimentation in the basins.
References in, and modified from, Pubellier et al. (2000). Bottom image: Compilation of cross sections
through the Haitian fold-and-thrust belt, modified from Pubellier et al. (2000). Abbreviations; LPSJFZ =
Los Pozos – San Juan Fault Zone, SMPFZ = Southern Mt. Paincrox Fault Zone, MNFZ = Montagnes Noires
Fault Zone, MFZ = Matheux Fault Zone. .............................................................................................. 169
Figure 5 - 5: Stratigraphy of the northern and southern flank of the Trou Caïman – Mirebalais cross
section. ................................................................................................................................................ 171
Figure 5 - 6: Stratigraphy of the northern and southern flank of the Titanyen - Dubon cross section.
............................................................................................................................................................. 173
Figure 5 - 7: Outcrop-scale style of deformation. A; u. Miocene calcareous sandstones, northern flank,
north of Gimballe. B: sigmoidal fractures in Burdigalian chalky limestones with top-to-SW motion,
northern flank, south of Triano. C; Syncline in u. Miocene (Tortonian?) marls and limestones, southern
flank, Ravine Sèche. D; SW-verging thrust in m. Miocene (Serravallian?) calcareous conglomerates,
southern flank, Ravine Sèche. E; folded Tortonian marls, limestones and basalts, southern flank, north
of Saint Phard. ..................................................................................................................................... 175
Figure 5 - 8: Outcrop-scale style of deformation. A; northeast verging box-fold in u. Eocene – Aquitanian
cherty limestones, northern flank, between Triano and Saintonge. B; NE-verging bedding-parallel and
upwards-ramping thrusts in u. Eocene – Oligocene cherty limestones overlain by MTD breccia,
northern flank, between Triano and Saintonge. C; SE-verging thrust with footwall drag fold, northern
flank, southwest of Hatte Palicarpe. D; Intra-formational slump in u. Eocene to Aquitanian well-
stratified cherty limestones, northern flank, southwest of Hatte Palicarpe....................................... 177
Figure 5 - 9: A; Thrust faults in l. – m. Eocene biostrome complex, northern flank, around Bois Dorme.
B; l. – m. Eocene biostrome complex cut by an upwards-ramping SE-verging thrust and two sets of
strike-slip faults (NNE – SSW and WNW – ESE trending), southern flank, west of Lerebours. ........... 178
Figure 5 - 10: Geological map of the eastern part of the Chaîne des Matheux. Abbreviations; BT =
Beauget Thrust, RST = Ravine Sèche Thrust, TT = Titanyen Thrust. ..................................................... 179
Figure 5 - 11: A: Titanyen – Dubon cross section. B; Trou Cayman – Mirebalais cross section. For location
see Figure 5 - 10. Both cross-sections are on the same, 1to1 scale. Abbreviations; BT = Beauget Thrust,
RST = Ravine Sèche Thrust, TT = Titanyen Thrust. ............................................................................... 180
Figure 5 - 12: Folded structures southwest of Hatte Palicarpe. A; Photo-interpretation. B; schematic
cross section through structure. C; Fold-axes and thrusts with motion. ............................................ 183
Figure 5 - 13: Joint sets and shear fractures from the structure southwest of Hatte Palicarpe (see Figure
5 - 12 for locations). A; Joints and shear fractures back-tilted to horizontal S0. B; Joint set back-tilted to

292
List of figures

horizontal S0. E – W to ENE – WSW trending joints (set 1a) are occasionally cemented by calcite veins.
C; Bedding parallel shear fractures in original orientation (left) and back-tilted to horizontal S0 (right)
with top-to-SE motion. D; Joint sets back-tilted to horizontal S0. ...................................................... 185
Figure 5 - 14: Joint sets in Aquitanian (?) cherty limestones, close to Ca Michaud. A; original joint
dataset. B; joint dataset back-tilted to horizontal S0. Set 1a most dominant, set 1b mainly abutting.
Note that set 1a joints trend parallel to the face of the outcrop and are thus not shown. Red: Joint set
1a. Yellow: Joint set 1b. Blue: Reactivated bedding and thrusts. Calcite mineralizations mainly observed
on set 1a joints. ................................................................................................................................... 187
Figure 5 - 15: Compilation chart of the Chaîne des Matheux displaying relative sea level, stratigraphy,
deformation events, uplift and subsidence, and the associated stress regime. See text for discussion.
............................................................................................................................................................. 189

Figure 6 - 1: A/B; fault zone architecture, modified from Faulkner et al. (2010). A: Simple fault zone
structure, with high-strain core surrounded by fractured damage zone. B: Multiple fault core model,
with many strands of highs-strain material enclosing lenses of fractured host rock. C: Fault
displacement versus fault damage zone width, modified from, and references within, Savage and
Brodsky (2011). Thick black line is best fitting curve. Thin black lines within graph delineate data. Light
grey column are 80% confidence limits for changing point from linear to decreasing scaling relationship
between displacement and width. EPGFZ range from section 4.6. .................................................... 194
Figure 6 - 2: Damage zone microfracture density vs distance from fault core. A: Data from three faults
with different displacement, all within low porosity granodiorite, Coastal Cordillera, Northern Chili.
Solid lines are best fitting to the data. Note that maximum fracture density is similar for all three faults,
irrespective of displacement. Modified from, and references within, Mitchell and Faulkner (2009). B:
Data from four faults; two with low porosity (9%) and two with high porosity (25%) host rock. Solid
lines are best fitting to the data. Modified from, and references within, Faulkner et al. (2010). ...... 196
Figure 6 - 3: Fracture propagation modes. Mode-I; opening. Mode-II; sliding. Mode-III; tearing.
Modified from Fossen (2010). ............................................................................................................. 200
Figure 6 - 4: Mohr diagram and failure envelope. A: Mohr diagram with Coulomb-Griffith failure
envelope and modes of failure. B: Tensile failure under elevated fluid pressure Pf. C: Shear failure under
elevated fluid pressure Pf. Modified from Fossen (2010). .................................................................. 201
Figure 6 - 5: Mohr diagram with failure in anisotropic layered rocks. A: Layer-parallel extension in a
sequence of competent and incompetent lithologies. B: Layer-parallel shortening in a sequence of
competent and incompetent lithologies. Modified from (Bons et al., 2012). .................................... 203
Figure 6 - 6: Frictional sliding along pre-existing plane of weakness. A: insufficient effective differential
stress for failure; stable condition. B; sufficient effective differential stress for failure on pre-existing

293
List of figures

fracture. Note that if the angle between σ1 and the plane of the pre-existing fracture becomes large
enough, the Mohr-circle will touch the Coulomb failure envelope before reactivation occurs, thereby
promoting the creation of new fractures. Modified from Twiss and Moores (2007) and Fossen (2010).
............................................................................................................................................................. 203
Figure 6 - 7: Scheme linking vein opening mechanism (antitaxial, syntaxial or stretching) and crystal
morphology to the growth plane and number of crack-seal events. Crystal types; b) blocky, c) elongate
blocky, d) stretched crystals and e) fibrous. From Bons et al. (2012). ................................................ 206
Figure 6 - 8. Chronology of deformation and associated veins in the Chaîne des Matheux. Left;
schematic representation of deformation and veins within a multilayered volume of rock. Grey are
marly / less competent intervals, white are limestone / competent intervals. Center left; dominant
orientation of structures active during this time. Center right; schematic Mohr-diagram with Coulomb-
Griffith failure criterion for the rock. Right; fracture and vein development and associated stresses.
............................................................................................................................................................. 210
Figure 6 - 9: Sample locations with vein types and associated figures. A, B, C and D are fluid inclusion
samples................................................................................................................................................ 211
Figure 6 - 10: Calcite mineralization along bedding perpendicular orthogonal joints (set 1a) offset by
calcite-filled shear fracture with normal motion (set 2). For location see Figure 6 - 9....................... 212
Figure 6 - 11: Reactivated bedding vein (set 3a). For location see Figure 6 - 9. ................................. 213
Figure 6 - 12: Calcite vein along strike-slip fault. For location see Figure 6 - 9. .................................. 214
Figure 6 - 13: Calcite vein related to karst processes. For location see Figure 6 - 9. .......................... 215
Figure 6 - 14: Stable oxygen and carbon isotope data for veins and host rocks. A: Host rock δ18O vs δ13C.
B: Vein sets δ18O vs δ13C. C: δ13C vein vs host rock. D: δ18O vein vs host rock. E: Age of the host rock
vein δ18O vs δ13C. Vein stable isotope data labelled A, B, C and D correspond to the fluid inclusion
examples A, B, C and D in Figure 6 - 15 and section 6.4.3. ................................................................. 216
Figure 6 - 15: Fluid inclusion microthermometry results for the Chaîne des Matheux. A: Examples of
fluid inclusions from calcite veins. B: Distribution of filling factor (F) of the inclusion. C: Distribution of
homogenization temperature (Th). D: Parent fluid composition calculated using the homogenization
temperature of the inclusions, δ18O of the corresponding vein calcite, and oxygen isotope fractionation
curves calculated after Kim and O’Neil (1997). ................................................................................... 218
Figure 6 - 16: Rare earth elements + Yttrium (REY) data for samples from the Chaîne des Matheux,
PAAS normalized after Taylor and McLennan (1985). A: REY for host rocks. B: REY for veins. C: REY for
veins and host rocks. ........................................................................................................................... 220
Figure 6 - 17: Schematic diagram showing the timing of the veins and their characteristics. Left;
schematic representation of deformation and veins within a multilayered volume of rock. Grey are
marly / less competent intervals, white are limestone / competent intervals. T = homogenization

294
List of figures

temperature of fluid inclusions, question mark indicates estimated temperatures. Center left;
dominant orientation of structures active during this time. Center; schematic Mohr-diagram with
Coulomb-Griffith failure criterion for the rock. Center right; fracture and vein development and
associated stresses. Right; δ18O vs δ13C plot of veins. ......................................................................... 222
Figure 6 - 18: Schematic evolution of paleo-fluid circulation in the Chaîne des Matheux. All schematic
sections are oriented NE – SW, apart from section D which is oriented NW – SE and thus along-strike
of the decollement. ............................................................................................................................. 227
Figure 6 - 19: Chronology of deformation and associated veins on the Southern Peninsula. Left;
schematic representation of deformation and veins within a multilayered volume of rock. Grey are
marly / less competent intervals, white are limestone / competent intervals. Center left; dominant
orientation of structures active during this time. Center right; schematic Mohr-diagram with Coulomb-
Griffith failure criterion for the rock. Right; fracture and vein development and associated stresses.
............................................................................................................................................................. 229
Figure 6 - 20: Sample locations on the Southern Peninsula with vein types and associated figures. A, B,
C and D are fluid inclusion samples. .................................................................................................... 230
Figure 6 - 21: Examples of samples containing both silica and calcite veins. Host rocks are all Cretaceous
slope limestones. Abbreviations; HR = host rock, cc = calcite, si = silica............................................. 231
Figure 6 - 22: Example of thrust-related calcite-silica veins in Upper Cretaceous limestones. For location
see Figure 6 - 20. ................................................................................................................................. 232
Figure 6 - 23: Calcite mineralization along bedding perpendicular orthogonal joints. Host rock is a lower
Eocene platform limestone. For location see Figure 6 - 20. ............................................................... 233
Figure 6 - 24: Calcite vein with right-lateral shear motion. Host rock is lower Paleocene volcaniclastics.
For location see Figure 6 - 20. ............................................................................................................. 234
Figure 6 - 25: Calcite vein from the core of the EPGFZ, south of Port-au-Prince. Host rock is lower to
middle Miocene limestone. For location see Figure 6 - 20. ................................................................ 235
Figure 6 - 26: Calcite veins along a thrust fault. Host rock is Upper Cretaceous limestone. For location
see Figure 6 - 20. ................................................................................................................................. 235
Figure 6 - 27: Strike-slip related vein west of Miragoane Basin. Host rock is lower Eocene platform
limestone. For location see Figure 6 - 20. ........................................................................................... 236
Figure 6 - 28: Calcite vein along normal faults, west of Miragoane Basin. Host rock is middle to upper
Eocene limestone. For location see Figure 6 - 20. .............................................................................. 237
Figure 6 - 29: Stable oxygen and carbon isotope data for veins and host rocks. A: Host rock δ18O vs δ13C.
B: Vein sets δ18O vs δ13C. C: δ13C vein vs host rock. D: δ18O vein vs host rock. E: Age of the host rock
vein δ18O vs δ13C. Vein stable isotope data labelled A, B, C and D correspond to the fluid inclusion
examples A, B, C and D in Figure 6 - 32 and section 6.6.3. Samples labelled E, F, G and H; see text. 239

295
List of figures

Figure 6 - 30: Vein deformation and cathodoluminescence intensity. A; Chert layer from upper Eocene
limestones west of Miragoane Basin. B; Core of EPGFZ, same location as Figure 6 - 25.................... 240
Figure 6 - 31: Vein deformation intensity and cathodoluminescence. A: δ18O vs δ13C plot with
deformation intensity of the veins. B: Plot of vein deformation intensity vs cathodoluminescence
intensity overlain with the δ13C values of vein calcite. C: Vein cathodoluminescence versus Mn and Fe
concentrations. D: Same as C but overlain with vein deformation intensity. For C and D, background
image from Boggs and Krinsley (2006) shows divalent Fe and Mn, whereas the overlain samples are
plotted with their respective ppm concentrations with unknown valence. Red box indicates detection
limit for samples, with those below the limit plotted in the lower left corner of the diagram. ......... 241
Figure 6 - 32: Fluid inclusion microthermometry results from the Southern Peninsula. A: Examples of
fluid inclusions from calcite veins. B: Distribution of filling factor (F) of the inclusion. C: Distribution of
homogenization temperature (Th). D: Parent fluid composition calculated using the homogenization
temperature of the inclusions, δ18O of the corresponding vein calcite, and oxygen isotope fractionation
curves calculated after Kim and O’Neil (1997). ................................................................................... 243
Figure 6 - 33: Rare earth elements + Yttrium (REY) data for samples from the Southern Peninsula, PAAS
normalized after Taylor and McLennan (1985). A: REY for host rock sediments. B: REY for basalts.. 245
Figure 6 - 34: Rare earth elements + Yttrium (REY) data for samples from the Southern Peninsula, PAAS
normalized after Taylor and McLennan (1985). Host rock and veins with the same line pattern are from
the same sample. A: REY for set 2 calcite veins. B: REY for set 3 calcite veins. C; REY for set 6b calcite
veins. D; REY for silica veins and chert layers...................................................................................... 246
Figure 6 - 35: Schematic diagram showing the timing of the veins and their characteristics. Left;
schematic representation of deformation and veins within a multilayered volume of rock. Grey are
marly / less competent intervals, white are limestone / competent intervals. T = homogenization
temperature of fluid inclusions, question mark indicates estimated temperatures. Center left;
dominant orientation of structures active during this time. Center; schematic Mohr-diagram with
Coulomb-Griffith failure criterion for the rock. Center right; fracture and vein development and
associated stresses. Right; δ18O vs δ13C plot of veins. ......................................................................... 248
Figure 6 - 36: Schematic evolution of paleo-fluid circulation in the Southern Peninsula. .................. 252

Figure 7 - 1: Shortening (%) calculated from cross sections in this study (A, B, C, K, and J) and published
cross sections (Appendix 2). ................................................................................................................ 257
Figure 7 - 2: Inferred amounts of displacements in a direction parallel to the strike of the EPGFZ, for
the Massif de la Selle and Sierra de Bahoruco. ΔL is the amount of Miocene to present-day shortening
perpendicular to the structural trend in the corresponding region. GPS vectors in blue from Calais et
al. (2016).............................................................................................................................................. 261

296
List of figures

Figure 7 - 3: Average geological shortening rates in southwestern Hispaniola, based on the shortening
in Figure 7 - 1 and temporal constraints in Table 7 - 2........................................................................ 263
Figure 7 - 4: Present-day strain rates in southwestern Hispaniola from GPS velocity modeling by
Symithe and Calais (2016). .................................................................................................................. 266
Figure 7 - 5: Geological compilation map of the Southern Peninsula, merging the three small-scale
geological maps of section 4.2 with the large-scale geological maps of the BME. ............................. 269
Figure 7 - 6: Paleogeographic reconstruction maps of the Southern Peninsula of Haiti for the Miocene
to present-day. .................................................................................................................................... 274
Figure 7 - 7: Moho depth in Haiti, based on the work by Corbeau et al. (2017). Morphotectonic zones
and faults are taken from section 3.1, Figure 3 – 1. Abbreviations; HFTB = Haitian Fold-and-Thrust Belt,
SP = Southern Peninsula. ..................................................................................................................... 277
Figure 7 - 8: Simplified crustal stratigraphy of the Caribbean Large Igneous Province, based on the work
by Mauffret and Leroy (1997), Diebold et al. (1999), Révillon et al. (2000), Mauffret et al. (2001). . 277
Figure 7 - 9: Geological and morphological map of the composite crustal-scale cross section. Moho
depth values are from Corbeau et al. (2017), springs (Fortin and Sources Puantes) are from (Battani et
al. (2016). Section A is Massif de la Selle section (section 4.2). Sections C and D are from the Trou
Caiman – Mirebalais cross section (section 5.5). Section B is the projected Cul-de-Sac part of the
Titanyen – Dubon cross section (section 5.5). .................................................................................... 279
Figure 7 - 10: Crustal-scale cross section through the Massif de la Selle and Chaîne des Matheux. See
Figure 7 - 9 for location of subsections A, B, C, and D......................................................................... 282

Table 7 - 1: Shortening in southwestern Hispaniola. Abbreviations; VdB-83 = Van den Berghe (1983a),
HH-07 = Hernaiz Huerta et al. (2007), HH-02 = Hernaiz Huerta and Pérez-Estaún (2002), Pub-00 =
Pubellier et al. (2000), Corb-16 = Corbeau et al. (2016). ..................................................................... 258
Table 7 - 2: Shortening rates in southwestern Hispaniola. Abbreviations are the same as in Table 7 - 1.
Additional abbreviations; B-88 = Bourgueil et al. (1988), Pub-91 = Pubellier et al. (1991)................. 265
Table 7 - 3: Shortening and strike-slip displacement on the Southern Peninsula. ............................. 272
Table 7 - 4: Displacement along the EPGFZ and associated strike-slip faults. .................................... 272

297
List of figures

298
References

References
Abbott, R.N., Draper, G., and Broman, B.N., 2006: P-T Path for Ultrahigh-Pressure Garnet Ultramafic
Rocks of the Cuaba Gneiss, Rio San Juan Complex, Dominican Republic. International Geology
Review, 48, 778–790.
Abbott, R.N., Broman, B.N., and Draper, G., 2007: UHP Magma Paragenesis Revisited, Olivine
Clinopyroxenite and Garnet-Bearing Ultramafic Rocks from the Cuaba Gneiss , Rio San Juan
Complex, Dominican Republic. International Geology Review, 49, 572–586.
Abbott, R.N., Bandy, B.R., and Rajkumar, A., 2013: Cenozoic burial metamorphism in eastern Jamaica.
Caribbean Journal of Earth Science, 46, 13–30.
Agosta, F., and Kirschner, D.L., 2003: Fluid conduits in carbonate-hosted seismogenic normal faults of
central Italy. Journal of Geophysical Research, 108 (B4), 1–13.
Agosta, F., and Aydin, A., 2006: Architecture and deformation mechanism of a basin-bounding normal
fault in Mesozoic platform carbonates, central Italy. Journal of Structural Geology, 28, 1445–
1467.
Aharonov, E., Tenthorey, E., and Scholz, H., 1998: Precipitation sealing and diagenesis: 2. Theoretical
analysis. Journal of Geophysical Research, 103 (B10), 969–981.
Ali, S.T., Freed, A.M., Calais, E., Manaker, D.M., and McCann, W.R., 2008: Coulomb stress evolution in
Northeastern Caribbean over the past 250 years due to coseismic, postseismic and interseismic
deformation. Geophysical Journal International, 174 (3), 904–918.
Alva-Valdivia, L., Goguitchaichvili, a., Cobiella-Reguera, J., Urrutia-Fucugauchi, J., Fundora-Granda, M.,
Grajales-Nishimura, J.M., and Rosales, C., 2001: Palaeomagnetism of the Guaniguanico Cordillera,
western Cuba: a pilot study. Cretaceous Research, 22, 705–718.
Amilcar, H., 1997: Etude géologique du Massif de La Hotte occidental, Presqu’île du Sud d’Haïti
(Grandes Antilles). Implications géodynamiques: 188 p.
Amrouch, K., Lacombe, O., Bellahsen, N., Daniel, J.-M., and Callot, J.-P., 2010: Stress and strain
patterns, kinematics and deformation mechanisms in a basement-cored anticline: Sheep
Mountain Anticline, Wyoming. Tectonics, 29 (1), 1–27.
Anders, M.H., and Wiltschko, D. V., 1994: Microfracturing, paleostress and the growth of faults. Journal
of Structural Geology, 16 (6), 795–815.
Anderson, E.M., 1905: The dynamics of faulting. Transactions of the Edinburgh Geological Society, 8,
387–402.
Andreieff, P., 1981: Valeur stratigraphique des grands foraminiferes de la region Caraibe a l’Oligo-
Miocene: Un premier essai de mise au point. Princip. Résult. Scient. Techn. SGN, Ed. BRGM,, 56–
57.
Andreieff, P., 1983: Extension stratigraphique des grands foraminiferes Neogenes de la region Caraibe :
Paraspiroclypeus Chawneri (Palmer) et operculinoides Cojimarensis (Palmer). Bulletin de la
Societe Géologique de France, 25 (7), 885–888.
Andreieff, P., and Desreumaux, C., 1984: Le passage Maestrichtien - Danien dans le sud d’Haïti
(Grandes Antilles). 109 ème congr. soc. sav. Dijon, Section Sciences 1,, 311–322.
Andreieff, P., 1985: Stratigraphic range of Caribbean larger foraminifera from Oligocene to Pliocene:
State of knowledge. Symp. geodyn. Caraïbes, Paris, Ed. Technip,, 99–100.

299
References

Andreieff, P., Bouysse, P., and Westercamp, D., 1987: Geologie de l’arc insulaire de Petites Antilles et
evolution geodynamique de l’est Caraïbe: 359 p.
Andreu, E., Torró, L., Proenza, J. a., Domenech, C., García-Casco, A., Villanova de Benavent, C., Chavez,
C., Espaillat, J., and Lewis, J.F., 2015: Weathering profile of the Cerro de Maimón VMS deposit
(Dominican Republic): textures, mineralogy, gossan evolution and mobility of gold and silver. Ore
Geology Reviews, 65 (November), 165–179.
Angelier, J., 1984: Tectonic analysis of fault slip data sets. Journal of Geophysical Research, 89 (B7),
5835–5848.
Antonellini, M.A., Aydin, A., and Pollard, D.D., 1994: Microstructure of deformation bands in porous
sandstones at Arches National Park, Utah. Journal of Structural Geology, 16 (7), 941–959.
Ayala, A., 1959: Estudios de algunos microfosiles planctonicos de las calizas del Cretacico superior de
la Republica de Haiti. Paleontologia Mexicana, 4, 42.
Bahat, D., and Grossmann, H.N.F., 1988: Regional jointing and paleostresses in Eocene chalks around
Beer-Sheva, Israel. Israel Journal of Earth Sciences, 37, 1–11.
Bai, T., Maerten, L., Gross, M.R., and Aydin, A., 2002: Orthogonal cross joints: Do they imply a regional
stress rotation? Journal of Structural Geology, 24 (1), 77–88.
Baker, A., Ito, E., Smart, P.L., and McEwan, R.F., 1997: Elevated and variable values of 13C in
speleothems in a British cave system. Chemical Geology, 136 (3–4), 263–270.
Bakun, W.H., Flores, C.H., and ten Brink, U.S., 2012: Significant Earthquakes on the Enriquillo Fault
System, Hispaniola, 1500-2010: Implications for Seismic Hazard. Bulletin of the Seismological
Society of America, 102 (1), 18–30.
Balsamo, F., Storti, F., Salvini, F., Silva, A.T., and Lima, C.C., 2010: Structural and petrophysical evolution
of extensional fault zones in low-porosity, poorly lithified sandstones of the Barreiras Formation,
NE Brazil. Journal of Structural Geology, 32 (11), 1806–1826.
Balsamo, F., Clemenzi, L., Storti, F., Mozafari, M., Solum, J., Swennen, R., Taberner, C., and
Tueckmantel, C., 2016: Anatomy and paleofluid evolution of laterally restricted extensional fault
zones in the Jabal Qusaybah anticline, Salakh arch, Oman. Geological Society of America Bulletin,
128 (5/6), 957–972.
Banks, V.J., and Jones, P.F., 2012: Hydrogeological significance of secondary terrestrial carbonate
deposition in karst environments, in Kazemi, G.A. ed., Hydrogeology - A Global Perspective,
IntechOpen, 43–78.
Bar-Matthews, M., Ayalon, a., Matthews, a., Sass, E., and Halicz, L., 1996: Carbon and oxygen isotope
study of the active water-carbonate system in a karstic Mediterranean cave.pdf. Geochimica et
Cosmochimica Acta, 60 (2), 337–347.
Bar-Matthews, M., Ayalon, A., and Kaufman, A., 1997: Late Quaternary Paleoclimate in the Eastern
Mediterranean Region from Stable Isotope Analysis of Speleothems at Soreq Cave, Israel.
Quaternary Research, 168 (47), 155–168.
Bar-Matthews, M., Ayalon, A., Gilmour, M., Matthews, A., and Hawkesworth, C.J., 2003: Sea - land
oxygen isotopic relationships from planktonic foraminifera and speleothems in the Eastern
Mediterranean region and their implication for paleorainfall during interglacial intervals.
Geochimica et Cosmochimica Acta, 67 (17), 3181–3199.
Barnett, J.A.M., Mortimer, J., Rippon, J.H., Walsh, J.J., and Watterson, J., 1987: Displacement Geometry
in the Volume Containing a Single Normal Fault. American Association of Petroleum Geologists

300
References

Bulletin, 71 (8), 925–937.


Bastesen, E., and Braathen, A., 2010: Extensional faults in fine grained carbonates - analysis of fault
core lithology and thickness - displacement relationships. Journal of Structural Geology, 32 (11),
1609–1628.
Bates, R.L., and Jackson, J.A., 1987: Glossary of geology. Alexandria, Virginia, American Geological
Institute, 788.
Battani, A., Burnard, P.G., Ellouz-Zimmermann, N., Blard, P.-H., Pillot, D., Mercier de Lepinay, B.,
Ruffine, L., and Momplaisir, R., 2016: Fluid migration in the area of Haiti: Evidence for deeply
rooted active faults highlighted by noble gas measurements, in Dingue, 1.
Bau, M., Möller, P., and Dulski, P., 1997: Yttrium and lanthanides in eastern Mediterranean seawater
and their fractionation during redox-cycling. Marine Chemistry, 56 (1–2), 123–131.
Bau, M., and Dulski, P., 1999: Comparing yttrium and rare earths in hydrothermal fluids from the Mid-
Atlantic Ridge: Implications for Y and REE behaviour during near-vent mixing and for the Y/Ho
ratio of proterozoic seawater. Chemical Geology, 155 (1–2), 77–90.
Bau, M., Balan, S., Schmidt, K., and Koschinsky, A., 2010: Rare earth elements in mussel shells of the
Mytilidae family as tracers for hidden and fossil high-temperature hydrothermal systems. Earth
and Planetary Science Letters, 299 (3–4), 310–316.
Baumgartner, P.O., Florer, K., Bandini, A.N., Girault, F., and Cruz, D., 2008: Upper Triassic to Cretaceous
adiolaria from Nicaragua and Northern Costa Rica - The Mesquito Composite Oceanic Terran.
Ofioliti, 33 (1), 1–19.
Beaudoin, N., Bellahsen, N., Lacombe, O., and Emmanuel, L., 2011: Fracture-controlled
paleohydrogeology in a basement-cored, fault-related fold: Sheep Mountain Anticline, Wyoming,
United States. Geochemistry, Geophysics, Geosystems, 12 (6), 1–15.
Bellahsen, N., Fiore, P., and Pollard, D.D., 2006: The role of fractures in the structural interpretation of
Sheep Mountain Anticline, Wyoming. Journal of Structural Geology, 28 (5), 850–867.
Bellon, H., Vila, J.M., and Mercier de Lépinay, B., 1985: Chronologie 40K - 40Ar et affinites geochemique
des manifestations magmatiques au Cretace et au Paleogene dans l’ile d’Hispaniola. Symp.
geodyn. Caraïbes, Paris, Ed. Technip,, 329–340.
Benford, B., DeMets, C., and Calais, E., 2012a: GPS estimates of microplate motions, northern
Caribbean: evidence for a Hispaniola microplate and implications for earthquake hazard.
Geophysical Journal International, 191 (2), 481–490.
Benford, B., DeMets, C., Tikoff, B., Williams, P., Brown, L., and Wiggins-Grandison, M., 2012b: Seismic
hazard along the southern boundary of the Gônave microplate: block modelling of GPS velocities
from Jamaica and nearby islands, northern Caribbean. Geophysical Journal International, 190 (1),
59–74.
Benford, B., Tikoff, B., and DeMets, C., 2012c: Character of the Caribbean-Gônave-North America plate
boundaries in the upper mantle based on shear-wave splitting. Geophysical Research Letters, 39
(24), 1–6.
Benford, B., Tikoff, B., and DeMets, C., 2015: Fault interaction of reactivated faults within a restraining
bend: Neotectonic deformation of southwest Jamaica. Lithosphere, 7 (1), 21–39.
Bennett, R. a., Rodi, W., and Reilinger, R.E., 1996: Global Positioning System constraints on fault slip
rates in southern California and northern Baja, Mexico. Journal of Geophysical Research, 101
(B10), 21943.

301
References

van Benthem, S., Govers, R., Spakman, W., and Wortel, R., 2013: Tectonic evolution and mantle
structure of the Caribbean. Journal of Geophysical Research: Solid Earth, 118 (6), 3019–3036.
van Benthem, S., Govers, R., and Wortel, R., 2014: What drives microplate motion and deformation in
the northeastern Caribbean plate boundary region? Tectonics, 33, 1–24.
Bergbauer, S., and Pollard, D.D., 2004: A new conceptual fold-fracture model including prefolding
joints, based on the Emigrant Gap anticline, Wyoming. Bulletin of the Geological Society of
America, 116 (3–4), 294–307.
Van den Berghe, B., 1983a: Evolution sedimentaire et structurale depuis le Paleocene du secteur
“Massif de la Selle” (Haïti) - “Bahoruco” (République Dominicaine) au nord de la ride de Beata,
dans l’orogene nord Caraïbe (Hispaniola-Grandes Antilles): 305 p.
Van den Berghe, B., 1983b: Decrochement senestre sud-Haïtien : Analyses structurales dans le Massif
de la Selle. Ann. Soc. Géol. Nord, 103, 317–323.
Bermudez, P.J., 1949: Tertiary smaller foraminifera of the Dominican Republic. Cush. Lab. Foram. Res.,
Special publications, 25, 322.
Berthier, F., and Janjou, D., 1984: Evaluation des ressources geothermiques basse energie en
republique d’Haïti, recherche d’un project type - rapport geologique. Rapport B.R.G.M. 84 SGN
206 GTH,.
Bevan, T.G., and Hancock, P.L., 1986: A late Cenozoic regional mesofracture system in southern
England and northern France. Journal of the Geological Society, 143 (2), 355–362.
Bien-Aime Momplaisir, R., 1986: Contribution à l’étude géologique de la partie orientale du Massif de
la Hotte (Presqu’île du Sud d’Haïti); Synthèse structurale des marges de la Presqu’île à partir des
données sismiques: 261 p.
Bien-Aime Momplaisir, R., Amilcar, H., Murat Pierre, G., and Cenatus Amilcar, H., 1988: Les Cayes 1:250
000 geological map sheet.
Biju-Duval, B., Chartier, A., Mascle, A., and Montadert, L., 1980: Etude des series sedimentaires
Cretacees et Tertiaires de la Jamaique et d’Hispaniola, Grandes Antilles. Rapport interne Institute
Francais du Petrol,.
Biju-Duval, B., Bizon, G., Mascle, A., and Muller, C., 1982: Active margin processes: Field observations
in Southern Hispaniola. AAPG Memoir, 34, 325–344.
Bilham, R., 2010a: Lessons from the Haiti earthquake. Nature, 463 (February), 878–879.
Bilham, R., 2010b: Invisible faults under shaky ground. Nature Geoscience, 3 (11), 743–745.
Billi, A., Salvini, F., and Storti, F., 2003: The damage zone-fault core transition in carbonate rocks:
Implications for fault growth, structure and permeability. Journal of Structural Geology, 25 (11),
1779–1794.
Billi, A., 2010: Microtectonics of low-P low-T carbonate fault rocks. Journal of Structural Geology, 32
(9), 1392–1402.
Bird, D.E., Hall, S. a., Burke, K., Casey, J.F., and Sawyer, D.S., 2007: Early Central Atlantic Ocean seafloor
spreading history. Geosphere, 3 (5), 282–298.
Bizon, G., Bizon, J.J., Calmus, T., Muller, C., and Van den Berghe, B., 1985: Stratigraphie du Tertiaire du
sud d’Hispaniola (Grandes Antilles). Influence de la tectonique déchrochante sur la
paléogéographie et l’histoire sédimentaire. Symp. Geodyn. Caraïbes, Paris, Ed. Technip,, 371–380.

302
References

Blenkinsop, T.G., 2008: Relationships between faults, extension fractures and veins, and stress. Journal
of Structural Geology, 30, 622–632.
Bodnar, R.J., Reynolds, T.J., and Kuehn, C.A., 1985: Fluid inclusion systematics in epithermal systems.
Reviews in Economic Geology, 2, 73–97.
Bodnar, R.J., 1993: Revised equation and table for determining the freezing point depression of H2O-
NaCl solutions. Geochimica et Cosmochimica Acta, 57 (3), 683–684.
Boggs, S., and Krinsley, D., 2006: Application of cathodoluminescence imaging to the study of
sedimentary rocks. Cambridge University Press, 165.
Boisseau, M., 1987: Le flanc nord-est de la Cordilliere Dominicaine (Hispaniola, Grandes Antilles): Un
edifice de nappe Cretace polyphase: 200 p.
Boisson, D., and Vila, J.M., 1982: Premieres donnees stratigraphiques sur le flysch dans le Massif du
Nord d Haïti (Grandes Antilles). 9e R.A.S.T. Paris,.
Boisson, D., and Vila, J.M., 1984: Une coupe interpretative du versant sud du Massif du Nord d’Haiti
pres de la frontiere Dominicano - Haïtienne. 10 ème R.A.S.T., Bordeaux, Soc. Géol. Fr., Edit.,, 71.
Boisson, D., 1987: Etude geologique du Massif du Nord d’Haiti, Hispaniola - Grandes Antilles: Universite
Pierre et Marie Curie, 296 p.
Boisson, D., and Pubellier, M., 1987: Cap Haitian 1:250 000 geological map sheet.
Boisson, D., and Bien-Aime Momplaisir, R., 1987: Port-au-Prince 1:250 000 geological map sheet. , 1.
Boisson, D., and Pubellier, M., 1988: Mole Saint Nicolas 1:250 000 geological map sheet. , 1.
Boisson, D., Bien-Aime Momplaisir, R., Rigaud, J.G., D’Meza, S., Baptiste, J., Amilcar, H., Calmus, T., van
den Berghe, B., Dubreuilh, P., Pubellier, M., and Laclede-Carrere, J.-M., 1989: Carte geologique
de la Republique d’Haïti; scale 1:250.000.
Van den Bold, W.A., 1981: Distribution of Ostracoda in the Neogene of Central Haïti. Bulletin of
American Paleontology, 79 (312), 7–136.
Bolhar, R., Kamber, B.S., Moorbath, S., Fedo, C.M., and Whitehouse, M.J., 2004: Characterisation of
early Archaean chemical sediments by trace element signatures. Earth and Planetary Science
Letters, 222 (1), 43–60.
Bonnet, E., Bour, O., Odling, N.E., Davy, P., Main, I., Cowie, P., and Berkowitz, B., 2001: Scaling of
fracture systems in geoligcal media. Review of Geophysics, 39, 347–383.
Bons, P.D., 2000: The formation of veins and their microstructures, in Jessell, M. and Urai, J. eds.,
Stress, Structure and Strain: a volume in honour of Win. D. Means, Journal of the Virtual Explorer,
49.
Bons, P.D., 2001: The formation of large quartz veins by rapid ascent of fluids in mobile hydrofractures.
Tectonophysics, 336 (1–4), 1–17.
Bons, P.D., and Montenari, M., 2005: The formation of antitaxial calcite veins with well-developed
fibres, Oppaminda Creek, South Australia. Journal of Structural Geology, 27, 231–248.
Bons, P.D., Druguet, E., Castaño, L., and Elburg, M.A., 2008: Finding what is now not there anymore:
Recognizing missing fluid and magma volumes. Tectonics, 36 (11), 851–854.
Bons, P.D., Montenari, Æ.M., Bakker, R.J., and Elburg, M.A., 2009: Potential evidence of fossilised
Neoproterozoic deep life: SEM observations on calcite veins from Oppaminda Creek, Arkaroola,

303
References

South Australia. International Journal of Earth Sciences, 98, 327–343.


Bons, P.D., Elburg, M. a., and Gomez-Rivas, E., 2012: A review of the formation of tectonic veins and
their microstructures. Journal of Structural Geology, 43, 33–62.
Boschman, L.M., van Hinsbergen, D.J.J., Torsvik, T.H., Spakman, W., and Pindell, J.L., 2014: Kinematic
reconstruction of the Caribbean region since the Early Jurassic. Earth-Science Reviews, 138, 102–
136.
BouDagher-Fadel, M.K., 2008: Evolution and Geological significance of Larger Benthic Foraminifera.
Elsevier B.V., 1-547.
BouDagher-Fadel, M.K., 2015: Biostratigraphic and Geological Significance of Planktonic Foraminifera.
UCLpress, 1-320.
Boullier, A., and Robert, F., 1992: Palaeoseismic events recorded in Archaean gold-quartz vein
networks, Val d’Or, Abitibi, Quebec, Canada. Journal of Structural Geology, 14 (2), 161–179.
Bourbié, T., Coussy, O., and Zinszner, B., 1987: Acoustics of Porous Media. Éditions Technip, 334.
Bourgueil, B., Andreieff, P., Lasnier, J., Gonnard, R., Le Métour, J., and Rançon, J.-P., 1988: Synthèse
géologique de la République d’Haïti, in Technical report- ATN/SF 2506 HA, Port-au-Prince, Bureau
des Mines et de l’Énergie d’Haïti, 332.
Bouysse, P., 1988: Opening of the Grenada back-arc Basin and evolution of the Caribbean plate during
the Mesozoic and Early Paleogene. Tectonophysics, 149, 121–143.
Bowin, C.O., 1966: Geology of central Dominican Republic: A case history of part of an island arc.
Geological Society of America Memoir, 98, 11–85.
Bowin, C.O., 1968: Geophysical study of the Cayman Trough. Journal of Geophysical Research, 73 (16),
5159–5173.
Bowin, C.O., 1975: The geology of Hispaniola, in Nairn, A.E.M. and Stehli, F.G. eds., The Gulf of Mexico
and the Caribbean (The ocean basins and margins), New York, Plenum, 501–552.
Bradbury, K.K., Davis, C.R., Shervais, J.W., Janecke, S.U., and Evans, J.P., 2015: Composition, Alteration,
and Texture of Fault-Related Rocks from Safod Core and Surface Outcrop Analogs: Evidence for
Deformation Processes and Fluid-Rock Interactions. Pure and Applied Geophysics, 172 (5), 1053–
1078.
De Bresser, J.H.P., and Spiers, C.J., 1997: Strength characteristics of the r, f, and c slip systems in calcite.
Tectonophysics, 272, 1–23.
ten Brink, U.S., Coleman, D.F., and Dillon, W.P., 2002: The nature of the crust under Cayman Trough
from gravity. Marine and Petroleum Geology, 19, 971–987.
ten Brink, U.S., Marshak, S., and Granja Bruna, J.-L., 2009: Bivergent thrust wedges surrounding oceanic
island arcs: Insight from observations and sandbox models of the northeastern Caribbean plate.
Geological Society of America Bulletin, 121 (11–12), 1522–1536.
ten Brink, U.S., Bakun, W.H., and Flores, C.H., 2011: Historical perspective on seismic hazard to
Hispaniola and the northeast Caribbean region. Journal of Geophysical Research, 116 (B12318),
1–15.
ten Brink, U.S., Bakun, W.H., and Flores, C.H., 2013: Seismic hazard from the Hispaniola subduction
zone: Correction to “Historical perspective on seismic hazard to Hispaniola and the northeast
Caribbean region.” Journal of Geophysical Research: Solid Earth, 116 (12), 1–15.

304
References

Budai, J.M., Lohmann, K.C., and Owen, R.M., 1983: Burial dedolomite in the Mississippian Madison
limestone, Wyoming and Utah thrust belt. Journal of Sedimentary Petrology, 54, 276–288.
Bullock, R.J., Paola, N. De, Holdsworth, R.E., and Trabucho-Alexandre, J., 2014: Lithological controls on
the deformation mechanisms operating within carbonate-hosted faults during the seismic cycle.
Journal of Structural Geology, 58, 22–42.
Burke, K., Fox, P.J., and Sengör, A.M.C., 1978: Buoyant Ocean Floor and the Evolution of the Caribbean.
Journal of Geophysical research, 83 (B8), 3949–3954.
Burke, K., 1988: Tectonic evolution of the Caribbean. Annual Review of Earth and Planetary Sciences,
16, 201–230.
Burkhard, M., 1990: Ductile deformation mechanisms in micritic limestones naturally deformed at low
temperatures (150-350 ° C), in Knipe, R.J. and Rutter, E.H. eds., Deformation Mechanisms,
Rheology and Tectonics, Geological Society, London, Special Publications, 241–257.
Burkhard, M., 1993: Calcite twins, their geometry, appearance and significance as stress-strain markers
and inidcators of tectonic regime: a review. Journal of Structural Geology, 15 (3–5), 351–368.
Bussolotto, M., Benedicto, A., Invernizzi, C., Micarelli, L., Plagnes, V., and Deiana, G., 2007: Deformation
features within an active normal fault zone in carbonate rocks: The Gubbio fault (Central
Apennines, Italy). Journal of Structural Geology, 29, 2017–2037.
Bussolotto, M., Benedicto, A., Moen-maurel, L., and Invernizzi, C., 2015: Fault deformation
mechanisms and fault rocks in micritic limestones: Examples from Corinth rift normal faults.
Journal of Structural Geology, 77, 191–212.
Butterlin, J., 1950: Sur la structure de l’île d’Haïti, Grandes Antilles. C.R. Ac. Sc. Paris, 231 (20), 1074–
1076.
Butterlin, J., 1953: Données nouvelles sur la géologie de la République d’Haïti. Bulletin de la Societe
Géologique de France,, 283–291.
Butterlin, J., 1954: La géologie de la République d’Haïti et ses rapports avec celle des régions voisines.
Publ. Comité 150 ème Anniv. Indépendance, P-Au-P,, 446.
Butterlin, J., 1956: La constitution geologique et la structure des Antilles. C.N.R.S., Paris,, 453.
Butterlin, J., 1960: Géologie générale et régionale de la Républiqe d’Haïti. Univ. Paris, Travaux Mém.
Inst. Htes. Etud. Amér. Latine, VI,, 194.
Butterlin, J., 1972: Regards sur l’origine et l’evolution des unites structurales de la region des Caraïbes.
Bulletin de la Societe Géologique de France, 14 (1–5), 46–54.
Butterlin, J., Maurrasse, F.J., Pierre Louis, F., and Sigal, J., 1976: Les basaltes Cretaces de la Presqu’île
du Sud de la Republique d’Haïti et leurs relations avec ceux du soubassement de la mer des
Caraïbes. 4 ème R.A.S.T., Paris, Soc. Géol. Fr. Edit.,, 83.
Butterlin, J., 1981: Claves para la determinacion de macroforaminiferos de Mexico y del Caraïbe, del
Cretacico superior al Miocene medio. Inst. Mex. Petral.,, 219.
Caine, J.S., Evans, J.P., and Forster, C.B., 1996: Fault zone architecture and permeability structure.
Geology, 24 (11), 1025–1028.
Calais, E., Béthoux, N., and Mercier de Lépinay, B., 1992: From transcurrent faulting to frontal
subduction: A seismotectonic study of the northern Caribbean plate boundary from Cuba to
Puerto Rico. Tectonics, 11 (1), 114–123.

305
References

Calais, E., and Mercier de Lépinay, B., 1995: Strike-Slip Tectonic Processes in the Northern Caribbean
Between Cuba and Hispaniola (Windward Passage). Marine Geophysical Researches, 17, 63–95.
Calais, E., Perrot, J., and Mercier de Lépinay, B., 1998: Strike-slip tectonics and seismicity along the
northern Caribbean plate boundary from Cuba to Hispaniola, in Dolan, J.F. and Mann, P. eds.,
Active Strike-Slip and Collisional Tectonics of the Northern Caribbean Plate Boundary Zone,
Boulder, Geological Society of America Special Papers, 125–169.
Calais, E., Freed, A., Mattioli, G., Amelung, F., Jónsson, S., Jansma, P., Hong, S.-H., Dixon, T., Prépetit,
C., and Momplaisir, R., 2010: Transpressional rupture of an unmapped fault during the 2010 Haiti
earthquake. Nature Geoscience, 3 (11), 794–799.
Calais, E., Symithe, S., Mercier de Lépinay, B., and Prépetit, C., 2016: Plate boundary segmentation in
the northeastern Caribbean from geodetic measurements and Neogene geological observations.
C.R. Geoscience, 348, 42–51.
Calmus, T., 1983: Contribution à l’étude géologique du Massif de Macaya (Sud-ouest d’Haïti, Grandes
Antilles). Sa place dans l’évolution de l’orogène Nord-Caraïbe: Universite Pierre et Marie Curie,
222 p.
Calmus, T., 1984: Decrochement senestre sud-Haïtien: Analyses et consequences paleogeographiques
dans la region de Camp-Perrin (Massif de Macaya, Presqu’île du sud d’Haïti). Ann. Soc. Géol. Nord,
103, 309–316.
Calmus, T., 1987: Études géochimiques des volcanismes crétacé et Tertiare du Massif de Macaya
(presqu’île du Sud d’Haïti): Leur place dans l’évolution tectonique de la région nord-caraïbe. C.R.
Ac. Sc., Paris, II (16), 981–986.
Calmus, T., and Vila, J.-M., 1988: The Massif of Macaya (Haiti): the evolution of a Laramide structure in
the left-handed-strike-slip boundary between North-America and Caribbean Plates. Bol. Depto.
Geol. Uni-Son., 5 (1), 63–69.
Capezzuoli, E., Ruggieri, G., Rimondi, V., Brogi, A., Liotta, D., Alçiçek, M.C., Alçiçek, H., Bülbül, A.,
Gandin, A., Meccheri, M., Shen, C.-C., and Baykara, M.O., 2018: Calcite veining and feeding
conduits in a hydrothermal system: Insights from a natural section across the Pleistocene
Gölemezli travertine depositional system (western Anatolia, Turkey). Sedimentary Geology, 364,
180–203.
Carpenter, B.M., Marone, C., and Saffer, D.M., 2011: Weakness of the San Andreas Fault revealed by
samples from the active fault zone. Nature Geoscience, 3 (2), 1–4.
Carpenter, B.M., Saffer, D.M., and Marone, C., 2012: Frictional properties and sliding stability of the
San Andreas fault from deep drill core. Geology, 40, 759–762.
Case, J.E., Holcombe, T.L., and Martin, R.G., 1984: Map of geologic provinces in the Caribbean region.
Geological Society of America Memoir, 162, 1–30.
Cello, G., 2000: A quantitative structural approach to the study of active fault zones in the Apennines
(Peninsular Italy). Journal of Geodynamics, 29, 265–292.
Cello, G., Tondi, E., Micarelli, L., and Invernizzi, C., 2001: Fault zone fabrics and geofluid properties as
indicators of rock deformation modes. Journal of Geodynamics, 32, 543–565.
Cervantes, P., and Wiltschko, D. V, 2010: Tip to midpoint observations on syntectonic veins, Ouachita
orogen, Arkansas: Trading space for time. Journal of Structural Geology, 32 (8), 1085–1100.
Chauvin, A., Bazhenov, M.L., and Beaudouin, T., 1994: A reconnaissance paleomagnetic study of
Cretaceous rocks from Central Cuba. Geophysical Research Letters, 21 (16), 1691–1694.

306
References

Cheilletz, A., 1976: Etude geologique et metallogenique des indices a cuivre et molybdene de type
porphyre cuprifere de la zone de vert-de-gris-Jean-Rabel, Presqu’île du Nord-Ouest d’Haïti: 117
p.
Cheilletz, A., and Lewis, J.F., 1976: Contribution a l’etude de la bordure meridionale du massif du NNE
d’Haïti. 7 ème Caribbean Geological Conference Transactions, 7, 243–247.
Cheilletz, A., Kachrillo, J.J., Sonet, J., and Zimmermann, J.L., 1978: Petrographie et geochronologie de
deux complexes intrusifs a pophyres cupriferes d’Haïti. Contribution a la connaissance de la
province cuprifere Laramienne de l’arc insulaire des Grandes Antilles. Bulletin de la Societe
Géologique de France, 20 (7), 907–914.
Chester, F.M., and Logan, J.M., 1986: Implications for Mechanical Properties of Brittle Faults from
Observations of the Punchbowl Fault Zone, California. Pure and Applied Geophysics, 124 (1–2),
79–106.
Childs, C., Manzocchi, T., Walsh, J.J., Bonson, C.G., Nicol, A., and Scho, M.P.J., 2009: A geometric model
of fault zone and fault rock thickness variations. Journal of Structural Geology, 31, 117–127.
Cianfarra, P., and Salvini, F., 2016: Quantification of fracturing within fault damage zones affecting Late
Proterozoic carbonates in Svalbard. Rendiconti Lincei, 27, 229–241.
Cilona, A., Baud, P., Tondi, E., Agosta, F., Vinciguerra, S., Rustichelli, A., and Spiers, C.J., 2012:
Deformation bands in porous carbonate grainstones: Field and laboratory observations. Journal
of Structural Geology, 45, 137–157.
Clemenzi, L., Molli, G., Storti, F., Muchez, P., Swennen, R., and Torelli, L., 2014: Extensional deformation
structures within a convergent orogen: The Val di Lima low-angle normal fault system (Northern
Apennines, Italy). Journal of Structural Geology, 66, 205–222.
Coleman, A.J., and Winslow, M.A., 1999: Tertiary tectonics of the Hispaniola Fault Zone in the
northwestern piedmont of the Cordillera Central, Dominican Republic, in Mann, P., Grindlay, N.,
and Dolan, J. eds., Penrose Conference: Subduction to strike-slip transitions on plate boundaries,
Geological Society of America, 1–2.
Corbeau, J., Rolandone, F., Leroy, S., Meyer, B., and Mercier de Lépinay, B Momplaisir, R., 2016a: How
transpressive is the northern Caribbean plate boundary? Tectonics, 35, 1032–1046.
Corbeau, J., Rolandone, F., Leroy, S., Mercier de Lépinay, B., Meyer, B., Ellouz-Zimmermann, N., and
Momplaisir, R., 2016b: The northern Caribbean plate boundary in the Jamaica Passage: Structure
and seismic stratigraphy. Tectonophysics, 675, 209–226.
Corbeau, J., Rolandone, F., Leroy, S., Guerrier, K., Keir, D., Stuart, G., Clouard, V., Gallacher, R., Ulysse,
S., Boisson, D., Momplaisir, R., Saint Preux, F., Prépetit, C., Saurel, J., et al., 2017: Crustal structure
of western Hispaniola (Haiti) from a teleseismic receiver function study. Tectonophysics, 709, 9–
19.
Cordey, F.A., and Cornée, J.-J., 2009: New radiolarian assemblages from La Désirade Island basement
complex (Guadeloupe, Lesser Antilles arc) and Caribbean tectonic implications. Bulletin de la
Societe Géologique de France, 180 (5), 399–409.
Cowan, D.S., 1999: Do faults preserve a record of seismic slip: A field geologist’s opinion. Journal of
Structural Geology, 21, 995–1001.
Cox, F., 1995: Faulting processes at high fluid pressures: An example of fault valve behavior from the
Wattle Gully Fault, Victoria, Australia. Journal of Geophysical Research, 100 (B7), 12841–12859.
Crawford, B.R., Faulkner, D.R., and Rutter, E.H., 2008: Strength, porosity, and permeability

307
References

development during hydrostatic and shear loading of synthetic quartz-clay fault gouge. Journal
of Geophysical Research, 113, 1–14.
Cruz-Orosa, I., Sàbat, F., Ramos, E., Rivero, L., and Vázquez-Taset, Y.M., 2012: Structural evolution of
the La Trocha fault zone: Oblique collision and strike-slip basins in the Cuban Orogen. Tectonics,
31 (TC5001), 1–23.
D’Meza, S., 1989: Etude pétrographique de quelques laves cénozoïques de Centre d’Haïti.
Van Daalen, M., Heilbronner, R., and Kunze, K., 1999: Orientation analysis of localized shear
deformation in quartz fibres at the brittle-ductile transition. Tectonophysics, 303 (1–4), 83–107.
Daniel, J., Moretti, I., Micarelli, L., Chuyne, S.E., and Piane, C.D., 2004: Macroscopic structural analysis
of AG10 well (Gulf of Corinth, Greece). Tectonics, 336, 435–444.
Darin, M.H., and Dorsey, R.J., 2013: Reconciling disparate estimates of total offset on the southern San
Andreas fault. Geology, 41 (9), 975–978.
Dawers, N.H., and Anders, M.H., 1995: Displacement-length scaling and fault linkage. Journal of
Structural Geology, 17 (5), 607–614.
Delle Piane, C., Clennell, M. Ben, Keller, J.V.A., Giwelli, A., and Luzin, V., 2017: Carbonate hosted fault
rocks: A review of structural and microstructural characteristic with implications for seismicity in
the upper crust. Journal of Structural Geology, 103, 17–36.
DeMets, C., and Dixon, T.H., 1999: New kinematic models for Pacific-North America motion from 3 Ma
to present, I: Evidence for steady motion and biases in the NUVEL-1A model. Geophysical
Research Letters, 26 (13), 1921–1924.
DeMets, C., Jansma, P.E., Mattioli, G.S., Dixon, T.H., Farina, F., Bilham, R., Calais, E., and Mann, P., 2000:
GPS geodetic constraints on Caribbean-North America plate motion. Geophysical Research
Letters, 27 (3), 437–440.
DeMets, C., 2001: A new estimate for present-day Cocos-Caribbean plate motion: Implications for slip
along the Central American volcanic arc. Geophysical Research Letters, 28 (21), 4043–4046.
DeMets, C., Mattioli, G., Jansma, P., Rogers, R.D., Tenorio, C., and Turner, H.L., 2007: Present motion
and deformation of the Caribbean plate: Constraints from new GPS geodetic measurements from
Honduras and Nicaragua. Geological Society of America Special Paper, 428 (2), 21–36.
DeMets, C., and Wiggins-Grandison, M., 2007: Deformation of Jamaica and motion of the Gonâve
microplate from GPS and seismic data. Geophysical Journal International, 168 (1), 362–378.
DeMets, C., Gordon, R.G., and Argus, D.F., 2010: Geologically current plate motions. Geophysical
Journal International, 181 (1), 1–80.
Desreumaux, C., 1985a: Sur la presence d’une serie carbonatee contiue de type bassin Caraïbe du
Cretace terminal au Miocene dans la Presqu’île du sud d’Haïti, Grandes Antilles. C. R. Ac. Sc., Paris,
5 (II), 319–322.
Desreumaux, C., 1985b: Haïti: Un modele recent et actuel de systeme compressif a effet centripete.
Géodynamique des Caraïbes, ed. Technip, Paris,, 391–402.
Desreumaux, C., 1987: Contribution a l’etude de l’histoire geologique des regions centrale et
meridional d’Haïti (Grandes Antilles) du Cretace à l’actuel: 507 p.
Diebold, J.B., Stoffa, P.L., Buhl, P., and Truchan, M., 1981: Venezuela Basin Crustal Structure. Journal
of Geophysical Research, 86 (B9), 7901–7923.

308
References

Diebold, J., Driscoll, N., Abrams, L., Buhl, P., Donnelly, T., Laine, E., Leroy, S., and Toy, A., 1999: New
Insights on the Formation of the Caribbean Basalt Province Revealed by Multichannel Seismic
Images of Volcanic Structures in the Venezuelan Basin, in Mann, P. ed., Sedimentary Basins of the
World; Caribbean Basins, Amsterdam, Elsevier Science B.V., 561–589.
Dillon, W.P., Austin, J.A., Scanlon, K.M., and Parson, L.M., 1992: Accretionary margin of north-western
Hispaniola: morphology, structure and development of part of the northern Caribbean plate
boundary. Marine and Petroleum Geology, 9 (February), 70–88.
Dixon, T.H., Farina, F., DeMets, C., Jansma, P., Mann, P., and Calais, E., 1998: Relative motion between
the Caribbean and North American plates and related boundary zone deformation from a decade
of GPS observations. Journal of Geophysical Research, 103 (B7), 15157–15182.
Dodge, R.E., Fairbanks, R.G., Benninger, L.K., and Maurrasse, F.. J., 1983: Pleistocene sea levels from
raised coral reefs of Haïti. Science, 219 (4591), 1423–1425.
Dolan, J., Mann, P., de Zoeten, R., Heubeck, C., Shiroma, J., and Monechi, S., 1991: Sedimentologic,
stratigraphic, and tectonic synthesis of Eocene-Miocene sedimentary basins, Hispaniola and
Puerto Rico, in Mann, P., Draper, G., and Lewis, J.F. eds., Geologic and Tectonic Development of
the North America-Caribbean Plate Boundary in Hispaniola, Boulder, Geological Society of
America Special Papers, 217–263.
Dolan, J.F., and Wald, D.J., 1998: The 1943 – 1953 north-central Caribbean earthquakes: Active tectonic
setting, seismic hazards, and implications for Caribbean – North America plate motions, in Dolan,
J.F. and Mann, P. eds., Active Strike-Slip and Collisional Tectonics of the Northern Caribbean Plate
Boundary Zone, Boulder, Geological Society of America Special Papers, 143–169.
Dong, G., Morrison, G., and Jaireth, S., 1995: Quartz textures in epithermal veins, Queensland -
classification, origin, and implication. Economic Geology, 90 (6), 1841–1856.
Donnelly, T.W., Melson, W., Kay, R., and Rogers, J.J.W., 1973: Basalts and dolerites of late cretaceous
age from the central caribbean, in Edgar, N.T. and Saunders, J.B. eds., Initial reports DSDP,
Washington D.C., US Government Printing Office, 989–1011.
Donnelly, T.W., 1994: The Caribbean Sea Floor, in Caribbean Geology: An Introduction, Kingston, U.W.I.
Publishers’ Association, 41–64.
Douilly, R., Haase, J.S., Ellsworth, W.L., Bouin, M.-P., Calais, E., Symithe, S.J., Armbruster, J.G., Mercier
de Lépinay, B., Deschamps, A., Mildor, S.-L., Meremonte, M.E., and Hough, S.E., 2013: Crustal
structure and fault geometry of the 2010 Haiti earthquake from temporary seismometer
deployments. Bulletin of the Seismological Society of America, 103 (4), 2305–2325.
Draper, G., 1987: A revised tectonic model for the evolution of Jamaica, in Ahmad, R. ed., Proceedings
of a Workshop on the Status of Jamaican Geology, Kingston, Special Publications of the Geological
Society of Jamaica, 151–169.
Draper, G., and Lewis, J.F., 1991: Metamorphic belts in central Hispaniola. Geological Society of
America Special Paper, 262, 29–45.
Draper, G., Gutiérrez, G., and Lewis, J.F., 1996: Thrust emplacement of the Hispaniola peridotite belt:
Orogenic expression of the mid-Cretaceous Caribbean arc polarity reversal? Geology, 24 (12),
1143–1146.
Draper, G., 1999: The transition from Cretaceous convergence to Cenozoic strike-slip in Hispaniola and
eastern Cuba, in Mann, P., Grindlay, N., and Dolan, J. eds., Penrose Conference: Subduction to
strike-slip transitions on plate boundaries, Geological Society of America, 27.

309
References

Draper, G., 2008: Some speculations on the Paleogene and Neogene tectonics of Jamaica. Geological
Journal, 43, 563–572.
Draper, G., and Lewis, J., 2008: Field trip guide to the Median Belt and subduction zone rocks,
Dominican Republic, in Abbott, R.A., Escuder-Viruete, J., Krebs, M., Maresch, W.V., Schertl, P.,
Perez-Estaun, A., and Pindell, J.L. eds., 18th Caribbean Geological Conference, Santo Domingo,
59.
Driscoll, N.W., and Diebold, J.B., 1999: Tectonic and Stratigraphic Development of the Eastern
Caribbean: New Constraints from Multichannel Seismic Data, in Mann, P. ed., Sedimentary Basins
of the World; Caribbean Basins, Amsterdam, Elsevier Science B.V., 591–626.
Dubreuilh, P., 1982: Contribution à l’etude du bassin Neogene du Plateau Central d’Haïti: 156 p.
Dubreuilh, P., 1984: Evolution du bassin sedimentaire du centre d’Haiti au Neogene. Bulletin de la
Institute Géologique du Bassin Aquitaine, 35, 53–76.
Duncan, R.A., and Hargraves, R.B., 1984: Plate tectonic evolution of the Caribbean region in the mantle
reference frame. Geological Society of America, 162, 81–93.
Dunlap, W.J., Hirth, G., and Teyssier, C., 1997: Thermomechanical evolution of a ductile duplex.
Tectonics, 16 (6), 983–1000.
Dunne, W.M., and North, C.P., 1990: Orthogonal fracture systems at the limits of thrusting: an example
from southwestern Wales. Journal of Structural Geology, 12 (2), 207–215.
Duplan, L., 1975: Etude photogéologique de la région sud de la République d’Haïti (Nations-Unies),
projet de développement minier.
Durney, D.W., and Ramsay, J.G., 1973: Incremental strains measured by syntectonic crystal growths,
in De Jong, K.A. and Scholten, K. eds., Gravity and Tectonics, New York, Wiley, 67–96.
Van Dusen, S.R., and Doser, D.I., 2000: Faulting processes of historic (1917-1962) M >= 6.0 earthquakes
along the north-central Caribbean margin. Pure and Applied Geophysics, 157 (5), 719–736.
Edgar, N.T., Saunders, J.B., Bolli, H.M., Donnelly, T.W., Hay, W.W., Maurrasse, F., Silva, I.P., Riedel,
W.R., and Schneidermann, N., 1973: DSDP Leg 15, Introduction, in Edgar, N.T. and Saunders, J.B.
eds., Initial Reports of the Deep Sea Drilling Project, 3–15.
Edgar, N.T., 1991: Structure and geologic development of the Cibao Valley, northern Hispaniola, in
Mann, P., Draper, G., and Lewis, J.F. eds., Geologic and Tectonic Development of the North
America-Caribbean Plate Boundary in Hispaniola, Boulder, Geological Society of America Special
Papers, 281–299.
Ehlert, K.W., 2003: Tectonic significance of the middle Miocene Mint Canyon and Caliente Formations,
southern California, in Crowell, J.C. ed., Evolution of Ridge Basin, soutern California: An interplay
of sedimentation and tectonics, Geological Society of America Special paper 367, 113–130.
Ehlig, P.L., Ehler, K.W., and Crowe, B.M., 1975: Offset of the Upper Miocene Caliente and Mint Canyon
Formations along the San Gabriel and San Andreas faults, in Crowell, J.C. ed., San Andreas fault
in southern California, California Division of Mines and Geology Special Report, 83–92.
Einsele, G., 2000: The Interplay Between Sediment Supply, Subsidence, and Basin Fill, in Sedimentary
Basins: Evolution, Facies, and Sediment Budget, Springer-Verlag, 792.
Electroconsult, 1983: Estudio de pre-factibilidad de area geotérmica, Yaya-Constanza, informe
fotogeologico. Dirección General de Minería: Santo Domingo, Republica Dominicana,.
Ellouz-Zimmermann, N., Leroy, S., Momplaisir, R., Mercier de Lépinay, B., and Group, H.-S., 2013: From

310
References

2012 HAITI-SIS Survey: thick-skin versus thin-skin tectonics partitioned along offshore strike-slip
faults - Haiti, in AGU San Francisco, 1.
Ellouz-Zimmermann, N., Hamon, Y., Deschamps, R., Battani, A., Schmitz, J., Ruffine, L., Momplaisir, R.,
and Leroy, S., 2014: Along-Fault Deformation Partitioning of NW Haiti: Implication on Fluid
Transfer, in AGU San Francisco, 1.
Ellouz-Zimmermann, N., and Pubellier, M., 2015: 889 IODP Haiti-DRILL Pre-proposal.:
Engelder, T., 1982: Is there a genetic relationship between selected regional joints and contemporary
stress within the lithosphere of North America. Tectonics, 1 (2), 161–177.
Engelder, T., 1985: Loading paths to joint propagation during a tectonic cycle: an example from the
Appalachian Plateau, U.S.A. Journal of Structural Geology, 7 (3–4), 459–476.
Erikson, J.P., Pindell, J.L., Karner, G.D., Sonder, L.J., Fuller, E., and Dent, L., 1998: Neogene
Sedimentation and Tectonics in the Cibao Basin and Northern Hispaniola: An Example of Basin
Evolution Near A Strike-Slip-Dominated Plate Boundary. The Journal of Geology, 106 (4), 473–
494.
Escuder-Viruete, J., Hernaiz Huerta, P.P., Draper, G., Gutiérrez, G., Lewis, J.F., and Pérez-Estaún, A.,
2002: The metamorphism and structure of the Maimón Formation and Duarte and Río Verde
Complexes, Dominican Central Cordillera: implications for the structure and evolution of the
primitive Caribbean Island Arc. Acta Geologica Hispanica, 37 (2–3), 123–162.
Escuder-Viruete, J., and Pérez-Estaún, A., 2004: Trayectoria metamórfica P-T relacionada con
subducción en eclogitas del Complejo de Basamento de Samaná, Cordillera Septentrional,
República Dominicana. Geo-Temas, 6, 37–44.
Escuder-Viruete, J., Díaz de Neira, A., Hernáiz Huerta, P.P., Monthel, J., Senz, J.G., Joubert, M., Lopera,
E., Ullrich, T., Friedman, R., and Mortensen, J., 2006a: Magmatic relationships and ages of
Caribbean Island arc tholeiites, boninites and related felsic rocks, Dominican Republic. Lithos, 90
(3–4), 161–186.
Escuder-Viruete, J., Contreras, F., Stein, G., Urien, P., Joubert, M., Ullrich, T., Mortensen, J., and Pérez-
Estaún, A., 2006b: Transpression and strain partitioning in the Caribbean Island-arc: Fabric
development, kinematics and Ar–Ar ages of syntectonic emplacement of the Loma de Cabrera
batholith, Dominican Republic. Journal of Structural Geology, 28 (8), 1496–1519.
Escuder-Viruete, J., and Pérez-Estaún, a., 2006: Subduction-related P–T path for eclogites and garnet
glaucophanites from the Samaná Peninsula basement complex, northern Hispaniola.
International Journal of Earth Sciences, 95 (6), 995–1017.
Escuder-Viruete, J., Pérez-Estaún, A., and Weis, D., 2007a: Geochemical constraints on the origin of the
late Jurassic proto-Caribbean oceanic crust in Hispaniola. International Journal of Earth Sciences,
98 (2), 407–425.
Escuder-Viruete, J., Contreras, F., Stein, G., Urien, P., Joubert, M., Pérez-Estaún, A., Friedman, R., and
Ullrich, T., 2007b: Magmatic relationships and ages between adakites, magnesian andesites and
Nb-enriched basalt-andesites from Hispaniola: Record of a major change in the Caribbean island
arc magma sources. Lithos, 99 (3–4), 151–177.
Escuder-Viruete, J., Pérez-Estaún, A., Contreras, F., Joubert, M., Weis, D., Ullrich, T.D., and Spadea, P.,
2007c: Plume mantle source heterogeneity through time: Insights from the Duarte Complex,
Hispaniola, northeastern Caribbean. Journal of Geophysical Research, 112 (B4), 1–19.
Escuder-Viruete, J., Joubert, M., Urien, P., Friedman, R., Weis, D., Ullrich, T., and Pérez-Estaún, A., 2008:

311
References

Caribbean island-arc rifting and back-arc basin development in the Late Cretaceous:
Geochemical, isotopic and geochronological evidence from Central Hispaniola. Lithos, 104 (1–4),
378–404.
Escuder-Viruete, J., and Pérez-Estaún, A., 2008: The Rio Verde Complex, Central Hispaniola: A fragment
of Lower Cretaceous back-arc basin of the primitive Caribbean island-arc, in Abstracts and
Program 18th Caribbean Geological Conference, Santo Domingo, 40–41.
Escuder-Viruete, J., Pérez-Estaún, A., Weis, D., and Friedman, R., 2010: Geochemical characteristics of
the Río Verde Complex, Central Hispaniola: Implications for the paleotectonic reconstruction of
the Lower Cretaceous Caribbean island-arc. Lithos, 114 (1–2), 168–185.
Escuder-Viruete, J., Pérez-Estaún, A., Joubert, M., and Weis, D., 2011a: The Pelona-Pico Duarte basalts
Formation, Central Hispaniola: an on-land section of Late Cretaceous volcanism related to the
Caribbean large igneous province. Geologica Acta, 9 (3–4), 307–328.
Escuder-Viruete, J., Pérez-Estaún, A., Booth-Rea, G., and Valverde-Vaquero, P., 2011b:
Tectonometamorphic evolution of the Samaná complex, northern Hispaniola: Implications for
the burial and exhumation of high-pressure rocks in a collisional accretionary wedge. Lithos, 125,
190–210.
Escuder-Viruete, J., Pérez-Estaún, A., Gabites, J., and Suárez-Rodríguez, Á., 2011c: Structural
development of a high-pressure collisional accretionary wedge: The Samaná complex, Northern
Hispaniola. Journal of Structural Geology, 33 (5), 928–950.
Escuder-Viruete, J., Friedman, R., Castillo-Carrión, M., Jabites, J., and Pérez-Estaún, A., 2011d: Origin
and significance of the ophiolitic high-P mélanges in the northern Caribbean convergent margin:
Insights from the geochemistry and large-scale structure of the Río San Juan metamorphic
complex. Lithos, 127, 483–504.
Escuder-Viruete, J., Valverde-Vaquero, P., Rojas-Agramonte, Y., Jabites, J., and Pérez-Estaún, A., 2013:
From intra-oceanic subduction to arc accretion and arc-continent collision: Insights from the
structural evolution of the Río San Juan metamorphic complex, northern Hispaniola. Journal of
Structural Geology, 46, 34–56.
Escuder-Viruete, J., Castillo-Carrión, M., and Pérez-Estaún, A., 2014: Magmatic relationships between
depleted mantle harzburgites, boninitic cumulate gabbros and subduction-related tholeiitic
basalts in the Puerto Plata ophiolitic complex, Dominican Republic: Implications for the birth of
the Caribbean island-arc. Lithos, 196–197, 261–280.
Evans, J.P., 1990: Thickness--displacement relationships for fault zones. Journal of Structural Geology,
12 (8), 1061–1065.
Evans, J.P., Forster, C.B., and Goddard, J. V., 1997: Permeability of fault-related rocks, and implications
for hydraulic structure of fault zones. Journal of Structural Geology, 19 (11), 1393–1404.
Ewing, J., Antoine, J., and Ewing, M., 1960: Geophysical measurements in the Western Caribbean Sea
and in the Gulf of Mexico. Journal of Geophysical Research, 65 (12), 4087–4126.
Famin, V., Raimbourg, H., Garcia, S., Bellahsen, N., Hamada, Y., Boullier, A.M., Fabbri, O., Michon, L.,
Uchide, T., Ricci, T., Hirono, T., and Kawabata, K., 2014: Stress rotations and the long-term
weakness of the Median Tectonic Line and the Rokko-Awaji Segment. Tectonics, 33 (10), 1900–
1919.
Faulkner, D.R., Rutter, E.H., Andrew, S., and Smith, F., 1998: The gas permeability of clay-bearing fault
gouge at 20°C. Geological Society, London, Special Publications, 147, 147–156.

312
References

Faulkner, D.R., and Rutter, E.H., 2000: Comparisons of water and argon permeability in natural clay-
bearing fault gouge under high pressure at 20°C. Journal of Geophysical Research, 105 (B7),
16415–16426.
Faulkner, D.R., Lewis, A.C., and Rutter, E.H., 2003: On the internal structure and mechanics of large
strike-slip fault zones: field observations of the Carboneras fault in southeastern Spain.
Tectonophysics, 367, 235–251.
Faulkner, D.R., 2004: A model for the variation in permeability of clay-bearing fault gouge with depth
in the brittle crust. Geophysical Research Letters, 31, 1–5.
Faulkner, D.R., Jackson, C.A.L., Lunn, R.J., Schlische, R.W., Shipton, Z.K., Wibberley, C.A.J., and Withjack,
M.O., 2010: A review of recent developments concerning the structure, mechanics and fluid flow
properties of fault zones. Journal of Structural Geology, 32 (11), 1557–1575.
Feinberg, H., and Vila, J.M., 1982: A propos de la coupure Oligocene - Miocene: Repartition
stratigraphique du genre Miogypsina à Hispaniola (Grandes Antilles). 9 ème R.A.S.T., Paris,.
Ferket, H., Swennen, R., Ortuño, S., and Roure, F., 2003: Reconstruction of the fluid flow history during
Laramide foreland fold and thrust belt development in eastern Mexico: cathodoluminescence
and d18O-d13C isotope trends of calcite-cemented fractures. Journal of Geochemical
Exploration, 78–79, 163–167.
Ferrill, D.A., Morris, A.P., Evans, M.A., Burkhard, M., Groshong, R.H., and Onasch, C.M., 2004: Calcite
twin morphology: a low-temperature deformation geothermometer. Journal of Structural
Geology, 26, 1521–1529.
Ferrill, D.A., Evans, M.A., McGinnis, R.N., Morris, A.P., Smart, K.J., Wigginton, S.S., Gulliver, K.D.H.,
Lehrmann, D., de Zoeten, E., and Sickmann, Z., 2017a: Fault zone processes in mechanically
layered mudrock and chalk. Journal of Structural Geology, 97, 118–143.
Ferrill, D.A., Morris, A.P., McGinnis, R.N., Smart, K.J., Wigginton, S.S., and Hill, N.J., 2017b: Mechanical
stratigraphy and normal faulting. Journal of Structural Geology, 94, 275–302.
Fisher, D.M., and Brantley, S.L., 1992: Models of Quartz Overgrowth and Vein Formation: Deformation
and Episodic Fluid Flow in an Ancient Subduction Zon. Journal of Geophysical Research, 97 (B13),
20043–20061.
Fisher, D.M., Brantley, S.L., Everett, M., and Dzvonik, J., 1995: Cyclic fluid flow through a regionally
extensive fracture network within the Kodiak accretionary prism. Journal of Geophysical
Research, 100 (B7), 12881–12894.
Fondriest, M., Smith, S.A.F., Candela, T., Nielsen, S.B., Mair, K., and Toro, G. Di, 2013: Mirror-like faults
and power dissipation during earthquakes. Geology, 41 (11), 1175–1178.
Fossen, M.C. Van, and Channell, J.E.T., 1988: Paleomagnetism of Late Cretaceous and Eocene
limestones and chalks from Haiti: Tectonic interpretations. tectonics, 7 (3), 601–612.
Fossen, H., Schultz, R.A., Shipton, Z.K., and Mair, K., 2007: Deformation bands in sandstone: a review.
Journal of the Geological Society, London, 164, 755–769.
Fossen, H., 2010: Structural Geology. Cambridge University Press, 463.
Frery, E., Gratier, J.-P., Ellouz-Zimmerman, N., Deschamps, P., Blamart, D., Hamelin, B., and Swennen,
R., 2017: Geochemical transect through a travertine mount: A detailed record of CO2-enriched
fluid leakage from Late Pleistocene to present-day - Little Grand Wash fault (Utah, USA).
Quaternary International, 437 (A), 98–106.

313
References

Friedman, M., 1972: Residual elastic strain in rocks. Tectonophysics, 15 (4), 297–330.
Frizzell, V.A., Mattinson, J.M., and Matti, J.C., 1986: Distinctive Triassic megaporphyritic monzogranite:
Evidence for only 160 km offset along the San Andreas Fault, southern California. Journal of
Geophysical Research, 91 (6), 14080–14088.
García-Casco, A., Iturralde-Vinent, M.A., and Pindell, J., 2008: Latest Cretaceous Collision/Accretion
between the Caribbean Plate and Caribeana: Origin of Metamorphic Terranes in the Greater
Antilles. International Geology Review, 50 (9), 781–809.
Geldmacher, J., Hanan, B.B., Blichert-Toft, J., Harpp, K., Hoernle, K., Hauff, F., Werner, R., and Kerr, a.
C., 2003: Hafnium isotopic variations in volcanic rocks from the Caribbean Large Igneous Province
and Galápagos hot spot tracks. Geochemistry, Geophysics, Geosystems, 4 (7), n/a-n/a.
van Gestel, J.-P., Mann, P., Dolan, J.F., and Grindlay, N.R., 1998: Structure and tectonic of the upper
Cenozoic Puerto Rico-Virgin Islands carbonate platform as determined from seismic reflection
studies. Journal of Geophysical Research, 103 (B12), 30505–30530.
Ghosh, S.K., 1988: Theory of chocolate tablet boudinage. Journal of Structural Geology, 10 (6), 541–
553.
Giger, S.B., Tenthorey, E., Cox, S.F., and Gerald, J.D.F., 2007: Permeability evolution in quartz fault
gouges under hydrothermal conditions. Journal of Geophysical Research, 112, 1–17.
Giorgetti, C., Collettini, C., Scuderi, M.M., Barchi, M.R., and Tesei, T., 2016: Fault geometry and
mechanics of marly carbonate multilayers: An integrated fi eld and laboratory study from the
Northern Apennines, Italy. Journal of Structural Geology, 93, 1–16.
Goldstein, R., and Reynolds, J., 1994: Systematics of Fluid Inclusions in Diagenetic Minerals. Society for
Sedimentary Geology (SEPM), 1-213.
Gómez, J.A., Martin, M., Olmo, A. del, Canales, L., Ruiz, T., Lewis, J., Draper, G., and Pérez-Estaún, A.,
2000: Memoria explicativa del Mapa Geológico a escala 1:50000 de Constanza (6072-I), in
Proyecto de Cartografia Geotemática de la República dominicana, Santo Domingo, 145.
Gordon, B., Mann, P., Cáceres, D., and Flores, R., 1997: Cenozoic tectonic history of the North America-
Caribbean plate boundary zone in western Cuba. Journal of Geophysical Research, 102 (B5),
10055–10082.
Gradstein, F.M., Ogg, J.G., Schmitz, J., and Ogg, G., 2012: The Geologic Time Scale 2012. Elsevier, 1176.
Graham, B., Antonellini, M., and Aydin, A., 2003: Formation and growth of normal faults in carbonates
within a compressive environment. Geological Society of America, 31 (1), 11–14.
Grandia, F., Cardellach, E., Canals, À., and Banks, D.A., 2003: Geochemistry of the Fluids Related to
Epigenetic Carbonate-Hosted Zn-Pb Deposits in the Maestrat Basin, Eastern Spain: Fluid Inclusion
and Isotope (Cl, C, O, S, Sr) Evidence. Economic Geology, 98, 933–954.
Granja Bruña, J.L., ten Brink, U.S., Carbó-Gorosabel, a., Muñoz-Martín, a., and Gómez Ballesteros, M.,
2009: Morphotectonics of the central Muertos thrust belt and Muertos Trough (northeastern
Caribbean). Marine Geology, 263 (1–4), 7–33.
Granja Bruña, J.L., Muñoz-Martín, a., ten Brink, U.S., Carbó-Gorosabel, a., Llanes Estrada, P., Martín-
Dávila, J., Córdoba-Barba, D., and Catalán Morollón, M., 2010: Gravity modeling of the Muertos
Trough and tectonic implications (north-eastern Caribbean). Marine Geophysical Researches, 31
(4), 263–283.
Granja Bruña, J.L., Carbó-Gorosabel, a., Llanes Estrada, P., Muñoz-Martín, a., ten Brink, U.S., Gómez

314
References

Ballesteros, M., Druet, M., and Pazos, a., 2014: Morphostructure at the junction between the
Beata ridge and the Greater Antilles island arc (offshore Hispaniola southern slope).
Tectonophysics, 618, 138–163.
Greene, A.R., Scoates, J.S., Weis, D., Katvala, E.C., and Nixon, G.T., 2010: The architecture of oceanic
plateaus revealed by the volcanic stratigraphy of the accreted Wrangellia oceanic plateau.
Geosphere, 6 (1), 47–73.
Groshong, R.H., 1972: Strain Calculated from Twinning in Calcite. Geological Society of America
Bulletin, 83, 2025–2038.
Guevara, N.O., García, A., and Arnaiz, M., 2013: Magnetic anomalies in the Eastern Caribbean.
International Journal of Earth Sciences, 102 (3), 591–604.
Guja, N.H., and Vincenz, S. a, 1978: Paleomagnetism of some late Cretaceous and Miocene igneous
rocks on Jamaica. Geophysical Journal of the Royal Astronomy Society, 52, 97–115.
Han, R., Shimamoto, T., Hirose, T., Ree, J.-H., and Ando, J. -i., 2007a: Ultralow Friction of Carbonate
Faults Caused by Thermal Decomposition. Science, 316, 878–882.
Han, R., Shimamoto, T., Ando, J.I., and Ree, J.H., 2007b: Seismic slip record in carbonate-bearing fault
zones: An insight from high-velocity friction experiments on siderite gouge. Geology, 35 (12),
1131–1134.
Hancock, P.L., 1969: Jointing in the jurassic limestones of the Cotswold Hills. Proceedings of the
Geologists’ Association, 80 (2), 219–241.
Hancock, P.L., Kadhi, A. Al, and Sha’at, N.A., 1984: Regional joint sets in the Arabian platform as
indicators of intraplate processes. Tectonics, 3 (1), 27–43.
Hancock, P.L., 1985: Brittle microtectonics: principles and practice. Journal of Structural Geology, 7
(3/4), 437–457.
Haq, B.U., Hardenbol, J., and Vail, P.R., 1988: Mesozoic and Cenozoic chronostratigraphy and cycles of
sea-level change. Science, 241, 71–108.
Harlow, G.E., Hemming, S.R., Avé Lallemant, H.G., Sisson, V.B., and Sorensen, S.S., 2004: Two high-
pressure-low-temperature serpentinite-matrix mélange belts, Motagua fault zone, Guatemala: A
record of Aptian and Maastrichtian collisions. Geology, 32 (1), 17–20.
Hastie, A.R., Kerr, A.C., Mitchell, S.F., and Millar, I.L., 2009: Geochemistry and tectonomagmatic
significance of Lower Cretaceous island arc lavas from the Devils Racecourse Formation, eastern
Jamaica. Geological Society, London, Special Publications, 328, 339–360.
Hastie, A.R., 2009: Is the Cretaceous primitive island arc series in the circum-Caribbean region
geochemically analogous to the modern island arc tholeiite series?, in James, K.H., Lorente, M.A.,
and Pindell, J.L. eds., The Origin and Evolution of the Caribbean Plate, Geological Society, London,
Special Publications, 399–409.
Hastie, A.R., and Kerr, A.C., 2010: Mantle plume or slab window?: Physical and geochemical constraints
on the origin of the Caribbean oceanic plateau. Earth-Science Reviews, 98 (3–4), 283–293.
Hastie, A.R., Mitchell, S.F., Treloar, P.J., Kerr, A.C., Neill, I., and Barfod, D.N., 2013: Geochemical
components in a Cretaceous island arc: The Th/La–(Ce/Ce*)Nd diagram and implications for
subduction initiation in the inter-American region. Lithos, 162–163, 57–69.
Hausegger, S., Kurz, W., Rabitsch, R., Kiechl, E., and Brosch, F.-J., 2010: Analysis of the internal structure
of a carbonate damage zone: Implications for the mechanisms of fault breccia formation and fluid

315
References

flow. Journal of Structural Geology, 32, 1349–1362.


Hayes, G.P., Briggs, R.W., Sladen, a., Fielding, E.J., Prentice, C., Hudnut, K., Mann, P., Taylor, F.W.,
Crone, a. J., Gold, R., Ito, T., and Simons, M., 2010: Complex rupture during the 12 January 2010
Haiti earthquake. Nature Geoscience, 3 (11), 800–805.
Hayman, N.W., Grindlay, N.R., Perfit, M.R., Mann, P., Leroy, S., and de Lépinay, B.M., 2011: Oceanic
core complex development at the ultraslow spreading Mid-Cayman Spreading Center.
Geochemistry, Geophysics, Geosystems, 12 (3), 1–21.
Healy, D., 2008: Damage patterns, stress rotations and pore fluid pressures in strike-slip fault zones.
Journal of Geophysical Research, 113, 1–16.
Hennings, P.H., Olson, J.E., and Thompson, L.B., 2000: Combining outcrop data and three-dimensional
structural models to characterize fractured reservoirs: An example from Wyoming. AAPG Bulletin,
84 (6), 830–849.
Hernaiz Huerta, P.P., Lewis, J.F., Escuder-Viruete, J., Gutiérrez, G., Mortenson, J., Hames, W., Solé, J.,
Martínez, A., and Graper, G., 2000: Memoria explicativa del Mapa Geológico a escala 1:50000 de
Arroyo Caña (6172-III), in Proyecto de Cartografia Geotemática de la República dominicana, Santo
Domingo, Dirección General de Minería, 186.
Hernaiz Huerta, P.P., and Pérez-Estaún, A., 2002: Structure of the Peralta thrust and fold belt,
Dominican Republic. Acta Geologica Hispanica, 37 (2–3), 183–205.
Hernaiz Huerta, P.P., Ardévol Oró, L., Granados, L., Calvo, J.P., Escuder-Viruete, J., Escuer, J., Caballero,
E.L., and Pacheco, C.A., 2004: Mapa geológico de la República Dominicana, escala 1:50.000. La
Descubierta (5871-I). , 156.
Hernaiz Huerta, P.P., Díaz de Neira, J.A., García-Senz, J., Deschamps, I., Genna, A., Nicole, N., Lopera,
E., Escuder Viruete, J., Ardévol Oró, L., and Pérez-Estaún, A., 2007: La estructura del suroeste de
la República Dominicana: un ejemplo de deformación en régimen transpresivo. Boletín Geológico
y Minero, 118 (2), 337–357.
Hernaiz Huerta, P.P., Pérez-Valera, F., Abad, M., Monthel, J., and Diaz de Neira, A., 2012: Mélanges and
olistostromes in the Puerto Plata area (northern Dominican Republic) as a record of subduction
and collisional processes between the Caribbean and North-American plates. Tectonophysics,
568–569, 266–281.
Heubeck, C., and Mann, P., 1991a: Geologic evaluation of plate kinematic models for the North
American-Caribbean plate boundary zone. Tectonophysics, 191, 1–26.
Heubeck, C., Mann, P., Dolan, J., and Monechi, S., 1991: Diachronous uplift and recycling of
sedimentary basins during Cenozoic tectonic transpression, northeastern Caribbean plate
margin. Sedimentary Geology, 70, 1–32.
Heubeck, C., and Mann, P., 1991b: Structural geology and Cenozoic tectonic history of the
southeastern termination of the Cordillera Central, Dominican Republic. Geological Society of
America Special Paper, 262, 315–336.
Hickman, S., and Zoback, M., 2004: Stress orientations and magnitudes in the SAFOD pilot hole.
Geophysical Research Letters, 31 (15), 13–16.
Hilgen, F.J., Lourens, L.J., Van Dam, J.A., Beu, A.G., Boyes, A.F., Cooper, R.A., Krijgsman, W., Ogg, J.G.,
Piller, W.E., and Wilson, D.S., 2012: The Neogene Period, in Gradstein, F.M., Ogg, J.G., Schmitz,
M.D., and Ogg, G.M. eds., The Geologic Time Scale, Boston, Elsevier, 923–978.
Hill, R.I., 1993: Mantle plumes and continental tectonics. Lithos, 30 (3–4), 193–206.

316
References

van Hinsbergen, D.J.J., Iturralde-Vinent, M.A., van Geffen, P.W.G., García-Casco, A., and van Benthem,
S., 2009: Structure of the accretionary prism, and the evolution of the Paleogene northern
Caribbean subduction zone in the region of Camagüey, Cuba. Journal of Structural Geology, 31
(10), 1130–1144.
Hodell, D.A., Curtis, J.H., Jones, G.A., Higuera-Gundy, A., Brenner, M., Binford, M.W., and Dorsey, K.T.,
1991: Reconstruction of Caribbean climate change over the past 10,500 years. Nature, 352
(August), 790–793.
Hoefs, J., 2009: Stable Isotope Geochemistry. 285.
Holcombe, T.L., Vogt, P.R., Matthews, J.E., and Murchison, R.R., 1973: Evidence for sea-floor spreading
in the Cayman Trough. Earth and Planetary Science Letters, 20, 357–371.
Holland, M., and Urai, J.L., 2010: Evolution of anastomosing crack – seal vein networks in limestones:
Insight from an exhumed high-pressure cell, Jabal Shams, Oman Mountains. Journal of Structural
Geology, 32 (9), 1279–1290.
Horan, S.L., 1995: The geochemistry and tectonic significance of the Maimón-Maina schists, Cordillera
Central, Dominican Republic: University of Florida, 172 p.
Hudson, J.D., 1977: Stable istopes and limestone lithofication. Journal of the Geological Society of
London, 133, 637–660.
International Seismological Centre, 2014: On-line Bulletin. http://www.isc.ac.uk,.
Irving, A.J., and Wyllie, P.J., 1975: Subsolidus and melting relationships for calcite, magnesite and the
join CaCO3-MgCO3 to 36 kb. Geochimica et Cosmochimica Acta, 39 (1), 35–53.
Ishii, E., Funaki, H., Tokiwa, T., and Ota, K., 2010: Relationship between fault growth mechanism and
permeability variations with depth of siliceous mudstones in northern Hokkaido, Japan. Journal
of Structural Geology, 32 (11), 1792–1805.
Iturralde-Vinent, M., Millán, G., Korpas, L., Nagy, E., and Pajón, J., 1996: Geological interpretation of
the Cuban K-Ar database, in In: Iturralde-Vinent, M.A. (ed.), Ofiolitas y Arcos Volcánicos de Cuba.
Miami, USA IGCP Project 364 Special Contribution, 48–69.
Jackson, J., and McKenzie, D., 1999: A hectare of fresh striations on the Arkitsa Fault, central Greece.
Journal of Structural Geology, 21, 1–6.
Jackson, T.A., 2013: A review of volcanic island evolution and magma production rate: an example from
a Cenozoic island arc in the Caribbean. Journal of the Geological Society, 170 (3), 547–556.
James, K.H., 2006a: Arguments for and against the Pacific origin of the Caribbean Plate: discussion,
finding for an inter-American origin. Geologica Acta, 4 (1–2), 279–302.
James, K.H., 2006b: The Caribbean Ocean Plateau – an overview, and a different understanding. , 1–
28.
James-Williamson, S. a., Mitchell, S.F., and Ramsook, R., 2014: Tectono-stratigraphic development of
the Coastal Group of south-eastern Jamaica. Journal of South American Earth Sciences, 50, 40–
47.
Jean Poix, C., 1980: Bilan des formations d’Haïti. 1 ère Coll. Géologie d’Haïti, Impr. Le Natal, P-au-P,,
171–183.
Jolly, W.T., Lidiak, E.G., Dickin, A.P., and Wu, T.-W., 2001: Secular Geochemistry of Central Puerto Rican
Island Arc Lavas: Constraints on Mesozoic Tectonism in the Eastern Greater Antilles. Journal of
Petrology, 42 (12), 2197–2214.

317
References

Joubert, M., Ardévol Oró, L., Escuder-Viruete, J., Draper, G., Lewis, J.F.F., Bernárdez Rodríguez, E.,
Ardévol Oró, L., Granados, L., Serra Kiel, J., Joubert, M., Ardévol Oró, L., Escuder-Viruete, J.,
Draper, G., Lewis, J.F.F., et al., 2004: Memoria explicativa del Mapa Geológico a escala 1:50000
de Lamedero (5973-II), in Proyecto de Cartografia Geotemática de la República dominicana, Santo
Domingo, Dirección General de Minería, 192.
Joussineau, G. De, and Aydin, A., 2007: The evolution of the damage zone with fault growth in
sandstone and its multiscale characteristics. Journal of Geophysical Research, 112, 1–19.
Kamenov, G.D., Perfit, M.R., Lewis, J.F., Goss, A.R., Jr., R.A., and Shuster, R.D., 2011a: Ancient
lithosperic source for Quaternary lavas in Hispaniola (Supplementary information). Nature
Geoscience, 4 (8), 11.
Kamenov, G.D., Perfit, M.R., Lewis, J.F., Goss, A.R., Arévalo, R., and Shuster, R.D., 2011b: Ancient
lithospheric source for Quaternary lavas in Hispaniola. Nature Geoscience, 4 (8), 554–557.
Kennedy, L.A., and Logan, J.M., 1997: The role of veining and dissolution in the evolution of fine-
grained mylonites: The McConnell thrust, Alberta. Journal of Structural Geology, 19 (6), 785–797.
Kennedy, L.A., and Logan, J.M., 1998: Microstructures of cataclasites in a limestone-on-shale thrust
fault: implications for low-temperature recrystallization of calcite. Tectonophysics, 295, 167–186.
Kennedy, L.A., and White, J.C., 2001: Low-temperature recrystallization in calcite: Mechanisms and
consequences. Geology, 29 (11), 1027–1030.
Kerr, A.C., Iturralde-Vinent, M.A., Saunders, A.D., Babbs, T.L., and Tarney, J., 1999: A new plate tectonic
model of the Caribbean: Implications from a geochemical reconnaissance of Cuban Mesozoic
volcanic rocks. Geological Society of America Bulletin, 111 (11), 1581–1599.
Kerr, A.C., Tarney, J., Kempton, P.D., Spadea, P., Nivia, A., Marriner, G.F., and Duncan, R.A., 2002:
Pervasive mantle plume head heterogeneity : Evidence from the late Cretaceous Caribbean-
Colombian oceanic plateau. Journal of Geophysical Research: Solid Earth, 107 (B7), 13.
Kerr, A.C., White, R. V, Thompson, P.M.E., and Saunders, A.D., 2003: No Oceanic Plateau — No
Caribbean Plate? The seminal role of an Oceanic Plateau in Caribbean plate evolution. AAPG
Memoir, 79, 126–168.
Kerr, A.C., Pearson, D.G., and Nowell, G.M., 2009: Magma source evolution beneath the Caribbean
oceanic plateau: new insights from elemental and Sr-Nd-Pb-Hf isotopic studies of ODP Leg 165
Site 1001 basalts. Geological Society, London, Special Publications, 328 (1), 809–827.
Kerr, A.C., 2013: Oceanic Plateaus, in Treatise on Geochemistry, 2nd Edition: Geochemistry of the
Earth’s Crust, 1–96.
Kesler, S.E., and Fleck, R.J., 1967: Age and possible origin of a granitic intrusion in the Great Antilles
Island Arc. Geological Society of America Special Paper, 115, 482.
Kesler, S.E., 1971: Petrology of the Terre Neuve igneous province, northern Haiti. Geological Society of
America Memoirs, 130, 119–137.
Kesler, S.E., Sutter, J.F., Jones, L.M., and Walker, R.L., 1977a: Early Cretaceous basement rocks in
Hispaniola. Geology, 5 (4), 245–247.
Kesler, S.E., Lewis, J.F., Jones, L.M., and Walker, R.L., 1977b: Early island-arc intrusive activity, Cordillera
Central, Dominican Republic. Contributions to Mineralogy and Petrology, 65, 91–99.
Kesler, S.E., Russell, N., Reyes, C., Santos, L., Rodriguez, A., and Fondeur, L., 1991a: Geology of the
Maimon Formation, Dominican Republic. Geological Society of America Special Paper, 262, 173–

318
References

185.
Kesler, S.E., Sutter, J.F., Barton, J.M., and Speck, R.C., 1991b: Age of intrusive rocks in northern
Hispaniola. Geological Society of America Special Paper, 262, 165–172.
Kesler, S.E., Campbell, I.H., and Allen, C.M., 2005: Age of the Los Ranchos Formation, Dominican
Republic: Timing and tectonic setting of primitive island arc volcanism in the Caribbean region.
Geological Society of America Bulletin, 117 (7), 987.
Kim, S.-T., and O’Neil, J.R., 1997: Equilibrium and nonequilibrium oxygen isotope effects in synthetic
carbonates. Geochimica et Cosmochimica Acta, 61 (16), 3461–3475.
Kim, Y.-S., and Sanderson, D.J., 2005: The relationship between displacement and length of faults: a
review. Earth-Science Reviews, 68 (3–4), 317–334.
Kim, Y., and Sanderson, D.J., 2010: Inferred fluid flow through fault damage zones based on the
observation of stalactites in carbonate caves. Journal of Structural Geology, 32 (9), 1305–1316.
Kirschner, D.L., and Kennedy, L.A., 2001: Limited syntectonic fluid flow in carbonate-hosted thrust
faults of the Front Ranges, Canadian Rockies, inferred from stable isotope data and structures.
Journal of Geophysical Research, 106 (B5), 8827–8840.
Koehn, D., and Passchier, C.W., 2000: Shear sense indicators in striped bedding-veins. Journal of
Structural Geology, 22, 1141–1151.
Kolbe, A.R., Hutson, R.A., Shannon, H., Trzcinski, E., Miles, B., Levitz, N., Puccio, M., James, L., Noel,
J.R., and Muggah, R., 2010: Mortality, crime and access to basic needs before and after the Haiti
earthquake: A random survey of Port-au-Prince households. Medicine, Conflict and Survival, 26
(4), 281–297.
Krebs, M., Maresch, W.V., Schertl, H.-P., Münker, C., Baumann, a., Draper, G., Idleman, B., and Trapp,
E., 2008: The dynamics of intra-oceanic subduction zones: A direct comparison between fossil
petrological evidence (Rio San Juan Complex, Dominican Republic) and numerical simulation.
Lithos, 103 (1–2), 106–137.
Krebs, M., Schertl, H.-P., Maresch, W.V., and Draper, G., 2011: Mass flow in serpentinite-hosted
subduction channels: P–T–t path patterns of metamorphic blocks in the Rio San Juan mélange
(Dominican Republic). Journal of Asian Earth Sciences, 42 (4), 569–595.
Lane, C.S., Horn, S.P., Orvis, K.H., and Thomason, J.M., 2011: Oxygen isotope evidence of Little Ice Age
aridity on the Caribbean slope of the Cordillera Central, Dominican Republic. Quaternary
Research, 75 (3), 461–470.
Lapierre, H., Dupuis, V., Lépinay, B.M. De, Tardy, M., Ruíz, J., Maury, R.C.R.C.R.C., Hernandez, J., Loubet,
M., Lepinay, B.M. De, Ruiz, J., and Maury, R.C.R.C.R.C., 1997: Is the Lower Duarte Igneous
Complex (Hispaniola) a remnant of the Caribbean plume-generated Oceanic Plateau? The Journal
of Geology, 105 (1), 111–120.
Lapierre, H., Dupuis, V., de Lépinay, B.M., Bosch, D., Monié, P., Tardy, M., Maury, R.C., Hernandez, J.,
Polvé, M., Yeghicheyan, D., Cotten, J., Lewis, J.F., Hames, W.E., and Draper, G., 1999: Late Jurassic
Oceanic Crust and Upper Cretaceous Caribbean Plateau Picritic Basalts Exposed in the Duarte
Igneous Complex, Hispaniola. The Journal of Geology, 107 (4), 193–207.
Lapierre, H., Bosch, D., Dupuis, V., Polvé, M., Maury, R.C., Hernandez, J., Monié, P., Yeghicheyan, D.,
Jaillard, E., Tardy, M., Mercier de Lépinay, B., Mamberti, M., Desmet, A., Keller, F., et al., 2000:
Multiple plume events in the genesis of the peri-Caribbean Cretaceous oceanic plateau province.
Journal of Geophysical Research, 105 (B4), 8403–8421.

319
References

Laubach, S.E., and Lorenz, J.C., 1992: Preliminary assessment of natural fracture patterns in Frontier
Formation sandstones, sotuhwest Wyoming, in Mullen, C.E. ed., Wyoming Geological
Association, 43rd Annual Field Conference, Guidebook, 87–96.
Lebrón, M.C., and Perfit, M.R., 1994: Petrochemistry and tectonic significance of Cretaceous island-arc
rocks, Cordillera Oriental, Dominican Republic. Tectonophysics, 229, 69–100.
Lee, Y.-J., and Wiltschko, D. V, 2000: Fault controlled sequential vein dilation: competition between
slip and precipitation rates in the Austin Chalk, Texas. Journal of Structural Geology, 22, 1247–
1260.
Lena, G., Barchi, M.R., Alvarez, W., Felici, F., and Minelli, G., 2015: Mesostructural analysis of S-C fabrics
in a shallow shear zone of the Umbria – Marche Apennines (Central Italy), in Faulkner, D.R.,
Mariani, E., and Macklenburgh, J. eds., Rock Deformation from Field, Experiments and Theory: A
Volume in Honour of Ernie Rutter, Geological Society, London, Special Publications, 328, 149–166.
Leroy, S., 1995: Structure et origine de la Plaque Caraïbe: Implications géodynamiques: Université
Pierre et Marie Curie - Paris 6, 230 p.
Leroy, S., Mercier de Lépinay, B., Mauffret, A., and Pubellier, M., 1996: Structural and Tectonic
Evolution of the Eastern Cayman Trough (Caribbean Sea) From Seismic Reflection Data. AAPG
Bulletin, 80 (2), 222–247.
Leroy, S., and Mauffret, A., 1996: Intraplate deformation in the Caribbean region. Journal of
Geodynamics, 21 (1), 113–122.
Leroy, S., Mauffret, A., Patriat, P., and Mercier de Lépinay, B., 2000: An alternative interpretation of
the Cayman trough evolution from a reidentification of magnetic anomalies. Geophysical Journal
International, 141 (3), 539–557.
Leroy, S., Ellouz-Zimmermann, N., Corbeau, J., Rolandone, F., Mercier de Lépinay, B., Meyer, B.,
Momplaisir, R., Granja Bruña, J.L., Battani, A., Baurion, C., Burov, E., Clouard, V., Deschamps, R.,
Gorini, C., et al., 2015: Segmentation and kinematics of the North America-Caribbean plate
boundary offshore Hispaniola. Terra Nova, 27 (6), 467–478.
Lewis, J.F., and Draper, G., 1990: Geology and tectonic evolution of the northern Caribbean margin, in
Dengo, G. and Case, J.E. eds., The geology of North America, Boulder, Geological Society of
America, 77–140.
Lewis, J.F., Amarante, A., Bloise, G., Jiménez, J.G., and Dominguez, H.D., 1991: Lithology and
stratigraphy of Upper Cretaceous volcanic and volcaniclastic rocks of the Tireo Group, Dominican
Republic, and correlations with the Massif du Nord in Haiti. Geological Society of America Special
Paper, 262, 143–163.
Lewis, J.F., and Jimenéz, J.G., 1991: Duarte Complex in the La Vega-Jarabacoa-Janico area, central
Hispaniola; geologic and geochemical features of the sea floor during the early stages of arc
evolution. Geological Society of America Special Paper, 262, 115–141.
Lewis, J.F., and Draper, G., 1995: Amphibolites and associated rocks of the Rio Verde Complex in the
median belt, Central Hispaniola: Their petrologic, structural and tectonic significance in the
placement of the Loma Caribe Peridotite, in 3rd Conference of the Geological Society and Trinidad
& Tobago, Port of Spain, 187.
Lewis, J.F., Escuder-Viruete, J., Hernaiz Huerta, P.P., Gutierrez, G., Draper, G., and Pérez-Estaún, A.,
2002: Geochemical subdivision of the Circum-Caribbean Island Arc, Dominican Cordillera Central;
implications for the crustal formation, accretion and growth within an intra-oceanic setting. Acta
Geologica Hispanica, 37 (2–3), 81–122.

320
References

Lewis, J.F., Proenza, J.A., Melgarejo, J.C., and Gervilla, F., 2003: The puzzle of Loma Caribe Chromitites
(Hispaniola), in Annual Report of IGCP Project No. 433, Reports of the field workshop in Cuba, 1.
Lewis, J.F., Proenza, J.A., and Longo, F., 2005: New petrological and geochemical constraints on the
origin of Loma Caribe Peridotite (Dominican Republic), in 17th Caribbean Geological Conference,
San Juan, 47–48.
Lewis, J.F., Draper, G., Proenza, J.A., Espaillat, J., Cristalografia, D. De, Geologia, F. De, and Dominicana,
F., 2006: Ophiolite-related ultramafic rocks (serpentinites) in the Caribbean region: A review of
their occurrence , composition, origin, emplacement and Ni-Laterite soil formation. Geologica
Acta, 4 (1–2), 237–263.
Lewis, J.F., Mattietti, G.K., Perfit, M., Kamenov, G., and Science, E., 2011: Geochemistry and petrology
of three granitoid rock cores from the Nicaraguan Rise, Caribbean Sea: implications for its
composition, structure and tectonic evolution. Geologica Acta, 9 (3–4), 467–479.
Lidiak, E.G., and Jolly, W.T., 1996: Circum-Caribbean Granitoids: Characteristics and Origin.
International Geology Review, 38 (12), 1098–1133.
Lister, G.S., and Snoke, A.W., 1984: S-C Mylonites. Journal of Structural Geology, 6 (6), 617–638.
Liu, J., Walter, J.M., and Weber, K., 2002: Fluid-enhanced low-temperature plasticity of calcite marble:
Microstructures and mechanisms. Geology, 30, 787–790.
Llanes Estrada, P., Ten Brink, U.S., Granja Bruña, J.L., Carbó-Gorosabel, A., Flores, C.H., Villasenor, A.,
Pazos, A., and Martin Davila, J.M., 2012: True subduction vs. underthrusting of the Caribbean
plate beneath Hispaniola, Northern Caribbean, in American Geophysical Union, Fall Meeting
2012, abstract #T41A-2567, 1.
Logan, J.M., Friedman, M., Higgs, N., Dengo, C., and Shimamoto, T., 1979: Experimental studies of
simulated gouge and their application to studies of natural fault zones. U.S. Geological survey
Open File Report, 79 (1239), 305–343.
Lohmann, K.C., 1988: Geochemical Patterns of Meteoric Diagenetic Systems and Their Application to
Studies of Paleokarst, in James, N.P. ed., Paleokarst, New York, Springer-Verlag, 58–80.
Lunn, R.J., Willson, J.P., Shipton, Z.K., and Moir, H., 2008: Simulating brittle fault growth from linkage
of preexisting structures. Journal of Geophysical Research, 113, 1–10.
Machel, H.G., 2000: Application of Cathodoluminescence to Carbonate Diagenesis, in Pagel, M. ed.,
Cathodoluminescence in Gesciences, Springer, 31.
Main, I.G., Kwon, O., Ngwenya, B., and Elphick, S.C., 2000: Fault sealing during deformation-band
growth in porous sandstone. Geology, 28 (12), 1131–1134.
Manaker, D.M., Calais, E., Freed, a. M., Ali, S.T., Przybylski, P., Mattioli, G., Jansma, P., Prépetit, C., and
de Chabalier, J.B., 2008: Interseismic Plate coupling and strain partitioning in the Northeastern
Caribbean. Geophysical Journal International, 174 (3), 889–903.
Mann, P., Hempton, M.R., Bradley, D.C., and Burke, K., 1983: Development of Pull-Apart Basins. The
Journal of Geology, 91 (5), 529–554.
Mann, P., Burke, K., and Matumoto, T., 1984: Neotectonics of Hispaniola: plate motion, sedimentation,
and seismicity at a restraining bend. Earth and Planetary Science Letters, 70 (2), 311–324.
Mann, P., and Burke, K., 1984: Neotectonics of the Caribbean. Reviews of geophysics and space physics,
22 (4), 309–362.
Mann, P., Draper, G., and Burke, K., 1985: Neotectonics of a strike-slip restraining bend system,

321
References

Jamaica. The Society of Economic Paleontologists and Mineralogists,, 211–226.


Mann, P., and Burke, K., 1990: Transverse intra-arc rifting: Palaeogene Wagwater Belt, Jamaica. Marine
and Petroleum Geology, 7, 410–427.
Mann, P., Tyburski, S.A., and Rosencrantz, E., 1991a: Neogene development of the Swan Islands
restraining-bend complex, Caribbean Sea. Geology, 19, 823–826.
Mann, P., Draper, G., and Lewis, J.F., 1991b: An overview of the geologic and tectonic development of
Hispaniola. Geological Society of America Special Paper, 262, 1–27.
Mann, P., McLaughlin, P.P., and Cooper, C., 1991c: Geology of the Azua and Enriquillo basins,
Dominican Republic; 2, Structure and tectonics. Geological Society of America Special Paper, 262,
367–389.
Mann, P., Taylor, F.W., Edwards, R.L., and Ku, T.-L., 1995: Actively evolving microplate formation by
oblique collision and sideways motion along strike-slip faults: An example from the northeastern
Caribbean plate margin. Tectonophysics, 246 (1–3), 1–69.
Mann, P., Prentice, C.S., Burr, G., Peña, L.R., and Taylor, F.W., 1998: Tectonic geomorphology and
paleoseismology of the Septentrional fault system, Dominican Republic, in Dolan, J.F. and Mann,
P. eds., Active Strike-Slip and Collisional Tectonics of the Northern Caribbean Plate Boundary
Zone, Boulder, Colorado, Geological Society of America Special Papers, 63–123.
Mann, P., Calais, E., Ruegg, J.-C., DeMets, C., Jansma, P.E., and Mattioli, G.S., 2002: Oblique collision in
the northeastern Caribbean from GPS measurements and geological observations. Tectonics, 21
(6), 1–26.
Mann, P., Rogers, R.D., and Gahagan, L., 2007a: Overview of plate tectonic history and its unresolved
tectonic problems, in Buncdschud, J. ed., Central America: Geology, Resources and Natural
Hazards, Balkema Publishers, 205–241.
Mann, P., DeMets, C., and Wiggins-Grandison, M., 2007b: Toward a better understanding of the Late
Neogene strike-slip restraining bend in Jamaica: geodetic, geological, and seismic constraints, in
Cunningham, W.D. and Mann, P. eds., Tectonics of Strike-Slip Restraining and Releasing Bends,
Geological Society, London, Special Publications, 239–253.
Mann, P., 2007: Overview of the tectonic history of northern Central America, in Mann, P. ed., Geologic
and Tectonic Development of the Caribbean plate boundary in northern Central America,
Geological Society of America Special Papers, 1–9.
Marchesi, C., Garrido, C.J., Godard, M., Proenza, J. a., Gervilla, F., and Blanco-Moreno, J., 2006:
Petrogenesis of highly depleted peridotites and gabbroic rocks from the Mayarí-Baracoa
Ophiolitic Belt (eastern Cuba). Contributions to Mineralogy and Petrology, 151 (6), 717–736.
Marone, C., and Scholz, C.H., 1989: Particle-size distribution and microstructures within simulated fault
gouge. Journal of Structural Geology, 11 (7), 799–814.
Marshak, S., and Engelder, T., 1985: Development of cleavage in limestones of a fold-thrust belt in
eastern New York. Journal of Structural Geology, 7 (3/4), 345–359.
Matti, J.C., and Morton, D.M., 1993: California: A reconstruction based on a new cross-fault
correlation, in Powell, R.E. ed., The San Andreas fault system: Displacement, palinspastic
reconstruction, and geologic evolution, Geological Society of America Memoir 178, 107–159.
Mattinson, J.M., Pessango, E.A., Montgomery, H., and Hopson, C.A., 2008: Late Jurassic age of oceanic
basement at La Désirade Island, Lesser Antilles arc. Geological Society of America Special Paper,
438, 175–190.

322
References

Mattioli, G., Miller, J., Demets, C., and Jansma, P., 2014: Rigidity and definition of Caribbean plate
motion from COCONet and campaign GPS observations. Geophysical Research Abstracts, 16
(EGU2014-EGU14546).
Mauffret, A., and Leroy, S., 1997: Seismic stratigraphy and structure of the Caribbean igneous province.
Tectonophysics, 283 (1–4), 61–104.
Mauffret, A., and Leroy, S., 1999: Neogene Intraplate Deformation of the Caribbean Plate at the Beata
Ridge, in Mann, P. ed., Sedimentary Basins of the World; Caribbean Basins, Amsterdam, Elsevier
Science B.V., 627–669.
Mauffret, A., Leroy, S., Vila, J., Hallot, E., L, B.M. De, and Duncan, R.A., 2001: Prolonged Magmatic and
Tectonic Development of the Caribbean Igneous Province Revealed by a Diving Submersible
Survey. Marine Geophysical Researches, 22, 17–45.
Maurrasse, F.J., Sayeed, U.A., Georges, G., Pierre Louis, F., and Rigaud, J., 1977: Ophiolite complex of
the southern peninsula of Haïti: A view at the Caribbean crust. 8th Caribbean Geological
Conference Abstracts, 9, 115–116.
Maurrasse, F., Husler, J., Georges, G., Schmitt, R., and Damond, P., 1979a: Upraised Caribbean seafloor
below acoustic reflector B" at the Southern Peninsula of Haiti. Geologie en Mijnbouw, 58 (1), 71–
83.
Maurrasse, F., Pierre Louis, F., and Rigaud, J.G., 1979b: Upper Cretaceous to lower Paleocene pelagic
calcareous deposits in the Southern Peninsula of Haïti: Their bearing on the problem of the
Cretaceous - Tertiary boundary. 4th Latin American Geological Congress, Trinidad and Tobago,,
328–337.
Maurrasse, F., 1980: New data on the stratigraphy of the Southern Peninsula of Haïti. 1 ère Coll.
Géologie d’Haïti, Impr. Le Natal, P-au-P,, 184–198.
Maurrasse, F., Pierre-Louis, F., and Rigaud, J.-G., 1980: Cenozoic facies distribution in the southern
peninsula of Haiti, in Transactions of the 9th Caribbean Geological Conference, Santo Domingo,
15.
Maurrasse, F., 1982: Survey of the geology of Haïti: Guide to the field excursions in Haïti. Miami
Geological Society,, 103.
Maurrasse, J.-M.R.F., and Sen, G., 1991: Impacts, Tsunamis, and the Haitian Cretaceous-Tertiary
Boundary Layer. Science, 252 (5013), 1690–1693.
Maus, S., 2010: An ellipsoidal harmonic representation of Earth’s lithospheric magnetic field to degree
and order 720. Geochemistry, Geophysics, Geosystems, 11 (6), 1–12.
McCann, W.R., 2006: Estimating the threat of tsunamigenic earthquakes and earthquake induced-
landslide tsunami in the Caribbean, in Aurelio, M. and Philip, L. eds., Caribbean Tsunami Hazard,
World Scientific Publishing, Singapore, 43–65.
McLaughlin, P.P., and Sen Gupta, B.K., 1991: Migration of Neogene marine environments,
southwestern Dominican Republic. Geology, 19 (3), 222–225.
Meighan, H.E., and Pulliam, J., 2013: Seismic anisotropy beneath the northeastern Caribbean:
implications for the subducting North American lithosphere. Bulletin de la Societe Géologique de
France, 184 (1–2), 67–76.
Meneghini, F., Botti, F., Aldega, L., Boschi, C., Corrado, S., Marroni, M., and Pandolfi, L., 2012: Hot fluid
pumping along shallow-level collisional thrusts: The Monte Rentella Shear Zone, Umbria
Apennine, Italy. Journal of Structural Geology, 37, 36–52.

323
References

Mercier de Lépinay, B., Labesse, B., Sigal, J., and Vila, J.M., 1979: Sedimentation chaotique et
tectonique tangentielle Maestrichtienne dans la Presqu’île du Sud d’Haïti (île d’Hispaniola,
Grandes Antilles). C.R. Ac. Sc. Paris, 289 (D), 887–890.
Mercier de Lépinay, B., Deschamps, A., Klingelhoefer, F., Mazabraud, Y., Delouis, B., Clouard, V., Hello,
Y., Crozon, J., Marcaillou, B., Graindorge, D., Vallée, M., Perrot, J., Bouin, M.-P., Saurel, J.-M., et
al., 2011: The 2010 Haiti earthquake: A complex fault pattern constrained by seismologic and
tectonic observations. Geophysical Research Letters, 38 (L22305), 1–7.
Meschede, M., and Frisch, W., 1998: A plate-tectonic model for the Mesozoic and Early Cenozoic
history of the Caribbean plate. Tectonophysics, 296 (3–4), 269–291.
Metcalf, R.V., and Shervais, J.W., 2008: Suprasubduction-zone ophiolites: Is there really an ophiolite
conundrum? Geological Society of America Special Paper, 438 (7), 191–222.
Micarelli, L., Moretti, I., and Daniel, J.M., 2003: Structural properties of rift-related normal faults: the
case study of the Gulf of Corinth, Greece. Journal of Geodynamics, 36, 275–303.
Micarelli, L., Benedicto, A., and Wibberley, C.A.J., 2006a: Structural evolution and permeability of
normal fault zones in highly porous carbonate rocks. Journal of Structural Geology, 28, 1214–
1227.
Micarelli, L., Moretti, I., Jaubert, M., and Moulouel, H., 2006b: Fracture analysis in the south-western
Corinth rift (Greece) and implications on fault hydraulic behavior. Tectonophysics, 426, 31–59.
Michard, A., Albarède, F., Michard, G., Minster, J.F., and Charlou, J.L., 1983: Rare-earth elements and
uranium in high-temperature solutions from east pacific rise hydrothermal vent field (13 °N).
Nature, 303, 795–797.
Michel, G., 1986: Caracterisation de l’indice aurifere de mont-organise (Haïti). Secteur Maman Noel et
Grenier. Mémoire de Maîtrise Sc. Appliquées. Ecole Polytechnique de Montréal,, 239.
Mitchell, S.F., 2003: Sedimentology and tectonic evolution of the Cretaceous rocks of Central Jamaica:
Relationships to the plate tectonic evolution of the Caribbean. AAPG Memoir, 79, 605–623.
Mitchell, T.M., and Faulkner, D.R., 2008: Experimental measurements of permeability evolution during
triaxial compression of initially intact crystalline rocks and implications for fluid flow in fault
zones. Journal of Geophysical Research: Solid Earth, 113 (11), 1–16.
Mitchell, T.M., and Faulkner, D.R., 2009: The nature and origin of off-fault damage surrounding strike-
slip fault zones with a wide range of displacements: A field study from the Atacama fault system
, northern Chile. Journal of Structural Geology, 31 (8), 802–816.
Mizoguchi, K., Hirose, T., Shimamoto, T., and Fukuyama, E., 2008: Internal structure and permeability
of the Nojima fault, southwest Japan. Journal of Structural Geology, 30, 513–524.
Moir, H., Lunn, R.J., Shipton, Z.K., and Kirkpatrick, J.D., 2010: Simulating brittle fault evolution from
networks of pre-existing joints within crystalline rock. Journal of Structural Geology, 32 (11),
1742–1753.
Molli, G., Clancy, J., Kennedy, L., and Taini, V., 2011: Low-temperature deformation of limestone , Isola
Palmaria , northern Apennine, Italy - The role of primary textures, precursory veins and
intracrystalline deformation in localization. Journal of Structural Geology, 33 (3), 255–270.
Molnar, P., and Sykes, L.R., 1969: Tectonics of the Caribbean and Middle America regions from focal
mechanisms and seismicity. Geological Society of America Bulletin, 80, 1639–1684.
Mondziel, S., Grindlay, N., Mann, P., Escalona, A., and Abrams, L., 2010: Morphology, structure, and

324
References

tectonic evolution of the Mona canyon (northern Mona passage) from multibeam bathymetry,
side-scan sonar, and seismic reflection profiles. Tectonics, 29 (2), 1–23.
Montgomery, H., Pessagno, E.A., Lewis, J.F., and Schellekens, J., 1994a: Paleogeography of Jurassic
fragments in the Caribbean. Tectonics, 13 (2), 725–732.
Montgomery, H., Pessango, E.A., and Pindell, J.L., 1994b: A 195 Ma terrane in a 165 Ma sea: Pacific
Origin of the Caribbean Plate. Geological Society of America Today, 4 (1), 1–6.
Montgomery, H., and Pessagno, E.A., 1999: Cretaceous microfaunas of the Blue Mountains, Jamaica,
and of the northern and central basement complexes of Hispaniola, in Mann, P. ed., Sedimentary
Basins of the World; Caribbean Basins, Amsterdam, Elsevier Science B.V., 237–246.
Montgomery, H., and Kerr, A.C., 2009: Rethinking the origins of the red chert at La Desirade, French
West Indies. Geological Society, London, Special Publications, 328 (1), 457–467.
Morad, S., Al-aasm, I.S., Sirat, M., and Sattar, M.M., 2010: Vein calcite in cretaceous carbonate
reservoirs of Abu Dhabi: Record of origin of fluids and diagenetic conditions. Journal of
Geochemical Exploration, 106 (1–3), 156–170.
Moreno, B., Grandison, M., and Atakan, K., 2002: Crustal velocity model along the southern Cuban
margin: implications for the tectonic regima at an active plate boundary. Geophysical Journal
International, 151, 632–645.
Morrow, C.A., and Byerlee, J.D., 1992: Permeability of Core Samples From Cajon Pass Scientific Drill
Hole: Results From 2100 to 3500 m Depth. Journal of Geophysical Research: Solid Earth, 97 (B4),
5145–5151.
Muchez, P., Slobodnik, M., Viaene, W.A., and Keppens, E., 1995: Geochemical constraints on the origin
and migration of palaeofluids at the northern margin of the Variscan foreland, southern Belgium.
Sedimentary Geology, 96, 191–200.
Mueller, K., 2017: Variation in slip rates on active faults: Natural growth or stress transients? Geology,
45 (3), 287–288.
Müller, R.D., Royer, J.-Y., Cande, S.C., Roest, W.R., and Maschenkov, S., 1999: New constraints on the
Late Cretaceous/Tertiary plate tectonic evolution of the Caribbean, in Mann, P. ed., Sedimentary
Basins of the World; Caribbean Basins, Amsterdam, Elsevier Science B.V., 33–59.
Munz, I.A., Yardley, B.W.D., Banks, D.A., and Wayne, D., 1995: Deep penetration of sedimentary fluids
in basement rocks from southern Norway: Evidence from hydrocarbon and brine inclusions in
quartz veins. Geochimica et Cosmochimica Acta, 59 (2), 239–254.
Myczynski, R., and Iturralde-Vinent, M., 2005: The Late Lower Albian Invertebrate Fauna of the Río
Hatillo Formation of Pueblo Viejo, Dominican Republic. Caribbean Journal of Science, 41 (4), 782–
796.
Nakatani, M., and Scholz, C.H., 2004a: Frictional healing of quartz gouge under hydrothermal
conditions: 1. Experimental evidence for solution transfer healing mechanism. Journal of
Geophysical Research, 109, 1–19.
Nakatani, M., and Scholz, C.H., 2004b: Frictional healing of quartz gouge under hydrothermal
conditions: 2. Quantitative interpretation with a physical model. Journal of Geophysical Research,
109, 1–10.
Narr, W., and Burruss, R.C., 1984: Origin of Reservoir Fractures in Little Knife Field, North Dakota. AAPG
Bulletin, 68 (9), 1087–1100.

325
References

Neill, I., Gibbs, J. a., Hastie, A.R., and Kerr, A.C., 2010: Origin of the volcanic complexes of La Désirade,
Lesser Antilles: Implications for tectonic reconstruction of the Late Jurassic to Cretaceous Pacific-
proto Caribbean margin. Lithos, 120 (3–4), 407–420.
Neill, I., Kerr, A.C., Hastie, A.R., Stanek, K.-P., and Millar, 2011: Origin of the Aves Ridge and Dutch–
Venezuelan Antilles: interaction of the Cretaceous “Great Arc” and Caribbean–Colombian
Oceanic Plateau ? Journal of the Geological Society, LondonLondon, 168, 333–347.
Neill, I., 2011: The tectono-magmatic evolution of the south-eastern Caribbean plate: Insights from La
Désirade, Trinidad and Tobago and the Aves Ridge: University of Cardiff, 401 p.
Nelson, C.E., Proenza, J.A., Lewis, J.F., and López-Kramer, J., 2011: The metallogenic evolution of the
Greater Antilles. Geologica Acta, 9 (3–4), 229–264.
Nickelsen, R.P., and Van Ness, D.H., 1967: Jointing in the Appalachian Plateau of Pennsylvania.
Geological Society of America Bulletin, 78, 609–630.
Nicolini, P., 1977: Les porphyres cupriferes et les complexes ultrabasiques du ne d’Haïti. Essai de
gitologie previsionnelle: 208 p.
Nishikawa, O., and Takeshita, T., 1999: Dynamic analysis and two types of kink bands in quartz veins
deformed under subgreenschist conditions. Tectonophysics, 301 (1–2), 21–34.
Nuriel, P., Rosenbaum, G., Uysal, T.I., Zhao, J., Golding, S.D., Weinberger, R., Karabacak, V., and Avni,
Y., 2011: Formation of fault-related calcite precipitates and their implications for dating fault
activity in the East Anatolian and Dead Sea fault zones. Geological Society, London, Special
Publications, 359 (1), 229–248.
Nuriel, P., Weinberger, R., Rosenbaum, G., Golding, S.D., Zhao, J.X., Tonguc Uysal, I., Bar-Matthews,
M., and Gross, M.R., 2012: Timing and mechanism of late-Pleistocene calcite vein formation
across the Dead Sea Fault Zone, northern Israel. Journal of Structural Geology, 36, 43–54.
Oda, M., Takemura, T., and Aoki, T., 2002: Damage growth and permeability change in triaxial
compression tests of Inada granite. Mechanics of Materials, 34 (6), 313–331.
Ogg, J.G., Hinnov, L.A., and Huang, C., 2012: Cretaceous, in Gradstein, F.M., Ogg, J.G., Schmitz, J., and
Ogg, G.M. eds., The Geologic time Scale, Boston, Elsevier, 793–853.
Oliver, N.H.S., and Bons, P.D., 2001: Mechanisms of fluid flow and fluid – rock interaction in fossil
metamorphic hydrothermal systems inferred from vein – wallrock patterns, geometry and
microstructure. Geofluids, 1, 137–162.
Olsen, M.P., Scholz, H., and Léger, A., 1998: Healing and sealing of a simulated fault gouge under
hydrothermal conditions: Implications for fault healing rubes. Journal of Geophysical Research,
103 (B4), 7421–7430.
Olsson, R.K., Hemleben, C., Berggren, W.A., and Huber, B.T., 1999: Atlas of Paleocene Planktonic
Foraminifera. Washington D.C., Smithsonian Institution Press, 1-252.
Ott, B., and Mann, P., 2012: Early origins of the Caribbean plate from deep seismic profiles across the
Nicaraguan Rise, in American Geophysical Union, Fall Meeting 2012, abstract #T41A-2575,
Pittsburg, 1.
Palmer, H.C., 1979: Geology of the Moncion-Jarabacoa area, Dominican Republic, in Lidz, B. and Nagle,
F. eds., Tectonic Focal Point of the North Caribbean, Miami, Miami Geological Society, 29–68.
Parker, J.M., 1942: Regional systematic jointing in slightly deformed sedimentary rocks. Bulletin of the
Geological Society of America, 53, 381–408.

326
References

Parlangeau, C., Lacombe, O., Schueller, S., and Daniel, J., 2018: Inversion of calcite twin data for
paleostress orientations and magnitudes: A new technique tested and calibrated on numerically-
generated and natural data. Tectonophysics, 722, 462–485.
Passchier, C.W., and Trouw, R.A.J., 2005: Microtectonics. 366.
Peacock, D.C.P., and Sanderson, D.J., 1991: Displacements, segment linkage and relay ramps in normal
fault zones. Journal of Structural Geology, 13 (6), 721–733.
Peacock, D.C.P., and Sanderson, D.J., 1992: Effects of layering and anisotropy on fault geometry.
Journal of the Geological Society, 149 (5), 793–802.
Peacock, D.C.P., and Sanderson, D.J., 1994: Strain and scaling of faults in the chalk at Flamborough
Head, U.K. Journal of Structural Geology, 16 (1), 97–107.
Pentecost, A., 2005: Travertine. Berlin, Springer, 445.
Pindell, J.L., and Dewey, J.F., 1982: Permo-Triassic reconstruction of western Pangea and the evolution
of the Gulf of Mexico/Caribbean region. Tectonics, 1 (2), 179–211.
Pindell, J.L., and Barrett, S.F., 1990: Geological evolution of the Caribbean region: A plate tectonic
perspective, in Dengo, G. and Case, J.E. eds., The Geology of North America - Vol. H, The
Caribbean Region, Boulder, Geological Society of America, 405–433.
Pindell, J., and Kennan, L., 2001a: Kinematic Evolution of the Gulf of Mexico, in Petroleum Systems of
Deep-Water Basins, GCSSEPM Foundation 21st Annual Bob F. Perkins Research Conference
Transactions, 193–220.
Pindell, J., and Kennan, L., 2001b: Processes and Events in the Terrane Assembly of Trinidad and
Eastern Venezuela, in Petroleum Systems of Deep-Water Basins, GCSSEPM Foundation 21st
Annual Bob F. Perkins Research Conference Transactions, 159–192.
Pindell, J., Kennan, L., Maresch, W. V, Stanek, K.-P., Draper, G., and Higgs, R., 2005: Plate-kinematics
and crustal dynamics of circum-Caribbean arc-continent interactions: Tectonic controls on basin
development in Proto-Caribbean margins, in Avé Lallemant, H.G. and Sisson, V.B. eds., Caribbean-
South American plate interactions, Venezuela, Geological Society of America Special Papers, 7–
52.
Pindell, J., Kennan, L., Stanek, K.P., Maresch, W.V., and Draper, G., 2006: Foundations of Gulf of Mexico
and Caribbean evolution: eight controversies resolved. Geologica Acta, 4 (1–2), 303–341.
Pindell, J.L., and Kennan, L., 2009: Tectonic evolution of the Gulf of Mexico, Caribbean and northern
South America in the mantle reference frame: an update, in James, K.H., Lorente, M.A., and
Pindell, J.L. eds., The Origin and Evolution of the Caribbean Plate, Geological Society, London,
Special Publications, 328, 1–55.
Pindell, J., Maresch, W. V., Martens, U., and Stanek, K., 2012: The Greater Antillean Arc: Early
Cretaceous origin and proposed relationship to Central American subduction mélanges:
implications for models of Caribbean evolution. International Geology Review, 54 (2), 131–143.
Piper, D.Z., and Bau, M., 2013: Normalized Rare Earth Elements in Water, Sediments, and Wine:
Identifying Sources and Environmental Redox Conditions. American Journal of Analytical
Chemistry, 4 (10), 69–83.
Pollard, D.D., and Segall, P., 1987: Theoretical displacement and stresses near fractures in rock: with
applications to faults, joints, veins, dikes, and solution surfaces, in Atkinson, B.K. ed., Fracture
Mechanisms of Rock, Academic Press, London, 277–349.

327
References

Pollard, D.D., and Aydin, A., 1988: Progress in understanding jointing over the past century. Geological
Society of America Bulletin, 100 (8), 1181–1204.
Poros, Z., Mindszenty, A., Molnár, F., Pironon, J., Gyori, O., Ronchi, P., and Szekeres, Z., 2012: Imprints
of hydrocarbon-bearing basinal fluids on a karst system: Mineralogical and fluid inclusion studies
from the Buda Hills, Hungary. International Journal of Earth Sciences, 101 (2), 429–452.
Prentice, C.S., Mann, P., Taylor, F.W., Burr, G., and Valastro, S., 1993: Paleoseismicity of the North
American-Caribbean plate boundary (Septentrional fault), Dominican Republic. Geology, 21, 49–
52.
Prentice, C.S., Mann, P., Peña, L.R., and Burr, G., 2003: Slip rate and earthquake recurrence along the
central Septentrional fault, North American–Caribbean plate boundary, Dominican Republic.
Journal of Geophysical Research, 108 (B3), 1–17.
Prentice, C.S., Mann, P., Crone, a. J., Gold, R.D., Hudnut, K.W., Briggs, R.W., Koehler, R.D., and Jean,
P., 2010: Seismic hazard of the Enriquillo–Plantain Garden fault in Haiti inferred from
palaeoseismology. Nature Geoscience, 3 (11), 789–793.
Prentice, C.S., Mann, P., and Peña, L.R., 2013: Comment on “Historical perspective on seismic hazard
to Hispaniola and the northeast Caribbean region” by U. ten Brink et al. Journal of Geophysical
Research: Solid Earth, 118, 1–4.
Proenza, J., Gervilla, F., Melgarejo, J., and Bodinier, J.-L., 1999: Al- and Cr-rich chromitites from the
Mayari-Baracoa ophiolitic belt (eastern Cuba); consequence of interaction between volatile-rich
melts and peridotites in suprasubduction mantle. Economic Geology, 94 (4), 547–566.
Proenza, J.A., Zaccarini, F., Lewis, J.F., Longo, F., and Garuti, G., 2007: Chromian spinel composition and
the platinum-group minerals of the PGE-rich Loma Peguera chromitites, Loma Caribe Peridotite,
Dominican Republic. The Canadian Mineralogist, 45 (3), 631–648.
Prokoph, A., Shields, G.A., and Veizer, J., 2008: Compilation and time-series analysis of a marine
carbonate δ18O, δ13C, 87Sr/86Sr and δ34S database through Earth history. Earth-Science
Reviews, 87, 113–133.
Pubellier, M., Vila, J.-M., and Boisson, D., 1991: North Caribbean neotectonic events: The Trans-Haitian
fault system. Tertiary record of an oblique transcurrent shear zone uplifted in Hispaniola.
Tectonophysics, 194 (3), 217–236.
Pubellier, M., Mauffret, A., Leroy, S., Vila, J.-M., and Amilcar, H., 2000: Plate boundary readjustment in
oblique convergence: Example of the Neogene of Hispaniola, Greater Antilles. Tectonics, 19 (4),
630–648.
Ramsay, J.G., 1980: The crack-seal mechanisms of rock deformation. Nature, 284, 135–135.
Ratschbacher, L., Franz, L., Min, M., Bachmann, R., Martens, U., Stanek, K., Stübner, K., Nelson, B.K.,
Herrmann, U., Weber, B., López-Martínez, M., Jonckheere, R., Sperner, B., Tichomirowa, M., et
al., 2009: The North American–Caribbean Plate boundary in Mexico – Guatemala – Honduras, in
James, K.H., Lorente, M.A., and Pindell, J.L. eds., The Origin and Evolution of the Caribbean Plate,
Geological Society, London, Special Publications, 219–293.
Reches, Z., and Lockner, D.A., 1994: Nucleation and growth of faults in brittle rocks. Journal of
Geophysical Research, 99 (B9), 18159–18173.
Reeside, J.B., 1944: Upper Cretaceous ammonites from Haïti. U.S. Geological Survey Prof. Paper, 214
(A), 1–5.
Reid, A., and Plumley, W., 1991: Paleomagnetic evidence for late Miocene counterclockwise rotation

328
References

of North Coast carbonated sequence, Puerto Rico. Geophysical Research Letters, 18 (3), 565–568.
Renard, F., Dysthe, D., Feder, J., Bjorlykke, K., and Jamtveit, B., 2001: Enhanced pressure solution creep
rates induced by clay particles: Experimental evidence in salt aggregates. Geophysical Research
Letters, 28 (7), 1295–1298.
Renard, F., Andréani, M., Boullier, A., and Labaume, P., 2005: Crack-seal patterns: records of
uncorrelated stress release variations in crustal rocks. Geological Society of London, Special
Publications, 243, 67–79.
Renne, P.R., Scott, G.R., Doppelhammer, S.K., Cala, E.L., and Hargraves, R.B., 1991: Discordant mid-
Cretaceous paleomagnetic pole from the Zaza terrane of Central Cuba. Geophysical Research
Letters, 18 (3), 445–458.
Révillon, S., Hallot, E., Arndt, N.T., Chauvel, C., and Duncan, R.A., 2000: A complex history for the
Caribbean Plateau: Petrology, geochemistry, and geochronology of the Beata Ridge, South
Hispaniola. The Journal of Geology, 108 (6), 641–661.
Rives, T., Rawnsley, K.D., and Petit, J.-P., 1994: Analogue simulation of natural orthogonal joint set
formation in brittle varnish. Journal of Structural Geology, 16 (3), 419–429.
Rogers, R.D., Mann, P., Emmet, P.A., and Venable, M.E., 2007: Colon fold belt of Honduras: Evidence
for Late Cretaceous collision between the continental Chortis block and intra-oceanic Caribbean
arc. Geological Society of America Special Papers, 428, 129–149.
Rojas-Agramonte, Y., Neubauer, F., Kröner, A., Wan, Y.S., Liu, D.Y., Garcia-Delgado, D.E., and Handler,
R., 2004: Geochemistry and early Palaeogene SHRIMP zircon ages for island arc granitoids of the
Sierra Maestra, southeastern Cuba. Chemical Geology, 213, 307–324.
Rojas-Agramonte, Y., Neubauer, F., Bojar, A.V., Hejl, E., Handler, R., and Garciá-Delgado, D.E., 2006:
Geology, age and tectonic evolution of the Sierra Maestra Mountains, southeastern Cuba.
Geologica Acta, 4 (1–2), 123–150.
Rojas-Agramonte, Y., Kröner, A., García-Casco, A., Somin, M., Iturralde-Vinent, M., Mattinson, J.M.,
Millán Trujillo, G., Sukar, K., Pérez Rodríguez, M., Carrasquilla, S., Wingate, M.T.D., and Liu, D.Y.,
2011: Timing and Evolution of Cretaceous Island Arc Magmatism in Central Cuba: Implications for
the History of Arc Systems in the Northwestern Caribbean. The Journal of Geology, 119, 619–640.
Rosencrantz, E., Ross, M.I., and Sclater, J.G., 1988: Age and spreading history of the Cayman Trough as
determined from depth, heat flow, and magnetic anomalies. Journal of Geophysical Research:
Solid Earth, 93 (B3), 2141–2157.
Rosencrantz, E., 1990: Structure and tectonics of the Yucatan Basin, Caribbean Sea, as determined
from seismic reflection studies. Tectonics, 9 (5), 1037–1059.
Rosencrantz, E., and Mann, P., 1991: SeaMARC II mapping of transform faults in the Cayman Trough,
Caribbean Sea. Geology, 19, 690–693.
Rosencrantz, E., 1996: Basement structures and tectonics in the Yucatan Basin, in In: Iturralde-Vinent,
M.A. (ed.), Ofiolitas y Arcos Volcánicos de Cuba. Miami, USA IGCP Project 364 Special
Contribution, 36–47.
Rotevatn, A., Sandve, T.H., Keilegavlen, E., Kolyukhin, D., and Fossen, H., 2013: Deformation bands and
their impact on fluid flow in sandstone reservoirs : the role of natural thickness variations.
Geofluids, 13, 359–371.
Rustichelli, A., Torrieri, S., Tondi, E., Laurita, S., Strauss, C., Agosta, F., and Balsamo, F., 2016: Fracture
characteristics in Cretaceous platform and overlying ramp carbonates: An outcrop study from

329
References

Maiella Mountain (central Italy). Marine and Petroleum Geology, 76, 68–87.
Rutter, E.H., Maddock, R.H., and Hall, S.H., 1986: Comparative Microstructures of Natural and
Experimentally Produced Clay-Bearing Fault Gouges. Pure and Applied Geophysics, 124 (1–2), 3–
30.
Saint-Fleur, N., 2014: Sismotectonique du système de failles d’Enriquilllo et du séisme du 12 janvier
2010 (Mw 7.0) en Haïti: 282 p.
Saint-Fleur, N., Feuillet, N., Grandin, R., Jacques, E., Weil-Accardo, J., and Klinger, Y., 2015:
Seismotectonics of southern Haiti: A new faulting model for the 12 January 2010 M7.0
earthquake. Geophysical Research Letters, 42 (23), 10273–10281.
de Saint-Méry, M.L.-E.M., 1798: Description topographique, physique, civile, politique et historique de
la partie française de l’isle Saint-Domingue. Société Française d’Histoire d’Outre-Mer, 3, 520p.
Salvini, F., Billi, A., and Wise, D.U., 1999: Strike-slip fault-propagation cleavage in carbonate rocks: the
Mattinata Fault Zone, Southern Apennines, Italy. Journal of Structural Geology, 21, 1731–1749.
Sammis, C., King, G., and Biegel, R., 1987: The Kinematics of Gouge Deformation. Pure and Applied
Geophysics, 125 (5), 777–812.
Samtani, M., Dollimore, D., and Alexander, K.., 2002: Comparison of dolomite decomposition kinetics
with related carbonates and the effect of procedural variables on its kinetic parameters.
Thermochimica Acta, 392, 135–145.
Sander, M. V., and Black, J.E., 1988: Crystallization and recrystallization of growth-zoned vein quartz
crystals from epithermal systems - implications for fluid inclusion studies. Economic Geology, 83
(5), 1052–1060.
Saunders, J.B., Jung, P., and Biju-Duval, B., 1986: Neogene paleontology in the northern Dominican
Republic; 1, Field surveys, lithology, environment, and age. Bulletins of American Paleontology,
89 (323), 110.
Savage, H.M., and Brodsky, E.E., 2011: Collateral damage: Evolution with displacement of fracture
distribution and secondary fault strands in fault damage zones. Journal of Geophysical Research,
116, 1–14.
Sayeed, V., Maurrasse, F., Husler, J., and Schmitt, R., 1978: Geochemistry and petrology of some mafic
rocks from Dumisseau, southern Haiti. EOS Transactions of the American Geophysical Union, 59
(4), 403.
Scherer, J., 1912: Great earthquakes in the island of Haiti. Bulletin of the Seismological Society of
America, 2 (3), 161–180.
Schmid, S.M., Casey, M., and Starkey, J., 1981: The microfabric of calcite tectonites from the Helvetic
Nappes (Swiss Alps). Geological Society, London, Special Publications, 9, 151–158.
Scholz, C.H., 1987: Wear and gouge formation in brittle faulting. Geology, 15, 493–495.
Sen, G., Hickey-Vargas, R., Waggoner, D.G., and Maurrasse, F., 1988: Geochemistry of basalts from the
Dumisseau Formation, southern Haiti: Implications for the origin of the Caribbean Sea crust. Earth
and Planetary Science Letters, 87, 423–437.
Sharp, Z.D., and Kirschner, D.L., 1994: Quartz-calcite oxygen isotope thermometry: A calibration based
on natural isotopic variations. Geochimica et Cosmochimica Acta, 58 (21), 4491–4501.
Shi, Z., Shi, Z., Yin, G., and Liang, J., 2014: Travertine deposits, deep thermal metamorphism and
tectonic activity in the Longmenshan tectonic region, southwestern China. Tectonophysics, 633,

330
References

156–163.
Shipton, Z.K., and Cowie, P.A., 2001: Damage zone and slip-surface evolution over μm to km scales in
high-porosity Navajo sandstone , Utah. Journal of Structural Geology, 23, 1825–1844.
Shipton, Z.K., Evans, J.P., Robeson, K.R., Forster, C.B., and Snelgrove, S., 2002: Structural heterogeneity
and permeability in faulted eolian sandstone: Implications for subsurface modeling of faults.
AAPG Bulletin, 86 (5), 863–883.
Shipton, Z.K., Evans, J.P., and Thompson, L.B., 2005: The Geometry and Thickness of Deformation-band
Fault Core and its Influence on Sealing Characteristics of Deformation-band Fault Zones. AAPG
Memoir, 85, 181–195.
Shipton, Z.K., Soden, A.M., Kirkpatrick, J.D., Bright, A.M., and Lunn, R.J., 2006: How Thick is a Fault ?
Fault Displacement-Thickness Scaling Revisited, in Earthquakes: Radiated Energy and the Physics
of Faulting, American Geophysical Union, 193–198.
Sigurdsson, H., Leckie, R.M., and Acton, G.D., 1997a: Site 1001. Proceedings of the Ocean Drilling
Program, Scientific Results, 165, 291–357.
Sigurdsson, H., Leckie, R.M., Acton, G.D., and et al., 1997b: Caribbean Volcanism, Cretaceous/Tertiary
Impact, and Ocean-Climate History: Synthesis of Leg 165. Proceedings of the Ocean Drilling
Program, Initial Reports, 165, 377–400.
Silliphant, L.J., Engelder, T., and Gross, M.R., 2002: The state of stress in the limb of the Split Mountain
anticline, Utah: Constraints placed by transected joints. Journal of Structural Geology, 24 (1), 155–
172.
Siman-Tov, S., Aharonov, E., Sagy, A., and Emmanuel, S., 2013: Nanograins form carbonate fault
mirrors. Geology, 41, 703–706.
Simon, J.L., Seron, F.J., and Casas, A.M., 1988: Sress deflection and fracture development in a
multidirectional extension regime. Mathematical and experimental approach with field
examples. Annales Tectonicae, 2 (1), 21–32.
Simpson, R.W., 1997: Quantifying Anderson’s fault types. Journal of Geophysical Research: Solid Earth,
102 (B8), 17909–17919.
Simpson, G., Gueguen, Y., and Schneider, F., 2001: Permeability enhancement due to microcrack
dilatancy in the damage regime. Journal of Geophysical Research-Solid Earth, 106 (B3), 3999–
4016.
Sinton, C.W., Duncan, R.A., Storey, M., Lewis, J., and Estrada, J.J., 1998: An oceanic flood basalt
province within the Caribbean plate. Earth and Planetary Science Letters, 155, 221–235.
Sinton, C.W., Sigurdsson, H., and Duncan, R.A., 2000: Geochronology and petrology of the igneous
basement at the lower Nicaraguan Rise, site 1001. Proceedings of the Ocean Drilling Program,
Scientific Results, 165, 233–236.
Smeraglia, L., Berra, F., Billi, A., Boschi, C., Carminati, E., and Doglioni, C., 2016: Origin and role of fluids
involved in the seismic cycle of extensional faults in carbonate rocks. Earth and Planetary Science
Letters, 450, 292–305.
Somoza, R., 2007: Eocene paleomagnetic pole for South America: Northward continental motion in the
Cenozoic, opening of Drake Passage and Caribbean convergence. Journal of Geophysical
Research: Solid Earth, 112 (3), 1–11.
Somoza, R., and Zaffarana, C.B., 2008: Mid-Cretaceous polar standstill of South America, motion of the

331
References

Atlantic hotspots and the birth of the Andean cordillera. Earth and Planetary Science Letters, 271
(1–4), 267–277.
Stanek, K.P., Maresch, W. V., and Pindell, J.L., 2009: The geotectonic story of the northwestern branch
of the Caribbean Arc: implications from structural and geochronological data of Cuba, in James,
K.H., Lorente, M.A., and Pindell, J.L. eds., The Origin and Evolution of the Caribbean Plate,
Geological Society, London, Special Publications, 361–398.
Stipp, M., Stünitz, H., Heilbronner, R., and Schmid, S.M., 2002: The eastern Tonale fault zone: A “natural
laboratory” for crystal plastic deformation of quartz over a temperature range from 250 to 700
°C. Journal of Structural Geology, 24 (12), 1861–1884.
Sulem, J., and Famin, V., 2009: Thermal decomposition of carbonates in fault zones: Slip-weakening
and temperature-limiting effects. Journal of Geophysical Research, 114 (B3), B03309.
Sykes, L.R., McCann, W.R., and Kafka, A.L., 1982: Motion of the Caribbean plate during the last 7 million
years and implications for earlier Cenozoic movements. Journal of Geophysical Research: Solid
Earth, 87 (B13), 10656–10676.
Symithe, S.J., Calais, E., Haase, J.S., Freed, A.M., and Douilly, R., 2013: Coseismic slip distribution of the
2010 M 7.0 Haiti earthquake and resulting stress changes on regional faults. Bulletin of the
Seismological Society of America, 103 (4), 2326–2343.
Symithe, S.J., Calais, E., de Chabalier, J.B., Robertson, R., and Higgins, M., 2015: Current block motions
and strain accumulation on active faults in the Caribbean. Journal of Geophysical Research: Solid
Earth, 120, 1–27.
Symithe, S., and Calais, E., 2016: Present-day shortening in Southern Haiti from GPS measurements
and implications for seismic hazard. Tectonophysics, 679, 117–124.
Tait, J., Rojas-Agramonte, Y., García-Delgado, D., Kröner, a., and Pérez-Aragón, R., 2009:
Palaeomagnetism of the central Cuban Cretaceous Arc sequences and geodynamic implications.
Tectonophysics, 470 (3–4), 284–297.
Tarasewicz, J.P.T., Woodcock, N.H., and Dickson, J.A.D., 2005: Carbonate dilation breccias: Examples
from the damage zone to the Dent Fault, northwest England. Geological Society of America
Bulletin, 117 (5/6), 736–745.
Tavani, S., Storti, F., Fernández, O., Muñoz, J.A., and Salvini, F., 2006: 3-D deformation pattern analysis
and evolution of the Añisclo anticline, southern Pyrenees. Journal of Structural Geology, 28 (4),
695–712.
Taylor, S.R., and McLennan, S.R., 1985: The Continental Crust: Its Composition and Evolution. Boston,
Blackwell Scientific Publications, 277p.
Teixell, A., Durney, D.W., and Arboleya, M., 2000: Stress and fluid control on décollement within
competent limestone. Journal of Structural Geology, 22, 349–371.
Tenthorey, E., Cox, S.F., and Todd, H.F., 2003: Evolution of strength recovery and permeability during
fluid-rock reaction in experimental fault zones. Earth and Planetary Science Letters, 206 (1–2),
161–172.
Terrier, M., Bialkowski, A., Nachbaur, A., Prépetit, C., Joseph, Y.F., and Joshph, Y.F., 2014: Revision of
the geological context of the Port-au-Prince metropolitan area, Haiti: Implications for slope
failures and seismic hazard assessment. Natural Hazards and Earth System Sciences, 14 (9), 2577–
2587.
Tesei, T., Collettini, C., Carpenter, B.M., Viti, C., and Marone, C., 2012: Frictional strength and healing

332
References

behavior of phyllosilicate-rich faults. Journal of Geophysical Research: Solid Earth, 117 (9), 1–13.
Tesei, T., Collettini, C., Viti, C., and Barchi, M.R., 2013: Fault architecture and deformation mechanisms
in exhumed analogues of seismogenic carbonate-bearing thrusts. Journal of Structural Geology,
55, 167–181.
Tesei, T., Collettini, C., Barchi, M.R., Carpenter, B.M., and Di Stefano, G., 2014: Heterogeneous strength
and fault zone complexity of carbonate-bearing thrusts with possible implications for seismicity.
Earth and Planetary Science Letters, 408, 307–318.
Tondi, E., Antonellini, M., Aydin, A., Marchegiani, L., and Cello, G., 2006: The role of deformation bands,
stylolites and sheared stylolites in fault development in carbonate grainstones of Majella
Mountain, Italy. Journal of Structural Geology, 28, 376–391.
Tondi, E., 2007: Nucleation, development and petrophysical properties of faults in carbonate
grainstones: Evidence from the San Vito Lo Capo peninsula (Sicily, Italy). Journal of Structural
Geology, 29 (4), 614–628.
Tondi, E., Cilona, a., Agosta, F., Aydin, a., Rustichelli, a., Renda, P., and Giunta, G., 2012: Growth
processes, dimensional parameters and scaling relationships of two conjugate sets of compactive
shear bands in porous carbonate grainstones, Favignana Island, Italy. Journal of Structural
Geology, 37, 53–64.
Torsvik, T.H., Van der Voo, R., Preeden, U., Niocaill, C. Mac, Steinberger, B., Doubrovine, P. V., van
Hinsbergen, D.J.J., Domeier, M., Gaina, C., Tohver, E., Meert, J.G., McCausland, P.J. a, and Cocks,
L.R.M., 2012: Phanerozoic polar wander, palaeogeography and dynamics. Earth-Science Reviews,
114 (3–4), 325–368.
Traineau, H., and Westercamp, D., 1980: Recherche geothermique en Republique d’Haïti, île
d’Hispaniola, Grandes Antilles. Rapport B.R.G.M. 80 SGN 920 GTH,, 42.
Trenkamp, R., Kellogg, J.N., Freymueller, J.T., and Mora, H.P., 2002: Wide plate margin deformation ,
southern Central America and northwestern South America, CASA GPS observations. Journal of
South American Earth Sciences, 15, 157–171.
Tritlla, J., Cardellach, E., and Sharp, Z.D., 2001: Origin of vein hydrothermal carbonates in triassic
limestones of the Espadán Ranges (Iberian Chain, E Spain). Chemical Geology, 172, 291–305.
Tschalenko, J.S., 1970: Similarities between Shear Zones of Different Magnitudes. Geological Society
Of America Bulletin, 81 (6), 1625–1640.
Tsutsumi, A., Nishino, S., Mizoguchi, K., Hirose, T., Uehara, S., Sato, K., Tanikawa, W., and Shimamoto,
T., 2004: Principal fault zone width and permeability of the active Neodani fault, Nobi fault
system, Southwest Japan. Tectonophysics, 379, 93–108.
Turner, F.J., Griggs, D.T., and Heard, H., 1954: Experimental deformation of calcite crystals. Geological
Society Of America Bulletin, 65 (9), 883–934.
Twiss, R.J., and Moores, E.M., 2007: Structural geology. 742.
Uehara, S. ichi, and Shimamoto, T., 2004: Gas permeability evolution of cataclasite and fault gouge in
triaxial compression and implications for changes in fault-zone permeability structure through
the earthquake cycle. Tectonophysics, 378 (3–4), 183–195.
Ünal-Ïmer, E., Tonguç Uysal, I., Zhao, J.-X., Isik, V., Shulmeister, J., Imer, A., and Feng, Y.-X., 2016: CO2
outburst events in relation to seismicity: Constraints from microscale geochronology,
geochemistry of late Quaternary vein carbonates, SW Turkey. Geochemica et Cosmochimica Acta,
187, 21–40.

333
References

Vandenberghe, N., Hilgen, F.J., Speijer, R.P., Ogg, J.G., Gradstein, F.M., Hammer, O., Hollis, C.J., and
Hooker, J.J., 2012: The Paleogene Period, in Gradstein, F.M., Ogg, J.G., Schmitz, M.D., and Ogg,
G.M. eds., The Geologic Time Scale, Boston, Elsevier, 855–921.
Vermilye, J.M., and Scholz, C.H., 1998: The process zone: A microstructural view of fault growth.
Journal of Geophysical Research, 103 (B6), 12223–12237.
Viganò, A., Tumiati, S., Recchia, S., Martin, S., Marelli, M., and Rigon, R., 2011: Carbonate
pseudotachylytes: Evidence for seismic faulting along carbonate faults. Terra Nova, 23 (3), 187–
194.
Vila, J.M., and Feinberg, H., 1982: Les discordances successives a la terminaison sud est de la coridllere
centrale Dominicaine. Bulletin de la Societe Géologique de France, 24 (1), 153–156.
Vila, J.M., Pubellier, M., Butterlin, J., and Boisson, D., 1986a: Le troncon central des decrochements
transhaïtiens: Chaine des Matheux et des Montagnes Noires. 11 ème R.A.S.T., Clermont-Ferrand,
Publ. Sp. Soc. Géol. Fr.,, 183.
Vila, J.M., Boisson, D., Butterlin, J., Feinberg, H., Laclede, J., and Pubellier, J., 1986b: Evolution
sedimentaire et structurale du bassin Oligo-Miocene de Trois Rivières, sur la frontiere
decrochante Nord-Caraïbe (Massif du Nord d’Haïti, Grandes Antilles). Rev. Géol. Dyn. Géogr.
Phys., 27 (3–4), 183–192.
Vila, J.M., Boisson, D., Butterlin, J., Feinberg, H., and Pubellier, M., 1987: Le complexe chaotique fini-
Eocene du Chouchou (Massif du Nord d’Haïti): Un enregistrement du debut des decrochements
senestres Nord-Caraïbes. C.R. Ac. Sc., Paris, 304 (II), 39–42.
Vila, J.-M., Pubellier, M., Jean-Poix, C., Feinberg, H., Butterlin, J., Boisson, D., Amilcar, H., and Amilcar,
H.C., 1988: Définition de la limite entre les blocs méridional et septentrional d’Hispaniola;
découverte d’un témoin de la nappe de Macaya dans l’anticlinal de Pierre Payen (centre d’Haïti,
Chaîne des Matheux, Grandes Antilles); implications géodynamiques. C.R. Ac. Sc. Paris, Ser. II,
307, 603–608.
Vincenz, S. a., and Dasgupta, S.N., 1978: Paleomagnetic study of some Cretaceous and tertiary rocks
on Hispaniola. Pure and Applied Geophysics, 116 (6), 1200–1210.
Viti, C., Collettini, C., and Tesei, T., 2014: Pressure solution seams in carbonatic fault rocks: mineralogy,
micro / nanostructures and deformation mechanism. Contributions to Mineralogy and Petrology,
167 (970), 1–15.
Vokes, E.H., 1979: The age of the Baitoa Formation, Dominican Republic, using mollusca for correlation:
Tulane Studies. Geology and Paleontology, 15 (4), 105–116.
Wadge, G., and Wooden, J.L., 1982: Late Cenozoic volcanism in the North-Western Caribbean: Tectonic
setting and Sr isotopic characteristics. Earth and Planetary Science Letters, 57, 35–45.
Walker, R.T., Talebian, M., Sloan, R.A., Rasheedi, A., Fattahi, M., and Bryant, C., 2010: Holocene slip-
rate on the Gowk strike-slip fault and implications for the distribution of tectonic strain in eastern
Iran. Geophysical Journal International, 181 (1), 221–228.
Wall, B.R.G., Girbacea, R., Mesonjesi, A., and Aydin, A., 2006: Evolution of fracture and fault-controlled
fluid pathways in carbonates of the Albanides fold-thrust belt. AAPG Bulletin, 90 (8), 1227–1249.
Walls, C., Rockwell, T., Mueller, K., Bock, Y., Williams, S., Pfanner, J., Dolan, J., and Fang, P., 1998:
Escape tectonics in the Los Angeles metropolitan region and implications for seismic risk. Nature,
394 (6691), 356–360.
Walsh, J.J., and Watterson, J., 1988: Analysis of the relationship between displacements and

334
References

dimensions of faults. Journal of Structural Geology, 10 (3), 239–247.


Walsh, J.J., Nicol, A., and Childs, C., 2002: An alternative model for the growth of faults. Journal of
Structural Geology, 24, 1669–1675.
Watkins, N.D., and Cambray, F.W., 1970: Palaeomagnetism of Cretaceous Dikes from Jamaica.
Geophysical Journal of the Royal Astronomy Society, 212, 163–179.
Wibberley, C.A.J., and Shimamoto, T., 2003: Internal structure and permeability of major strike-slip
fault zones: the Median Tectonic Line in Mie Prefecture, Southwest Japan. Journal of Structural
Geology, 25, 59–78.
Wibberley, C. a. J., Yielding, G., Di Toro, G., Toro, G.D.I., Wibberley, C. a. J., Yielding, G., Toro, G.D.I.,
Einstein, A., House, N.B., Lane, N.B., and Pe, L., 2008: Recent advances in the understanding of
fault zone internal structure : a review. Geological Society, London, Special Publications, 299 (1),
5–33.
Wiggins-Grandison, M.D., and Atakan, K., 2005: Seismotectonics of Jamaica. Geophysical Journal
International, 160 (2), 573–580.
Wilkins, S.J., Gross, M.R., Wacker, M., Eyal, Y., and Engelder, T., 2001: Faulted joints: kinematics,
displacement-length scaling relations and criteria for their identification. Journal of Structural
Geology, 23, 315–327.
Wilson, C.J.L., 1994: Crystal growth during a single-stage opening event and its implications for
syntectonic veins. Journal of Structural Geology, 16 (9), 1283–1296.
Wilson, J.E., Chester, J.S., and Chester, F.M., 2003: Microfracture analysis of fault growth and wear
processes, Punchbowl Fault, San Andreas system, California. Journal of Structural Geology, 25,
1855–1873.
Wiltschko, D. V, Lambert, G.R., and Lamb, W., 2009: Conditions during syntectonic vein formation in
the footwall of the Absaroka Thrust Fault, Idaho – Wyoming – Utah fold and thrust belt. Journal
of Structural Geology, 31 (9), 1039–1057.
Winslow, M.A., Guglielmo, G., Nadai, A.C., Vega, L.A., and McCann, W.R., 1991: Tectonic evolution of
the San Francisco Ridge of the eastern Cibao Basin, northeastern Hispaniol. Geological Society of
America Special Paper, 262, 301–313.
Witschard, M., and Dolan, J.F., 1990: Contrasting structural styles in siliciclastic and carbonate rocks of
an offscraped sequence: The Peralta accretionary prism, Hispaniola. Geological Society of
America Bulletin, 102 (6), 792–806.
van der Woerd, J., Klinger, Y., Sieh, K., Tapponnier, P., Ryerson, F.J., and M??riaux, A.S., 2006: Long-
term slip rate of the southern San Andreas Fault from 10Be-26Al surface exposure dating of an
offset alluvial fan. Journal of Geophysical Research: Solid Earth, 111 (4), 1–17.
Wojtal, S., and Mitra, G., 1986: Strain hardening and strain softening in fault zones from foreland
thrusts. Geological Society of America Bulletin, 97 (6), 674.
Woodring, W.P., 1922: Stratigraphy, structure and possible oil resources of the Miocene rocks of the
Central Plain. Geological Survey of the Republic of Haïti,, 19.
Woodring, W.P., Brown, J.S., and Burbank, W.S., 1924: Geology of the Republic of Haïti. Geological
Survey of the Republic of Haïti,, 631.
Woodward, N.B., Boyer, S.E., and Suppe, J., 1989: Balanced Geological Cross-Sections: An Essential
Technique in Geological Research and Exploration. Washington D.C., American Geophysical

335
References

Union, 132p.
Wu, S., and Groshong, R.H., 1991: Low-temperature deformation of sandstone, southern Appalachian
fold-thrust belt. Geological Society of America Bulletin, 103 (7), 861–875.
Yamaji, A., Sato, K., and Otsubo, M., 2005: Multiple Inverse Method Software Package. Division of Earth
& Planetary Sciences, Kyoto University,, 1–16.
Yasuhara, H., Marone, C., and Elsworth, D., 2005: Fault zone restrengthening and frictional healing:
The role of pressure solution. Journal of Geophysical Research, 110, 1–11.
Young, J.R., 1998: Neogene, in Brown, P.R. ed., Calcareous Nannofossil Biostratigraphy, London,
Chapman & Hall, 225–265.
Young, J.R., Bown, P.R., and Lees, J.A., 2018a: Nannotax3 website.
Young, J.R., Wade, B.S., and Hubert, B.T., 2018b: Pforams@mikrotax website.
Zhang, S., and Tullis, T.E., 1998: The effect of fault slip on permeability and permeability anisotropy in
quartz gouge. Tectonophysics, 295, 41–52.
Zoback, M.D., Zoback, M. Lou, Mount, V.S., Suppe, J., Eaton, J.P., Healy, J.H., Oppenheimer, D.,
Reasenberg, P., Jones, L., Raleigh, C.B., Wong, I.G., Scotti, O., and Wentworth, C., 1987: New
evidence on the state of stress of the San Andreas fault system. Science, 238 (4830), 1105–1111.
Zoback, M.D., and Healy, J.H., 1992: In situ stress measurements to 3.5 km depth in the Cajon Pass
Scientific Research Borehole: Implications for the mechanics of crustal faulting. Journal of
Geophysical Research, 97 (B4), 5039.
Zoback, M., Hickman, S., and Ellsworth, W., 2010: Scientific Drilling Into the San Andreas Fault Zone.
Eos, Transactions American Geophysical Union, 91 (22), 197–199.
Zoback, M., Hickman, S., and Ellsworth, W., 2011: Scientific drilling into the San Andreas fault zone -
An overview of SAFOD’s first five years. Scientific Drilling, (1), 14–28.
de Zoeten, R., and Mann, P., 1991: Structural geology and Cenozoic tectonic history of the central
Cordillera Septentrional, Dominican Republic, in Mann, P., Draper, G., and Lewis, J.F. eds.,
Geologic and Tectonic Development of the North America-Caribbean Plate Boundary in
Hispaniola, Boulder, Geological Society of America Special Papers, 265–279.
de Zoeten, R., and Mann, P., 1999: Cenozoic El Mamey Group of Northern Hispaniola: A Sedimentary
Record of Subduction, Collisional and Strike-Slip Events within the North America – Caribbean
Plate Boundary Zone, in Mann, P. ed., Sedimentary Basins of the World; Caribbean Basins,
Amsterdam, Elsevier Science B.V., 247–286.

336
Appendix 1

3 SSW
Enriquillo basin Sierra de Neiba NNE km
Cross section F 3
2 Sierra de Bahoruco NNE SSW Ns
2 Legend;
1 EPGFZ Enriquillo Cabritos 1 1 L. - m. Miocene
Js Ji Lake Ji
0 CVA
0 Limestone
-1 Ab-LS
Ni

-1
Eocene - Oligocene
-2 Limestone
Sc -1.951 m
-2 Upper Miocene
-3 Cretaceous island-arc
-3 Marl- and sandstone
- 40 5 10 15 20 25 30 km -4
Shortening line Fault
-5 -5
km
Cross section G Sierra de Neiba
3 SW NE S Enriquillo basin N SSW
km
NNE 3
2 Sierra de Bahoruco Montagnes Noires eq. 2
1 Charco Largo well
Matheux eq.
1
0 Ji
Qi
As
Tch
0
Ji Si

-1 Ns
Si -1
-2 Ab-LS
Ns
-2
Nb
-3 CVA
-3
Ni
-4 An
¿? Si
-4
-5 Tch
- 4.877 m
-5
km Sc

0 5 10 15 20 25 30 35 40 45 50 55 km

Cross section H Cross section I 10 km


Sierra de Azua Basin
SW Martín García
Sierra de Enriquillo NE S N SW NE
2 Bahoruco Basin

10 km
Palo Alto
1 Sc Proyectado + 6 km

0 Ns Qi

Ns
-1 Sc
Ab-LS

-2 -2.001 m
¿?

-3
-4
km

0 5 10 15 20 25 30 km

SW Cross section I NE
Peralta Belt 10 km
LPSJFZ SJRFZ

10 km
0 2 4 6 8 10
km

338
Appendix 2

Appendix 2 – Extended stratigraphic field observations


Southern Peninsula

The extended field observations on the Southern Peninsula listed below are an addition to the
stratigraphic work presented in section 4.2 of this thesis. The geographic places listed below can be
found on the geological compilation map (Figure 7 – 5) or on the small-scale geological maps in
appendix XX.

A2.1 Western domain (Tiburon)


Field observations
Around Les Anglais, in the Tiburon Valley, and east of Les Irois, strongly weathered basaltic flows and
pillows (Figure A2 – 1A), interbedded with pink to white limestones and radiolarites are observed. At
higher structural levels and apparently overlying the basalts are dm-bedded, well-stratified creamy
limestones with planktonic foraminifera and occasional chert nodules. These limestones are found
north of Bon Pas and on the southern side of the Tiburon Valley.

East of Anse-d’Hainault a series of cm-bedded grey claystones are observed. These claystones are in a
structurally higher position and appear to overly strongly weathered basalts found immediately to the
south.

Two isolated outcrops in the Tiburon Valley show a few 10’s of meters of dm-bedded grain- to
packstones rich in large benthic foraminifera and echinoderm debris indicating platform facies.
Massive limestones with similar fauna are found south of Rampe des Lions, which are impregnated
with bituminous residue.

Northwest of Port-a-Piment coarsely stratified white chalky limestones are found. Massive white
chalky limestones with intervals of rounded limestone pebbles and shell debris are found at Haut Fort
(Figure A2 – 1B). This sequence is directly overlying weathered basalts.

One kilometer south of Tiburon a 10 m thick succession of massive polymict breccia with meter-sized
clasts is overlain by at least 40 m of finer conglo-breccia containing predominantly basaltic material.

340
Appendix 2

conglomerates (Biju-Duval et al., 1980; Calmus, 1983). According to these authors the total thickness
of the Paleocene is laterally variable and ranges between 100 and 350 m. The Paleocene stratigraphy
is here divided into two parts; 1) a lower (Danian) part defined by predominantly (calcareous) clay- silt-
or sandstones, with or without volcaniclastics and basalts, and 2) an upper (Selandian and Thanetian)
part defined by predominantly silty calcarenites with occasional clastic intervals.

The platform limestones observed along the Tiburon Valley and at Rampe des Lions correspond well
with the description of Calmus (1983) and Amilcar (1997) for early Eocene limestones. Combining the
thicknesses of the lower (Biju-Duval et al., 1980) and middle to upper Eocene (Calmus, 1983) gives a
maximum thickness of 500 m for the Eocene interval.

The Oligocene is characterized by pelagic chalks with cherts and is part of the Jeremie Fm. (Maurrasse,
1980, 1982). Similar lithologies characterize the Aquitanian, which grade into chalky limestones
without cherts and marls of Burdigalian age (Bizon et al., 1985). The chalky limestones without cherts
at Port-a-Piment, which are overlain by 200m of Serravallian marls (Amilcar, 1997), are therefore
probably Burdigalian to Langhian in age. The chalky limestones with conglomerates at Haut Fort are,
by similarity, probably also Burdigalian to Langhian. The combined thickness of the Oligocene and
Aquitanian is estimated at 400m (Calmus, 1983; Amilcar, 1997), while the chalky limestones and marls
that make up the remainder of the Miocene estimated at a thickness of at least 300m (Calmus, 1983;
Bizon et al., 1985). The polymict breccia and basalts with m-sized clasts observed south of Tiburon are
probably Plio-Quaternary in age, although this is difficult to verify.

A2.2 Central domain (L’Asile)


Field observations
White dm-bedded chalky limestones with planktonic foraminifera and cherts, similar to the Macaya
Formation in the west, are observed southeast of Cavaillon and along the road between Bonne Fin and
Suzane (Figure A2 – 2A). A contact between these limestones and dark pillowed basalts is found south
of the EPGFZ at Mornes.

Within the Rivière Dose valley, around 1.5 km north of the EPGFZ, a volcaniclastic sequence is found
(Figure A2 – 2B). The basal part consists of a thick sequence of black volcaniclastics overlain by a
polymict breccia containing up to meter-sized clasts of black basalt and claystone, red siltstone and
minor limestones. The sequence is fining and deepening upwards, with several recurrences of
volcaniclastic breccia intervals, grading into turbiditic sand- and siltstones and calcarenites. The top is
capped by pillowed basalts rich in calcareous veins. The thickness of this sequence is at least 150m. A
similar breccia sequence is observed 3 km to the north in one of the tributaries to the Rivière Pins

342
Appendix 2

Interpretation
The age and thickness of the lithologies discussed below can be found on the stratigraphic columns in
appendices 4c – 4f of section 4.2, whose locations are indicated on the cross sections of figure 12.

The basalts southeast of l’Asile are similar to the Santonian basalts in the Aquin area (Amilcar, 1997)
and Upper Cretaceous basalts along the Grande Rivière de Nippes (Bien-Aime Momplaisir, 1986),
which are part of the CLIP. Pelagic limestones with cherts overlying these basalts have late Campanian
– early Maastrichtian ages southwest of Anse-a-Veau (Bien-Aime Momplaisir, 1986) and Campanian –
Maastrichtian ages in the l’Asile and Bonne Fin regions (Amilcar, 1997), and are interpreted to be part
of the Macaya Formation.

The breccia observed along the Rivière Dose and Pins has lithological similarities to the breccia
observed north of l’Asile by Bien-Aime Momplaisir (1986), which is overlain by Paleocene limestones.
This author also observed basalts along the Grande Rivière de Nippes with planktonic intercalations of
Globorotalia angulata, indicative of a latest Danian to middle Thanetian age (P3 – P4) (Olsson et al.,
1999). The Rivière Dose section is therefore tentatively placed in the early Paleocene and is expect to
be the lateral equivalent of the Rivière Glace Formation, with a thickness of several 100 meters. The
associated basalts are best interpreted as the equivalent of the lower Paleocene calc-alkaline volcanics
observed around Baradères and Morne des Orangers by Calmus (1983, 1987). The sand- and siltstones
directly overlying weathered CLIP basalts at Pallière are also interpreted to be the lateral equivalent of
the Rivière Glace Fm.

The multicolored clastic limestones further north appear stratigraphically higher in the sequence and
are placed in the upper Paleocene. This decision is made since platform limestones of early Eocene age
(Calmus, 1983; Bourgueil et al., 1988) that normally underlie these limestones were not observed,
which would otherwise have placed the clastic limestones in the Eocene or younger.

The cherty limestones west of Cavaillon are assigned a middle Eocene to Oligocene age (Calmus, 1983).
Following this author a thickness of some 500m is assigned to this sequence.

In the l’Asile basin, Bien-Aime Momplaisir (1986) found 300m of marls with lacustrine diatoms and
fresh water gastropods. In the same region Woodring et al. (1924) observed conglomerates and marls
with fresh water mollusks. These lacustrine deposits are apparently overlain by marls containing
Sphenolithus heteromorphus nannofossils (Bien-Aime Momplaisir, 1986) dated late Burdigalian to
Langhian (NN4 – NN5) (Young, 1998). The observations and timing fit well with the shallowing of facies
observed further west in the Camp Perrin area (Calmus, 1983). For now the l’Asile basin infill is
tentatively attributed to the Burdigalian and more recent.

344
Appendix 2

The fluvial clastics observed north of the Clonard Basin are unconformably overlying the Macaya
Formation and probably of Quaternary age (Mann et al., 1983), although no biostratigraphic dating is
available.

A2.3 Eastern domain (Marigot)


Field observations
The structurally lowest unit observed on the southern flank of the Massif de la Selle anticline, 3km to
the west of Morne la Visite, consists of massive basalts and pillow lavas with obsidian rims (Figure A2
– 3A). These basalts are overlain by around 250m of grey clay- and siltstones with quartz fragments
and planktonic foraminifera (Figure A2 – 3B). The upper 150m is more calcareous and interbedded
with microconglomerates and sandstones. The top of the ridge consists of around 150m of dm-bedded
yellow/orange limestones with marly intercalations. Along the road south towards Marche, white
packstones rich in large benthic foraminifera stratified in 50cm beds are outcropping. Stratigraphically
upwards and west of Marche these limestones grade into wacke- to mudstones with chert nodules. A
rhythmic alternation of m-bedded claystones and cm-bedded siltstones was observed at Ristache,
while at higher stratigraphic levels at Macary the silt- and claystones are both dm-bedded.

On the northern flank of the Massif de la Sell anticline the lowermost stratigraphic level is observed
around Furcy and consists of weathered basalt flows with sedimentary intercalations. West of Obleon
massive to dm-bedded yellow to beige platform limestones are observed in faulted contact with the
basalts. Upwards in the stratigraphy and towards Kenscoff pluri-dm bedded rose to beige limestones
with red algae and minor chert nodules are outcropping. At Fermate the facies are different with dm-
bedded chalky mudstones alternating with cm-beds of marl. Similar chalky limestones are observed at
La Boule, Toto and Boutillier quarries. At Les Cayettes quarry these durable chalky limestones make up
the hanging wall of a thrust fault, with the footwall consisting of yellow marlstones.

Interpretation
The age and thickness of the lithologies discussed below can be found on the stratigraphic columns in
appendices 4g and 4h of section 4.2, whose locations are indicated on the cross sections of figure 14.

The pillowed and massive basalts in the core of the Massif de la Selle anticline are part of the
Dumisseau Formation, with a minimum estimated thickness of 1500m (Maurrasse et al., 1979a) to
2000m (Desreumaux, 1987). In this region the basalts and intercalations are dated Albian to middle
Campanian, predominantly Turonian to Santonian (Sayeed et al., 1978; Mercier de Lépinay et al., 1979;
Van den Berghe, 1983a; Calmus, 1983; Bellon et al., 1985; Desreumaux, 1987; Sen et al., 1988).

345
Appendix 2

different nannofossil time scales, with all authors placing the platform limestones in nannofossil zone
NP 10, which at present is located in the early Eocene (Vandenberghe et al., 2012). I therefore follow
the division of Desreumaux (1987) and Bourgueil et al. (1988) and group the clay-, silt- and sandstones
in the lower Paleocene, the calcareous portion of this sequence in the upper Paleocene, and the
massive platform limestones in the lower Eocene.

Along the northern flank the massive platform limestones are in faulted contact with the basalts.
Following the same reasoning for these limestones on the southern flank they are placed in the early
Eocene, with a maximum thickness of 200m (Desreumaux, 1987). Van den Berghe (1983a) also
observed this contact and proposed that the Paleocene is missing in this area. If the Paleocene is
missing it would imply a major erosion event at the end of the Paleocene, which is not observed
anywhere else on the southern peninsula. The simplest solution is that the Paleocene is present in
most parts and that when absent, this is the result of later faulting.

The limestones observed around Kenscoff fit well with the description of the middle to upper Eocene
limestones in the northern Massif de la Selle (Van den Berghe, 1983a; Desreumaux, 1987). The
Oligocene in the northern Massif de la Selle is characterized by chalky limestones with occasional
cherts (Bourgueil et al., 1988). The total combined thickness of the Eocene and Oligocene is estimated
at 400m by Van den Berghe (1983).

The white chalky mudstones at La Boule, Toto, Boutillier and Les Cayettes quarries and at Fermate are
similar to the chalky mudstones described by Van den Berghe (1983a) in this region. According to this
author these limestones are deposited under bathyal conditions and can attain a maximum thickness
of 700m. The facies observed in the footwall at Les Cayettes quarry are similar to the mildly
consolidated calcareous marl- and sandstones of Van den Berghe (1983a), which can attain a thickness
of some 800m (Desreumaux, 1987). These calcareous – silty flysch-type detrital sediments are part of
the Rivière Grise Formation (Butterlin, 1954) and deposited during the late Messinian and Zanclean
(Andreieff et al., 1987).

At the onset of the Piacenzian (Desreumaux, 1987) sedimentation changes to sandstones and
conglomerates with interbedded mollusk-rich limestones (Jean Poix, 1980) of the Morne Delmas
Formation (Butterlin, 1960; Maurrasse, 1982). The Morne Delmas series locally reaches a thickness of
800m deposited in 1 Myr, with the upper 300m deposited in very shallow lagoonal conditions
(Bourgueil et al., 1988).

347
Appendix 2

Cul-de-Sac-1 (Atlantic) well


The Cul-de-Sac-1 dry well was drilled to a depth of 2408m below sea level by the Atlantic Refining
Company (now ARCO) in 1947. The exact location is difficult to find, but based on the maps and
information collected by Bourgueil et al. (1988) it is located at roughly 18.56°N and 72.30°W in the Cul-
de-Sac plain. The description below is based on the stratigraphic log provided in Bourgueil et al. (1988),
of which a schematic representation can be found in appendix 4i of section 4.2.

The Pliocene consists of an upper 150m of sands and conglomerates, which overlies 250m of sands,
silts and clays. The late Miocene has 400m of silts and clays overlying 300m of thinly bedded silty and
marly limestones. The middle Miocene consists of 250m of argillaceous limestones overlying 150m of
limestones with occasional cherts. The early Miocene contains 500m of cherty limestones. The
Oligocene consists of 150m of limestones with occasional cherts. The lowermost part is middle to late
Eocene and measures 250m of the same limestones with occasional cherts.

This sequence is similar to the descriptions for the northern flank of the Massif de la Selle (Van den
Berghe, 1983a; Desreumaux, 1987).

348
Appendix 4

Appendix 4: Representative bedding measurements


Bedding
Dip Dip
Long Lat Azimuth Dip Long Lat Azimuth Dip
-73,66116 18,34461 184 24 -72,291326 18,343135 155 30
-74,29650801 18,32172303 6 65 -72,259188 18,306479 212 17
-74,29656702 18,32146897 2 60 -72,266359 18,306553 178 28
-74,34273597 18,33652698 294 39 -72,272136 18,308199 190 19
-74,34169502 18,33655204 84 38 -72,277176 18,303173 189 8
-74,39773902 18,32204497 179 22 -72,288259 18,304997 226 33
-74,39760097 18,32209904 175 38 -72,300253 18,286725 302 22
-74,39624101 18,31630496 325 20 -72,56065 18,269173 150 24
-73,89326999 18,07157996 80 0 -72,568564 18,275456 183 21
-73,90075503 18,071773 270 10 -72,56913 18,279286 225 23
-73,87883099 18,07853401 137 35 -72,572079 18,281937 199 29
-73,87348903 18,08312503 167 16 -72,572429 18,282408 173 23
-73,87009403 18,08633597 194 31 -72,571978 18,284775 225 35
-73,86105398 18,09703404 254 39 -72,577378 18,292878 163 12
-73,85213404 18,11172299 346 37 -72,585016 18,302063 78 19
-73,74687502 18,23636801 75 10 -72,582111 18,319187 147 12
-73,74088003 18,24545299 285 16 -72,578913 18,326019 228 16
-73,73413897 18,26310603 312 18 -72,573535 18,334929 223 28
-73,70327798 18,28059697 205 47 -72,586531 18,34632 227 24
-73,51758798 18,38873098 25 68 -72,581615 18,362027 73 22
-73,50029703 18,37451099 225 40 -72,290734 18,429145 115 37
-73,38175597 18,35810903 42 40 -72,283058 18,433198 15 20
-73,364815 18,34798 324 27 -72,277432 18,434827 0 25
-73,36510397 18,34597598 265 30 -72,281801 18,438953 8 37
-73,70803203 18,26747099 233 20 -72,286241 18,453007 190 36
-73,69927302 18,28745504 47 20 -72,284349 18,456726 195 24
-73,69002996 18,293374 30 20 -72,30545 18,472678 25 29
-73,68534598 18,298738 255 11 -72,313432 18,486917 43 49
-73,65009497 18,29506103 352 12 -72,286883 18,50609 11 80
-73,64485603 18,29381799 210 28 -72,035664 18,67748 220 25
-73,63583601 18,29031897 211 30 -72,037847 18,677294 240 20
-72,28925804 18,74757199 0 20 -72,039161 18,677832 264 29
-72,29590204 18,74478199 175 43 -72,042243 18,678587 283 34
-72,67658704 18,93393598 230 56 -72,045923 18,676524 303 35
-72,31432303 18,67707297 356 26 -72,048056 18,673108 246 28
-72,16174498 18,69067898 197 52 -73,090865 18,381155 321 26
-72,33401396 18,50616597 21 26 -73,091574 18,350129 13 23
-72,32698197 18,50525896 31 49 -73,087929 18,303634 246 21
-72,35539704 18,50940004 352 39 -73,098438 18,295043 134 18
-72,284032 18,259505 141 25 -73,099382 18,292539 199 24

357
Appendix 4

-72,28556204 18,26263497 235 20 -73,102713 18,289111 215 51


-72,29491004 18,26580802 245 47 -73,114532 18,280024 188 31
-72,26128997 18,30672998 180 30 -73,119715 18,279031 29 41
-72,28862797 18,35315499 120 50 -73,126272 18,268064 297 29
-72,28767897 18,352983 147 43 -73,123336 18,259747 58 51
-72,28775902 18,35250003 163 46 -73,115815 18,252023 46 15
-72,28739097 18,35215696 138 39 -73,108101 18,245561 152 21
-72,28679401 18,35179604 115 29 -73,112167 18,231697 125 43
-73,656213 18,30425003 63 20 -74,106455 18,449342 195 25
-73,59192297 18,38497597 320 30 -74,108309 18,449473 208 30
-73,54542604 18,375547 184 30 -74,112533 18,473323 40 60
-73,50720498 18,41228697 189 31 -74,084077 18,512475 10 45
-73,50996598 18,42666303 48 40 -74,268324 18,553884 343 30
-73,49497796 18,43734903 15 34 -74,378411 18,486879 3 32
-73,49130099 18,44165498 55 35 -74,347944 18,483158 24 58
-73,48998402 18,44422697 16 65 -74,010001 18,586918 348 45
-73,48509997 18,45933401 5 85 -74,005816 18,584594 350 30
-73,10687003 18,43838403 256 21 -74,001229 18,564209 25 56
-73,060141 18,78777796 239 28 -73,984472 18,524250 10 40
-73,059249 18,78803202 183 60 -73,970745 18,496235 254 62
-73,05872102 18,78835296 357 25 -73,903321 18,424016 350 24
-73,05819103 18,788705 39 31 -73,852084 18,108494 26 30
-73,05845297 18,78916902 6 40 -72,765936 19,746871 213 15
-73,05691296 18,79120901 257 24 -72,844997 20,045961 240 30
-73,05199396 18,79327197 229 24 -72,852522 20,055351 20 22
-73,045601 18,800008 10 30 -72,862535 19,921337 220 15
-72,655123 19,03563403 170 19 -72,825058 19,894295 353 27
-72,66817697 19,04912001 190 21 -72,827173 19,887522 295 30
-72,68369196 19,072603 236 81 -72,828657 19,844943 15 30
-72,26159097 18,78117604 35 26 -72,800065 19,818896 258 50
-72,27662802 18,777271 187 11 -72,696199 19,710105 260 15
-72,28461597 18,76241399 45 21 -72,644986 19,571836 270 38
-72,28234598 18,75562096 340 41 -71,816624 18,367690 40 65
-72,30132999 18,74034301 75 25 -71,854154 18,390950 40 65
-72,30636701 18,73053996 47 18 -72,337655 18,484766 240 40
-72,317597 18,72692401 240 26 -72,333137 18,493555 92 20
-71,93073998 18,833704 25 24 -72,326979 18,496767 80 48
-71,94302198 18,76338696 212 61 -72,156731 18,692682 45 25
-71,94256702 18,76582299 35 60 -72,144180 18,757723 58 8
-71,94528401 18,765709 220 45 -72,153264 18,749854 45 18
-71,94740304 18,76510802 55 90 -72,160472 18,738473 65 22
-71,94922099 18,76597697 224 68 -72,166757 18,733088 37 18
-71,95515404 18,769851 42 52 -72,159210 18,712519 12 45
-71,95763097 18,77230003 248 70 -72,156207 18,708309 246 42
-71,95994002 18,77415402 263 49 -72,155857 18,706102 19 9

358
Appendix 4

-71,96303898 18,77761801 221 36 -72,152866 18,706020 229 33


-71,97309298 18,78584702 216 52 -72,153903 18,700678 38 45
-71,97107001 18,78881103 206 36 -72,160680 18,692875 351 30
-71,946045 18,48952596 258 26 -72,183310 18,661759 82 22
-73,587906 18,263066 232 21 -73,340684 19,673916 28 23
-73,645226 18,294149 220 20 -73,301568 19,817662 130 15
-73,777944 18,323536 90 35 -73,374295 19,820338 148 15
-73,849483 18,306820 200 25 -73,185081 19,835995 210 14
-74,099121 18,441430 4 35 -73,174477 19,817788 210 16
-74,099857 18,442666 338 20 -73,148782 19,803734 285 32
-74,102323 18,443574 15 28 -73,107692 19,771892 255 15
-73,089015 19,753239 63 28 -73,136180 18,442783 350 22
-73,082507 19,723366 55 28 -72,301234 18,686299 55 55
-73,483854 18,473289 330 76 -72,324249 18,726665 30 20
-73,361951 18,498926 180 10 -72,038787 18,505800 160 40
-73,229576 18,459561 20 66 -72,179929 18,848893 60 26
-73,235985 18,460505 16 80 -72,239551 18,784288 40 22
-73,180698 18,405396 144 26 -72,280103 18,754319 36 14

359
Appendix 5

Appendix 5: Whole-rock data


Trace elements Method LF200 LF200
HFSE REE Analyte SiO2 Al2O3
3d
LILE transition Unit wt % wt %
cg/g cg/g
MDL 0,01 0,01
Long Lat Sample Type Stop Description Type
-74,001405 18,560097 B1 Basalt 5.4a Pillow basalt, Baradères Formation Rock 46,49 14,25
-73,148782 19,803734 A2 Andesite JRAB 5 Andesite Rock 59,93 16,42
-72,907933 19,564119 B3 Basalt JRAB 12 Basalt Rock 61,97 15,19
-73,478894 18,466465 B4 Basalt 19.2a Cretaceous basalt Rock 45,45 12,98
-73,233588 18,459031 B5 Basalt 20.3b Cretaceous basalt Rock 47,96 14,31
-72,301752 18,688767 B6 Basalt MATH 5 Miocene basalt Rock 41,51 11,6
??? ??? B7 Basalt MATH8B Basalt (Vigie?) Rock 38,34 11,71
-72,045694 18,688767 B8 Basalt THOM5 Thomazeau basalts Rock 39,03 9,44
??? ??? B9 Basalt TRO1 Cretaceous basalt Rock 40,81 13,89
??? ??? BV9 Vein TRO1 Calcite vein in basalt Rock Pulp 12,42 5,61
-72,651972 18,344722 B10 Basalt TRO5 Cretaceous basalt Rock 59,58 0,34
-72,655722 18,370611 B11 Basalt TRO8 Cretaceous basalt Rock 23,67 8,28
-72,655722 18,370611 BZ11 Zeolite TRO8 Zeolites in basalts Rock Pulp 26,27 10,85
-72,655611 18,373917 B12 Basalt TRO9 Cretaceous basalt Rock 30,13 9,6
??? ??? B13 Basalt MATH10 Basalts Vigie Rock 42,21 11,84
-73,340297 19,673730 B14 Basalt BOMB3 Basalts Bombardopolis Rock 54,64 15,67
-74,345416 18,440436 B15 Basalt 4.5a Cretaceous basalt Rock 47,39 14,54
-72,644986 19,571836 4V Vein 9.9 Calcite veins Rock Pulp 0,28 0,04
-72,644986 19,571836 4HR Host rock 9.9 Limestone turbidites Rock 2,39 0,57

360
Appendix 5

LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200
Fe2O3 MgO CaO Na2O K2O TiO2 P2O5 MnO Cr2O3 Ba Ni Sc LOI Sum Be
wt % wt % wt % wt % wt % wt % wt % wt % wt % PPM PPM PPM wt % wt % PPM
cg/g cg/g cg/g cg/g cg/g cg/g cg/g cg/g cg/g μg/g μg/g μg/g cg/g cg/g μg/g
0,04 0,01 0,01 0,01 0,01 0,01 0,01 0,01 0,002 1 20 1 -5,1 0,01 1
Sample
B1 13,34 7 10,03 2,02 0,3 1,53 0,11 0,25 0,016 77 166 49 4,4 99,76
A2 6,32 2,58 4,24 4,59 2,56 0,83 0,28 0,09 1245 13 1,8 99,82 2
B3 5,63 3,52 5,77 3,74 0,83 0,74 0,27 0,09 0,01 606 83 9 2 99,8
B4 15,43 5,57 10,77 2,09 0,11 3,35 0,28 0,26 0,018 158 59 38 3,4 99,71 1
B5 12,39 5,58 10,68 2,23 0,25 2,6 0,21 0,15 0,018 81 78 33 3,3 99,74 2
B6 11,67 9,19 12,36 1,66 1,46 2,35 0,94 0,14 0,063 1378 232 23 6,2 99,34 2
B7 8,46 6,18 15,1 2,1 0,36 2,27 1,17 0,13 0,032 692 179 18 13,7 99,64 2
B8 13,11 11,68 13,66 3,31 2,06 3,95 2,22 0,17 0,046 1597 287 21 0,4 99,28
B9 10,39 6,04 13,42 2,24 0,82 1,02 0,11 0,13 0,032 35 93 42 10,9 99,77
BV9 0,09 0,37 41,79 2,98 0,03 0,01 0,02 0,15 10 2 36,5 99,98
B10 32,98 0,3 0,58 0,07 0,05 0,03 0,1 0,07 38 43 5,8 99,88
B11 8 2,97 29,35 0,64 1,07 0,8 0,11 0,21 0,019 31 46 25 24,8 99,86 2
BZ11 0,31 0,26 28,57 6,23 0,06 0,02 0,02 0,2 18 27,2 99,99
B12 8,46 4,06 22,59 1,16 1,32 0,88 0,17 0,18 0,018 26 51 35 21,2 99,81
B13 11,68 10,36 11,45 4,01 3,03 2,28 1,12 0,18 0,05 2526 259 21 0,9 99,45 3
B14 7,36 5,02 8,07 3,69 1,07 1,34 0,33 0,11 0,024 399 93 21 2,4 99,75 2
B15 11,7 7,97 12,77 1,89 0,06 1,13 0,08 0,2 0,048 32 122 45 1,9 99,75 2
4V 0,55 0,3 54,6 0,04 47 43,7 99,54
4HR 0,8 1,24 51,8 0,1 0,07 0,04 0,04 0,04 0,002 140 42,5 99,65

361
Appendix 5

LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200
Co Cs Ga Hf Nb Rb Sn Sr Ta Th U V W Zr Y
PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM
μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g
0,2 0,1 0,5 0,1 0,1 0,1 1 0,5 0,1 0,2 0,1 8 0,5 0,1 0,1
Sample
B1 76,2 18 2,3 4,8 6 1 133,1 0,3 0,4 0,2 368 82,2 23,2
A2 13 0,5 14,8 3 12,3 63,9 785,4 0,8 3,4 0,9 97 120,6 15,4
B3 16,3 0,2 16,9 4,6 19 8,1 1 612,2 1,3 6,2 1,6 81 0,8 216,6 12,6
B4 49,9 20,9 5,1 16,9 1,6 2 254,8 0,9 1,2 0,3 445 0,5 197 38,2
B5 42,8 19,7 3,8 13,5 3,2 2 264,5 0,7 0,9 0,2 378 155 29,4
B6 46 0,3 15,9 5,4 53,5 28 2 2989,5 2,8 20,9 3 238 0,6 235,7 22,1
B7 27,5 6,6 14,9 8 65,8 30,1 2 865,5 3,1 6,3 2 195 0,5 371,2 21,8
B8 51,1 0,4 18,5 10,4 122,2 42,2 3 2096,2 5,1 26,1 5,6 245 1,3 505,8 29,2
B9 37,2 0,2 18 1,4 4,2 12 324,6 0,3 0,4 0,5 284 51,9 19,1
BV9 0,5 1,1 48,8 0,6 4,4
B10 14,9 0,5 1 17,3 1 488 0,8 3,7 7
B11 20,5 7,9 1,3 3,2 12,9 201,5 0,2 0,3 0,2 170 43,7 16,4
BZ11 0,4 0,3 0,2 2,3 44,3 2,6 4,3
B12 26,7 0,3 9,2 1,5 3,2 15,6 3 287,9 0,2 0,4 0,3 248 47,3 44,6
B13 50 0,5 17,4 7,9 65,8 91,8 2 1615,2 3,4 20,2 3,6 246 1,9 370,1 22,6
B14 23,5 0,2 17,9 5,7 22,9 18,9 2 454,6 1,3 6,4 1,6 174 261,1 24,2
B15 46 14,7 1,6 3,7 0,5 111,6 0,1 0,3 322 55,8 17,9
4V 0,2 3798,8 1,3 0,3
4HR 0,9 0,1 0,1 0,4 2 2693,8 1,3 9 6,6 2,6

362
Appendix 5

LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 TC000
La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Tm Yb Lu TOT/C
PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM wt %
μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g cg/g
0,1 0,1 0,02 0,3 0,05 0,02 0,05 0,01 0,05 0,02 0,03 0,01 0,05 0,01 0,02
Sample
B1 3,8 10,2 1,49 7,9 2,63 1,09 3,85 0,66 4,33 0,93 2,76 0,4 2,75 0,44 0,07
A2 25,3 47,4 5,95 23,9 4,39 1,44 4,14 0,54 3,11 0,55 1,69 0,22 1,49 0,2 0,15
B3 34,1 58,5 5,6 19,1 3,3 1,12 2,94 0,41 2,31 0,45 1,19 0,17 1,04 0,16 0,09
B4 13,8 34,9 5 24,4 6,31 2,18 8,33 1,26 7,29 1,39 3,85 0,51 3,2 0,45 0,03
B5 11,3 27 3,82 18 4,91 1,73 6,2 0,93 5,13 1,01 2,82 0,38 2,47 0,37 <0,02
B6 103,9 192 20,05 71,1 10,54 3,03 8,34 0,99 4,86 0,75 1,93 0,23 1,54 0,2 0,05
B7 88,3 171,1 18,07 63,2 9,83 2,85 7,81 0,98 4,65 0,76 1,86 0,25 1,31 0,2 1,32
B8 224,1 431,3 45,93 162,7 21,96 5,84 15,09 1,51 6,38 0,96 2,3 0,26 1,7 0,21 <0,02
B9 4,7 10,7 1,4 7,2 1,8 0,72 2,66 0,47 3,02 0,68 2,14 0,29 1,91 0,29 1,04
BV9 0,3 0,5 0,07 0,4 0,17 0,09 0,35 0,06 0,54 0,14 0,44 0,06 0,38 0,07 9,8
B10 3,6 3,6 0,58 2,3 0,46 0,16 0,73 0,11 0,73 0,17 0,58 0,07 0,43 0,06 0,09
B11 3,9 8,9 1,24 6 1,85 0,72 2,61 0,45 2,89 0,63 1,88 0,25 1,6 0,25 6,01
BZ11 0,9 1,9 0,21 1,1 0,29 0,11 0,49 0,08 0,55 0,14 0,45 0,06 0,42 0,06 6,64
B12 6,8 19,3 2,85 14,5 4,53 1,77 7,03 1,3 8,64 1,93 6,24 0,89 6,08 0,93 4,15
B13 120,8 226,1 23,42 81,9 11,53 3,21 8,79 1,01 5,03 0,83 2,11 0,27 1,65 0,22 0,08
B14 33,2 60,2 6,54 25,3 4,81 1,51 4,94 0,72 4,39 0,87 2,5 0,34 2,12 0,33 0,09
B15 3,7 9 1,31 7 2,17 0,83 2,99 0,53 3,25 0,69 2 0,29 1,89 0,29 <0,02
4V 0,4 0,7 12,48
4HR 2,2 2,5 0,34 1,7 0,37 0,08 0,37 0,05 0,26 0,06 0,21 0,03 0,22 0,03 11,96

363
Appendix 5

TC000 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200
TOT/S Mo Cu Pb Zn Ni As Cd Sb Bi Ag Au Hg Tl Se
wt % PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPB PPM PPM PPM
cg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g ng/g μg/g μg/g μg/g
0,02 0,1 0,1 0,1 1 0,1 0,5 0,1 0,1 0,1 0,1 0,5 0,01 0,1 0,5
Sample
B1 0,2 74,3 8,4 100 131,3 1,4 0,1 3 0,01
A2 0,3 9,7 8,7 35 4,8 3,3 0,5 1,2
B3 0,03 0,6 65,6 7,4 29 12,4 2,5
B4 0,6 223,8 5,6 93 31,5 3,9 0,9
B5 0,3 170,5 5,1 88 44,4 0,1 2,7
B6 0,03 2,8 81,5 20,1 95 186 0,7 0,4 4,7 1,3
B7 1,6 49,3 4,2 96 159,6 1,3 0,6
B8 0,09 5,3 87 21,1 128 264,6 0,7 0,1 2,1
B9 0,1 67,7 2,7 48 50,8 0,8 0,1 1,3
BV9 1,5 0,6 1 6,3 0,5
B10 2,4 171,5 6,2 45 38,2 39,4 1,3 0,1 0,7 0,01
B11 84,9 2,5 43 27,2 1,9 0,3 6 0,9
BZ11 4,7 1,4 2 9,2 1,3 0,5
B12 112 1,8 61 28,8 1,7 0,1 0,02
B13 3,3 52,1 15,7 50 201,3 1,1 1,8 0,1
B14 0,8 111,7 3,3 42 66,8 0,8 0,1
B15 0,05 144,3 2,3 53 58 0,6 1,7
4V 0,7 1,6 0,5
4HR 0,02 15,6 1,7 10 2,7 1,9

364
Appendix 5

Method LF200 LF200


Analyte SiO2 Al2O3
wt % wt %
Unit
cg/g cg/g
MDL 0,01 0,01
Long Lat Sample Type Stop Description Type
-73,148782 19,803734 7V Vein JRAB 7 Calcite veins Rock Pulp 0,49 0,04
-73,148782 19,803734 7HR Host rock JRAB 7 Limestone Rock Pulp 13,09 1,44
-73,089015 19,753239 8V Vein JRAB 9 Calcite veins Rock 0,23
-73,089015 19,753239 8HR Host rock JRAB 9 Limestone Rock Pulp 28,44 0,79
-73,136180 18,442783 11V Vein 20.14 Calcite vein along normal fault Rock Pulp 0,46 0,02
-73,136180 18,442783 11HR Host rock 20.14 Limestone Rock Pulp 2,46 0,73
-72,298449 18,669361 13V Vein MATH 1 Calcite vein, 'orange' Rock 0,37
-72,298449 18,669361 13HR Host rock MATH 1 Limestone Rock 1,17 0,25
-72,305171 18,678424 14V Vein MATH 2a 'Vein' calcite Rock 0,34 0,03
-72,305171 18,678424 14HR Host rock MATH 2a Limestone Rock 2,85 0,91
-72,305171 18,678424 15V Vein MATH 2b 'Bedding parallel' calcite Rock 0,22
-72,305171 18,678424 15HR Host rock 21.2 Limestone Rock 0,65 0,18
-72,331732 18,724482 16V Vein 21.7 Calcite veins Rock Pulp 0,23 0,03
-72,331732 18,724482 16HR Host rock 21.7 Limestone Rock Pulp 4,02 1,14
-72,280103 18,754319 17V Vein 23.8b Calcite veins Rock Pulp 0,001
-72,280103 18,754319 17HR Host rock 23.8b Limestone Rock 0,2
Source
-73,284569 19,820982 18L Stalagtite Preston Stalagtitic limestone Rock 0,1
??? ??? 19V Vein Toro (Nadine) Calcite vein in cave Rock Pulp 1,35 0,03
??? ??? 19HR Host rock Toro (Nadine) Limestone from vein in cave Rock Pulp 3,87 1,13

365
Appendix 5

LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200
Fe2O3 MgO CaO Na2O K2O TiO2 P2O5 MnO Cr2O3 Ba Ni Sc LOI Sum Be
wt % wt % wt % wt % wt % wt % wt % wt % wt % PPM PPM PPM wt % wt % PPM
cg/g cg/g cg/g cg/g cg/g cg/g cg/g cg/g cg/g μg/g μg/g μg/g cg/g cg/g μg/g
0,04 0,01 0,01 0,01 0,01 0,01 0,01 0,01 0,002 1 20 1 -5,1 0,01 1
Sample
7V 0,07 0,08 54,74 0,01 0,06 18915 35 42,2 99,8
7HR 0,72 0,62 44,86 0,12 0,26 0,07 0,06 0,06 0,003 905 21 3 38,5 99,87
8V 0,3 0,1 55,65 0,05 16 43,5 99,83
8HR 0,21 0,26 38,02 0,1 0,16 0,03 0,04 0,03 342 1 31,8 99,94
11V 0,02 55,76 14 43,7 99,99
11HR 0,31 0,35 52,38 0,38 0,05 0,03 0,12 0,04 21 1 43,1 99,92
13V 0,18 0,17 55,51 26 43,7 99,99
13HR 0,22 0,19 54,18 0,1 0,02 0,02 0,01 58 43,8 99,97
14V 0,16 55,75 0,04 12 43,7 99,99
14HR 0,56 0,33 52,67 0,01 0,06 0,07 0,03 0,003 30 2 42,5 99,97
15V 0,08 0,05 55,85 11 43,8 99,99
15HR 0,23 0,17 55,23 0,02 0,02 0,03 14 43,5 99,98
16V 0,43 54,95 0,03 21 43,9 99,62
16HR 0,54 0,64 51,1 0,03 0,16 0,07 0,05 0,005 44 2 41,8 99,57
17V 0,38 55,61 44 99,97
17HR 0,08 0,32 55,71 0,02 0,004 10 43,6 99,94
18L 0,08 0,06 56,04 0,19 9 43,5 99,98
19V 0,05 0,03 34,86 0,03 <0,01 4 18,8 55,14
19HR 0,47 1,13 52,72 0,01 0,28 0,05 0,02 0,02 0,002 16 25 2 40,2 99,93

366
Appendix 5

LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200
Co Cs Ga Hf Nb Rb Sn Sr Ta Th U V W Zr Y
PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM
μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g
0,2 0,1 0,5 0,1 0,1 0,1 1 0,5 0,1 0,2 0,1 8 0,5 0,1 0,1
Sample
7V 0,1 0,2 0,4 1619,2 0,1 0,9 1,1
7HR 3,5 0,2 0,3 0,8 6,3 935,8 0,4 0,1 14 11,7 6,4
8V 1395,8 0,6 0,9
8HR 1,5 0,2 0,4 2,6 448 0,8 6,6 5,4
11V 0,2 15,1 4,3 1,9
11HR 2,6 0,2 0,3 0,7 2 531,1 0,5 0,2 9,5 8,7
13V 0,1 70,4 1,4
13HR 1,1 0,2 0,7 99,1 0,4 2,8 0,5
14V 44,1 0,3 0,8
14HR 2,5 0,2 0,7 1,6 104,4 0,2 0,5 15 8,5 1,8
15V 30,9 0,7 0,1
15HR 0,6 0,1 0,3 0,3 162,2 0,7 2,6 1,5
16V 0,2 3146,6 0,2 9 2,9 0,5
16HR 1,4 0,2 0,3 0,9 4,4 3335,3 0,6 4,4 34 10,2 8,3
17V 208,7 0,2 0,4 1,8
17HR 0,3 386,4 0,7 1 2,8
18L 0,2 92 0,7 0,1
19V 78,7 0,2 6 0,6
19HR 0,6 0,4 0,3 8,3 308,5 0,4 1 22 10,2 4,9

367
Appendix 5

LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 TC000
La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Tm Yb Lu TOT/C
PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM wt %
μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g cg/g
0,1 0,1 0,02 0,3 0,05 0,02 0,05 0,01 0,05 0,02 0,03 0,01 0,05 0,01 0,02
Sample
7V 0,7 0,6 0,07 0,4 0,34 0,1 0,02 0,06 0,01 0,06 12,04
7HR 5 5,5 1 3,9 0,97 0,24 1,01 0,15 0,9 0,18 0,56 0,07 0,52 0,06 10,63
8V 0,9 0,5 0,07 0,4 0,02 0,06 0,01 0,07 0,02 0,05 0,05 12,19
8HR 5,7 4 1,12 4,6 0,92 0,21 1,04 0,14 0,8 0,15 0,42 0,06 0,33 0,05 8,8
11V 0,8 0,2 0,07 0,4 0,07 0,02 0,14 0,04 0,15 0,02 0,15 0,03 12,37
11HR 6,2 5 1,18 4,8 1 0,26 1,28 0,18 1,11 0,23 0,68 0,08 0,61 0,09 11,68
13V 0,2 0,2 12,38
13HR 0,8 0,8 0,13 0,4 0,09 0,02 0,11 0,01 0,07 0,06 0,07 12,31
14V 0,3 0,1 12,33
14HR 1,5 2,6 0,31 1,5 0,24 0,07 0,32 0,04 0,27 0,06 0,14 0,02 0,11 0,02 11,9
15V 0,4 0,2 12,19
15HR 1,1 0,8 0,16 0,6 0,22 0,06 0,27 0,04 0,24 0,04 0,11 0,01 0,09 0,02 12,42
16V 0,7 0,4 0,05 0,05 12,44
16HR 6,1 7,3 1,24 5,1 1,06 0,25 1,18 0,16 1,11 0,22 0,7 0,08 0,44 0,07 11,53
17V 1,3 0,2 0,08 0,09 0,08 0,02 0,08 0,01 0,06 12,64
17HR 1,6 0,7 0,16 0,8 0,13 0,04 0,19 0,03 0,15 0,04 0,14 0,02 0,13 0,02 12,67
18L 0,4 0,2 12,16
19V 2,8 2,3 0,29 0,6 0,14 0,07 0,28 0,24 0,07 0,13
19HR 3,6 4,5 0,66 2,9 0,51 0,18 0,69 0,09 0,53 0,11 0,34 0,04 0,29 0,04 11,45

368
Appendix 5

TC000 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200
TOT/S Mo Cu Pb Zn Ni As Cd Sb Bi Ag Au Hg Tl Se
wt % PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPB PPM PPM PPM
cg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g ng/g μg/g μg/g μg/g
0,02 0,1 0,1 0,1 1 0,1 0,5 0,1 0,1 0,1 0,1 0,5 0,01 0,1 0,5
Sample
7V 0,49 1,3 0,8 2 18,6 1,6 0,6
7HR 0,03 9,6 1,8 13 19,5 1,1
8V 7,9 2,7 6 2,6
8HR 5,3 1,9 7 8,6 0,6 0,4 5,5
11V 0,5 0,4 8,4 0,5 0,1 4
11HR 0,02 4,5 1,8 11 22 0,8 0,4
13V 3,1 1,3 1,3 1,1
13HR 0,04 7,4 1,2 3 2,4 1,5 1,6
14V 0,8 0,2 1 1,3 1
14HR 0,03 6,8 0,7 4 5,7 0,8 1
15V 2,6 0,9 2 1,3 0,9
15HR 6,2 1,2 4 2,3 2 0,1
16V 0,4 0,6 42,2 1,2 0,7
16HR 3,3 7,5 1,2 7 11,8 2,7 0,2 0,1 1,6
17V 0,04 0,1 0,1 0,7
17HR 0,02 0,2 4,6 1 6 3,6 0,9 0,3 1
18L 4 1,1 4 1,2 0,5
19V 19,79 0,5 0,4 2 6,4 0,7 0,9 0,03
19HR 0,23 0,5 2,8 0,9 8 23,8 4,2 1,6 0,2 0,7 0,19

369
Appendix 5

Method LF200 LF200


Analyte SiO2 Al2O3
wt % wt %
Unit
cg/g cg/g
MDL 0,01 0,01
Long Lat Sample Type Stop Description Type
-73,517610 18,388723 B20 Basalt 17.6.0a Paleocene (?) basalt Rock 49,40 18,37
-73,517610 18,388723 BV20 Vein 17.6.0b Calcite vein in basalt Rock 1,06 0,22
-73,520563 18,386882 B21 Basalt 17.6.2 Basalt within conglomerate Rock 62,54 14,97
-72,288628 18,353155 B22 Basalt 17.13.3b Marigot Nord Cret pillow basalt Rock 46,62 14,46
-72,580436 18,358002 B23 Basalt 17.14.24a Diorite 'pillow' ? (Jacmel blocks) Rock 30,15 10,61
-72,580436 18,358002 B24 Basalt 17.14.24c Basalt pillow (Jacmel blocks) Rock 32,92 11,26
-72,306555 18,417560 B25 Basalt 17.15.1a Furcy Cretaceous basalt Rock 47,06 13,28
-73,507205 18,412287 B26 Basalt 17.19.17a Block of basalt in conglomerate Rock 58,71 17,14
-73,509966 18,426663 B27 Basalt 17.19.19b Paleocene (?) basalt Rock 48,07 17,20
-73,485100 18,459334 B28 Basalt 17.19.30 Paleocene (?) basalt Rock 65,93 4,54
-72,306367 18,730540 B29 Basalt 17.23.13e La Vigie basalts? Rock 40,53 12,62
-74,237116 18,310619 30HR Host rock 17.2.8a HR Cretaceous (?) slope carbonates Rock 48,38 2,94
-74,237116 18,310619 30V Vein 17.2.8a Silica vein (or chert vein?) Rock 88,33 1,13
-74,237116 18,310619 31V Vein 17.2.8b Calcite vein Rock 1,35 0,09
-74,373612 18,332768 32HR Host rock 17.3.3a HR Cretaceous slope carbonates Rock 20,15 0,22
-74,373612 18,332768 32V Vein 17.3.3a Silica vein (or chert vein?) Rock 95,68 0,18
-74,380675 18,331202 33HR Host rock 17.3.4 Cretaceous slope carbonates Rock 1,05 0,12
-74,380675 18,331202 33V Vein 17.3.4 Calcite vein Rock 2,82 0,03
-73,714789 18,261280 34HR Host rock 17.5.14b1 Eocene slope carbonates Rock 3,21 0,52
-73,517588 18,388731 35HR Host rock 17.6.1a1 Paleocene (slope) clastics Rock 25,32 6,79

370
Appendix 5

LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200
Fe2O3 MgO CaO Na2O K2O TiO2 P2O5 MnO Cr2O3 Ba Ni Sc LOI Sum Be
wt % wt % wt % wt % wt % wt % wt % wt % wt % PPM PPM PPM wt % wt % PPM
cg/g cg/g cg/g cg/g cg/g cg/g cg/g cg/g cg/g μg/g μg/g μg/g cg/g cg/g μg/g
0,04 0,01 0,01 0,01 0,01 0,01 0,01 0,01 0,002 1 20 1 -5,1 0,01 1
Sample
B20 7,41 5,12 8,62 3,34 0,33 1,10 0,19 0,09 0,020 85 68 23 5,8 99,80
BV20 0,38 0,49 53,87 0,03 0,02 0,02 0,02 0,47 119 <20 2 43,4 99,95 3
B21 7,42 1,67 6,36 3,96 0,68 1,05 0,26 0,08 0,010 295 25 16 0,8 99,89
B22 12,47 6,83 11,46 2,18 0,05 1,41 0,11 0,22 0,030 18 71 50 3,9 99,74
B23 5,63 3,42 26,11 0,98 1,26 0,78 0,16 0,17 0,024 64 39 29 20,6 99,87
B24 7,90 5,83 18,10 0,44 0,40 0,86 0,06 0,17 0,025 6 54 32 21,8 99,82
B25 13,94 6,16 10,42 2,23 0,09 3,35 0,27 0,19 0,016 43 69 35 2,7 99,67 3
B26 6,74 3,14 6,05 4,58 0,87 1,00 0,21 0,11 0,007 292 34 14 1,3 99,84
B27 8,90 3,81 11,16 2,99 0,20 2,76 0,29 0,11 0,017 296 56 42 4,2 99,74
B28 1,75 0,66 11,75 0,19 0,31 0,39 0,08 0,05 0,011 4138 11 13,8 99,92
B29 9,45 6,17 13,84 3,24 0,47 2,39 1,20 0,11 0,035 590 176 20 9,4 99,59 5
30HR 1,34 0,60 23,33 0,71 0,41 0,15 0,07 0,15 457 4 21,8 99,94
30V 1,76 0,30 3,99 0,30 0,07 0,13 0,08 0,04 0,003 76 1 3,9 99,99
31V 0,11 0,10 54,84 0,02 0,05 41 43,4 99,96
32HR 0,22 0,31 43,72 0,03 0,02 0,01 0,05 0,02 13 35,2 99,95
32V 2,19 0,02 0,67 0,05 0,03 0,01 0,04 0,02 0,004 296 1,1 100,02
33HR 0,17 0,77 53,90 0,02 0,02 0,01 0,02 5 43,8 99,91
33V 0,55 0,56 51,67 0,01 21 747 N.A. 55,76
34HR 0,35 0,30 52,96 0,04 0,07 0,02 0,07 0,04 332 1 42,3 99,92
35HR 1,51 0,73 34,98 1,43 0,29 0,29 0,12 0,22 0,003 565 5 28,2 99,90

371
Appendix 5

LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200
Co Cs Ga Hf Nb Rb Sn Sr Ta Th U V W Zr Y
PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM
μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g
0,2 0,1 0,5 0,1 0,1 0,1 1 0,5 0,1 0,2 0,1 8 0,5 0,1 0,1
Sample
B20 30,5 15,4 2,8 16,5 5,2 1 361,4 0,7 0,9 0,3 161 123,8 20,0
BV20 0,8 10,4 0,2 131,3 0,2 3,3 11,9
B21 15,2 0,1 12,9 2,5 9,2 12,8 352,8 0,7 0,8 0,2 72 121,4 18,9
B22 53,0 16,8 2,4 4,1 0,1 117,3 0,3 0,3 0,2 385 79,0 29,4
B23 22,5 0,2 10,5 1,3 2,5 16,4 126,5 0,2 0,3 0,4 207 40,7 15,2
B24 30,3 0,1 9,9 1,3 5,0 9,1 104,4 0,5 0,6 220 47,4 16,5
B25 64,7 21,7 5,4 15,9 1,9 2 263,0 1,1 1,4 0,3 441 210,6 47,1
B26 18,8 0,2 15,9 2,9 10,1 15,4 2 369,8 0,8 1,4 0,7 101 138,8 22,2
B27 40,9 22,2 4,6 30,8 1,0 1 366,8 1,7 2,0 0,5 410 171,5 27,8
B28 1,0 4,2 0,8 6,4 3,6 326,5 0,2 0,4 0,9 99 29,2 12,1
B29 40,2 4,2 18,1 9,3 68,7 25,6 2 1070,8 3,7 6,4 2,2 209 429,0 24,3
30HR 4,1 0,2 2,0 0,7 1,9 10,3 200,8 0,2 0,9 0,3 25 29,0 15,5
30V 3,8 0,2 1,4 0,5 4,8 2,2 60,3 0,4 0,4 0,2 18 23,1 4,5
31V 0,5 0,1 18,1 0,1 2,1 1,5
32HR 2,7 0,5 354,7 0,2 0,4 3,3 4,6
32V 2,5 3,3 0,3 1 13,3 0,4 2,1 2,5
33HR 0,5 1,0 0,3 606,2 0,9 1,5 3,4
33V 0,3 0,2 0,3 493,6 0,3 9,1 2,6
34HR 1,3 0,5 2,2 538,7 0,3 0,1 5,5 6,8
35HR 3,3 4,9 1,2 2,7 4,9 478,2 0,1 0,6 1,0 41 42,7 10,9

372
Appendix 5

LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 TC000
La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Tm Yb Lu TOT/C
PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM wt %
μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g cg/g
0,1 0,1 0,02 0,3 0,05 0,02 0,05 0,01 0,05 0,02 0,03 0,01 0,05 0,01 0,02
Sample
B20 8,8 18,5 2,35 10,3 2,67 1,05 3,23 0,56 3,31 0,81 2,29 0,32 2,10 0,34 0,08
BV20 2,7 3,2 0,44 2,0 0,39 0,16 0,85 0,17 1,30 0,36 1,16 0,19 1,29 0,22 12,55
B21 8,9 18,9 2,31 11,1 2,60 0,94 3,05 0,51 2,93 0,68 2,07 0,30 1,89 0,33
B22 3,9 10,6 1,66 8,7 2,92 1,13 4,27 0,80 5,16 1,16 3,39 0,49 3,41 0,50 0,35
B23 2,8 6,1 1,00 5,1 1,47 0,59 2,28 0,41 2,57 0,62 1,93 0,26 1,67 0,25 4,89
B24 3,0 7,8 1,12 6,3 1,85 0,69 2,47 0,44 2,83 0,68 2,00 0,27 1,81 0,28 3,28
B25 15,6 39,2 5,47 26,6 7,27 2,44 8,67 1,35 7,95 1,58 4,20 0,57 3,47 0,49 0,02
B26 13,0 23,6 2,88 12,0 2,78 1,02 3,26 0,59 3,65 0,78 2,38 0,37 2,29 0,38 0,02
B27 23,1 50,8 6,50 27,0 6,26 2,12 6,21 0,98 5,55 1,07 3,01 0,39 2,46 0,35 0,11
B28 5,4 6,0 0,99 4,1 1,00 0,34 1,39 0,23 1,40 0,36 0,96 0,14 0,96 0,13 2,46
B29 101,6 195,8 20,89 74,5 11,76 3,25 9,52 1,19 5,37 0,97 2,35 0,29 1,76 0,23 0,76
30HR 17,3 11,3 3,95 15,8 3,10 0,77 3,24 0,48 2,76 0,59 1,52 0,23 1,38 0,20 5,26
30V 7,0 10,6 1,48 5,7 1,02 0,26 1,00 0,14 0,76 0,13 0,32 0,05 0,33 0,04 0,78
31V 0,6 0,5 0,14 0,6 0,09 0,03 0,15 0,03 0,16 0,04 0,11 0,02 0,14 0,02 12,67
32HR 3,6 2,3 0,72 3,1 0,51 0,14 0,65 0,10 0,58 0,15 0,33 0,05 0,25 0,05 10,11
32V 2,2 1,9 0,51 1,9 0,39 0,10 0,43 0,06 0,39 0,06 0,20 0,02 0,16 0,01 0,16
33HR 2,7 1,3 0,38 1,7 0,28 0,07 0,34 0,06 0,29 0,07 0,24 0,03 0,22 0,03 12,65
33V 0,7 0,5 0,09 0,5 0,06 0,03 0,14 0,01 0,08 0,10 0,05 12,50
34HR 4,7 3,5 0,99 4,3 0,75 0,18 0,97 0,14 0,86 0,20 0,55 0,07 0,43 0,09 12,26
35HR 10,3 8,5 1,96 8,3 1,64 0,55 1,85 0,28 1,70 0,36 0,98 0,15 0,92 0,16 7,71

373
Appendix 5

TC000 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200
TOT/S Mo Cu Pb Zn Ni As Cd Sb Bi Ag Au Hg Tl Se
wt % PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPB PPM PPM PPM
cg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g ng/g μg/g μg/g μg/g
0,02 0,1 0,1 0,1 1 0,1 0,5 0,1 0,1 0,1 0,1 0,5 0,01 0,1 0,5
Sample
B20 2,0 71,6 3,3 59 66,6 0,5 2,5 0,5
BV20 1,1 118,5 2,9 13 1,8 1,3 1,5
B21 1,8 18,1 2,3 15 14,1 0,9
B22 0,04 0,7 221,4 1,4 81 47,0 6,1
B23 0,3 45,3 1,0 48 19,9 0,5 0,3 0,9
B24 0,2 109,0 1,1 59 49,6 0,2 5,4
B25 2,0 335,9 2,3 101 26,5 0,1 0,3 3,3
B26 3,4 22,1 1,1 17 7,0 0,6 1,1
B27 1,8 155,6 1,4 106 32,0 2,2 0,01
B28 0,6 44,4 3,6 73 13,4 1,1 0,4 0,1 0,8 0,02
B29 3,4 58,1 3,7 93 156,0 0,6 0,6
30HR 0,2 39,4 3,8 38 8,6 0,1 1,2
30V 0,02 3,2 23,0 2,1 12 17,2 1,2 0,1 1,2
31V 18,1 0,7 203 1,0 11,7 4,9
32HR 0,7 17,1 2,0 12 2,4 0,4 2,7
32V 3,4 24,7 4,7 6 6,4 1,9 0,1 0,3 1,9
33HR 0,02 0,4 14,8 1,7 10 1,0 0,4
33V 0,8 31,3 5,6 17 881,4 0,6 0,2 3,1
34HR 0,5 21,3 3,0 16 5,9 0,5 1,4
35HR 0,4 44,1 4,3 32 7,8 0,3 2,2 0,02

374
Appendix 5

Method LF200 LF200


Analyte SiO2 Al2O3
wt % wt %
Unit
cg/g cg/g
MDL 0,01 0,01
Long Lat Sample Type Stop Description Type
-73,517804 18,388011 36HR Host rock 17.6.1g Paleocene volcaniclastics Rock 61,19 16,27
-73,517804 18,388011 36V Vein 17.6.1g Quartz vein Rock 96,15 0,11
Middle Eocene platform
-72,299477 18,669212 37HR Host rock 17.9.2a carbonates Rock 1,03 0,23
-72,288086 18,353305 38V Vein 17.13.4b Mixed vein in Cret. Basalts Rock 88,85 0,31
-72,279116 18,433858 39V Vein 17.15.4 Weird mixed vein Rock 88,07 0,70
-72,278064 18,434264 40V Vein 17.15.5 Calcite vein Rock 0,37
-73,490462 18,443222 41HR Host rock 17.19.26c Paleocene slope carbonates Rock 24,61 1,28
-73,490462 18,443222 41V Vein 17.19.26c Calcite vein Rock 0,56
-72,658580 19,030411 42HR Host rock 17.22.2 Eocene (?) platform carbonates Rock 0,31
-72,658580 19,030411 42V1 Vein 17.22.2 Calcite vein (1) Rock 0,27
-73,490462 18,443222 43HR Host rock 17.19.26a2 Paleocene slope carbonates Rock 18,29 1,45
-73,490462 18,443222 43V Vein 17.19.26a2 Calcite vein Rock 0,47
-74,373612 18,332768 44V Vein 17.3.3b Silica vein (or chert vein?) Rock 96,66 0,08
-73,490462 18,443222 45HR Host rock 17.19.26a1 Paleocene slope carbonates Rock 18,35 1,98
-73,490462 18,443222 45V Vein 17.19.26a1 Calcite vein Rock 0,48 0,02
-73,123336 18,259747 B46 Basalt 17.18.13b Block of basalt in conglomerate Rock 42,79 15,23
-72,668177 19,049120 47HR Host rock 17.22.8 Volcaniclastic (?) sandstone Rock 27,48 7,46

375
Appendix 5

LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200
Fe2O3 MgO CaO Na2O K2O TiO2 P2O5 MnO Cr2O3 Ba Ni Sc LOI Sum Be
wt % wt % wt % wt % wt % wt % wt % wt % wt % PPM PPM PPM wt % wt % PPM
cg/g cg/g cg/g cg/g cg/g cg/g cg/g cg/g cg/g μg/g μg/g μg/g cg/g cg/g μg/g
0,04 0,01 0,01 0,01 0,01 0,01 0,01 0,01 0,002 1 20 1 -5,1 0,01 1
Sample
36HR 4,25 1,80 7,44 3,68 0,26 0,78 0,13 0,05 0,005 224 22 11 4,0 99,88
36V 2,01 0,01 0,42 0,05 0,02 0,01 0,02 0,004 6 1,2 100,02 5
37HR 0,16 0,29 54,76 0,05 0,02 0,02 0,002 8 43,4 99,92
38V 1,98 0,43 4,32 0,07 0,02 0,01 0,06 0,003 31 2 3,9 99,99
39V 2,29 0,38 4,09 0,12 0,11 0,07 0,04 0,04 0,004 8 1 4,1 100,00 1
40V 0,09 0,07 55,65 0,01 2 43,8 99,99
41HR 0,94 0,43 39,48 0,04 0,20 0,09 0,07 0,07 0,002 244 21 3 32,7 99,92
41V 0,18 0,12 55,30 0,16 134 43,6 99,96
42HR 0,11 0,28 55,61 0,17 3 43,5 99,96
42V1 0,10 0,03 56,05 0,06 3 43,5 99,99 2
43HR 1,01 0,42 42,95 0,04 0,66 0,11 0,07 0,07 0,002 926 21 4 34,8 99,92
43V 0,10 0,04 55,50 0,12 5590 43,1 99,96 2
44V 1,84 0,44 0,03 0,02 0,02 9 0,9 100,01
45HR 1,54 0,50 42,14 0,06 0,48 0,13 0,13 0,05 0,002 2001 24 5 34,3 99,90 2
45V 0,11 0,06 55,53 0,11 2051 43,4 99,96
B46 9,33 2,95 9,93 1,63 1,87 1,65 0,30 0,16 0,021 407 76 47 13,9 99,81 2
47HR 3,65 2,14 28,94 0,44 0,61 0,47 0,06 0,20 0,022 567 31 22 28,3 99,84 2

376
Appendix 5

LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200
Co Cs Ga Hf Nb Rb Sn Sr Ta Th U V W Zr Y
PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM
μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g
0,2 0,1 0,5 0,1 0,1 0,1 1 0,5 0,1 0,2 0,1 8 0,5 0,1 0,1
Sample
36HR 11,1 0,1 12,3 2,6 7,4 4,0 1 429,4 0,5 1,5 0,6 87 111,5 15,2
36V 2,4 1,5 0,1 7,1 0,3 0,5 1,4 0,5
37HR 0,6 548,8 0,7 2,6 1,7
38V 2,1 1,4 96,4 0,2 10 0,5 1,0 3,7
39V 4,0 0,2 1,2 1,2 18,8 0,2 14 5,4 3,2
40V 10,9 0,5
41HR 2,8 0,1 0,3 1,3 4,6 436,3 0,4 18 13,0 16,5
41V 0,3 246,0 0,6 12,0
42HR 3,5 219,5 1,5 1,3 1,8
42V1 8,7 0,5 0,1
43HR 3,2 0,4 2,9 4,1 445,3 0,2 0,4 25 14,0 16,3
43V 0,1 270,2 3,0 10,6
44V 1,9 1,1 3,0 0,1 2,9 0,2
45HR 4,0 0,1 1,7 0,4 1,2 4,7 511,6 0,5 25 21,8 26,8
45V 222,1 2,4 12,8
B46 31,7 0,4 16,1 2,3 4,6 32,1 121,6 0,3 0,5 1,3 310 0,6 83,8 32,2
47HR 10,7 0,1 5,6 0,7 1,4 8,3 688,1 0,2 0,2 137 23,9 11,5

377
Appendix 5

LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 LF200 TC000
La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Tm Yb Lu TOT/C
PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM wt %
μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g cg/g
0,1 0,1 0,02 0,3 0,05 0,02 0,05 0,01 0,05 0,02 0,03 0,01 0,05 0,01 0,02
Sample
36HR 11,0 19,9 2,37 9,9 2,32 0,81 2,66 0,44 2,47 0,53 1,68 0,26 1,57 0,25 0,33
36V 0,3 0,4 0,04 0,05 0,06 0,02 0,27 0,06 0,11
37HR 1,0 1,2 0,19 0,9 0,15 0,04 0,17 0,03 0,20 0,05 0,11 0,01 0,09 0,01 12,61
38V 0,9 2,1 0,27 1,4 0,43 0,14 0,54 0,08 0,47 0,12 0,38 0,07 0,64 0,13 0,98
39V 2,0 2,4 0,31 1,5 0,28 0,11 0,45 0,06 0,45 0,08 0,30 0,04 0,25 0,04 0,85
40V 0,2 12,93
41HR 9,9 6,1 1,60 7,0 1,31 0,39 1,84 0,29 1,70 0,45 1,40 0,19 1,34 0,20 9,17
41V 12,5 7,9 1,50 6,2 1,04 0,27 1,38 0,20 1,38 0,33 0,94 0,14 0,86 0,13 12,53
42HR 0,8 0,9 0,12 0,4 0,09 0,02 0,15 0,02 0,15 0,04 0,12 0,01 0,09 0,01 12,80
42V1 0,1 0,2 12,79
43HR 10,3 6,0 1,52 6,9 1,25 0,39 1,82 0,28 1,64 0,43 1,31 0,20 1,23 0,22 9,81
43V 8,6 4,5 0,92 3,2 0,58 0,39 0,77 0,14 0,89 0,25 0,81 0,12 0,70 0,11 12,48
44V 0,2 0,3 0,03 0,09
45HR 15,7 7,7 2,67 11,9 2,31 0,61 2,88 0,46 2,82 0,70 2,04 0,28 1,92 0,28 9,58
45V 13,3 5,5 1,37 5,3 0,66 0,25 0,92 0,15 1,17 0,30 1,03 0,16 1,08 0,16 12,61
B46 9,7 16,2 2,52 12,5 3,30 1,16 4,63 0,83 4,83 1,10 3,34 0,47 2,94 0,43 1,36
47HR 4,6 5,2 0,99 4,2 1,24 0,41 1,57 0,27 1,69 0,39 1,18 0,18 0,98 0,17 6,24

378
Appendix 5

TC000 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200 AQ200
TOT/S Mo Cu Pb Zn Ni As Cd Sb Bi Ag Au Hg Tl Se
wt % PPM PPM PPM PPM PPM PPM PPM PPM PPM PPM PPB PPM PPM PPM
cg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g μg/g ng/g μg/g μg/g μg/g
0,02 0,1 0,1 0,1 1 0,1 0,5 0,1 0,1 0,1 0,1 0,5 0,01 0,1 0,5
Sample
36HR 0,5 60,7 3,4 44 20,9 9,7
36V 6,6 15,9 0,9 5 10,9 1,6 0,2 2,0
37HR 0,2 7,4 0,9 5 1,6 0,1 1,1
38V 0,09 4,4 21,1 1,2 14 13,8 1,0 0,1 0,2
39V 3,8 17,7 0,9 15 14,9 1,3 0,1 1,6
40V 0,1 10,5 1,2 6 0,8 1,0
41HR 1,6 32,8 3,6 33 18,2 0,6 0,2 1,0
41V 0,3 21,0 2,2 12
42HR 0,5 10,8 1,0 8 0,9 0,8 7,8 0,8
42V1 0,2 13,3 1,1 9 1,1 1,0 2,8
43HR 0,4 69,7 5,0 52 20,3 0,2 0,1 0,1 0,6 0,01
43V 0,13 0,2 9,0 1,0 7 0,5 1,6
44V 2,7 20,4 2,5 6 4,7 1,4 0,2 0,9
45HR 0,04 0,4 38,8 4,7 51 20,8 0,3 0,1 0,5
45V 0,06 0,2 4,2 0,6 3 0,7 0,5
B46 0,2 61,1 2,1 251 62,3 0,5 0,1 1,2
47HR 0,5 35,1 1,4 35 25,8 1,6 0,9 0,03

379
Acknowledgements

Acknowledgements

I would like to take this opportunity to thank IFPEn, ISTeP, UEH, URGéo, BME, and all the people who’ve
I’ve been working with and that have helped to make this PhD into a success.

Nadine Ellouz-Zimmermann and Sylvie Leroy, my PhD supervisors. I first and foremost want to thank
you both for instigating this project and for choosing me to execute it. It is almost four years ago to
date that I learned about this project, which combined a whole suite of disciplines (sedimentology,
structural geology, geodynamics, and geochemistry) in an area that is was completely unfamiliar with
(the Caribbean in general, and Haiti in specific). It was challenging and ambitious from the onset, and
I am glad that I applied for it when the opportunity arose. Nadine, thank you for being the driving
force, for the continuous stream of comments and corrections, and for pushing me to continuously
reshape my understanding and explanation of the geology of Haiti. We did not always agree on the
same hypothesis, but that is part of doing science; challenging ideas, and progress while doing so. I
really enjoyed our discussions in the field and in the office, and I think we’ve come a long way in better
understanding the evolution of this part of the planet. Sylvie, thank you for giving me the space to
shape my own ideas about the Caribbean, and for fact-checking and exposing it when I took a wrong
turn. Your knowledge on the Caribbean proved invaluable for this project. I would also like to thank
you for the additional funding you managed to secure and push this project a few months into
(additional) overtime.

Claudio Rosenberg and Nicolas Bellahsen, my other (unofficial) supervisors. Somewhere along the
way both of you rolled into this project and haven’t let go ever since, and I am really grateful for that.
Claudio, thank you for being the voice of reason and tranquility, and for your positive attitude and
comments. Nico, thank you for the positive energy and optimism you brought to this project and to
the field. I have to especially thank the both of you for the efficient team work in the field, without
that it would not have been possible to collect 1500+ measurements in a month’s time (and for making
the best out of Day 9 in 2017, when I spend most of my day being ill in the back of the car…).

Claude Prepetit, Dominique Boisson and Roberte Momplaisir, my Haitian connection. Claude, thank
you for the possibility to work in Haiti and for providing the geological reports and maps, meetings,

380
Acknowledgements

the field logistics, and the permits. Dominique, thank you for sharing your knowledge and experience
of the geology of Haiti, with the field logistics, and helping us find the right outcrops. Roberte, thank
you for your wisdom, sharp comments, your geological advices, and thank you so much for the
hospitality and dinners at your place, for taking me around Port-au-Prince and showing me the culture
of Haiti.

Other members of the jury. Sveva and Manu, thank you for reviewing my thesis, and thank you for
the comments that undoubtedly will greatly improve this manuscript. James and Rudy, I am looking
forward to the discussions about the paleo-fluid circulation in southern Haiti, I am sure that will provide
some interesting new angles to the interpretation of the data. Olivier, thank you for finding the time
and agreeing to be the president of my jury.

Anne Battani and Anne Verlaguet, the geochemistry gurus, hereinafter abbreviated to Anne B. and
Anne V., to avoid confusion. Anne B., thank you for your insights into the present-day fluid-circulation
in Haiti, and for the discussions on how it got to be that way. A highlight of the 2015 fieldwork was
definitely the 10 km hike through the Rivière Bras à Droite in the pouring rain to collect that one specific
sample. Anne V., thank you for all the help on the fluid inclusions and geochemistry in general, how to
interpret the results, and how to improve them. You also came into this project along the way, and I
am grateful for all the efforts you’ve put in ever since.

Remy Deschamps, Youri Hamon and Daniel Pillot. Remy and Youri, thank you for all your work on the
stratigraphy in Haiti, which is so aptly summarized in an internal IFPEn report. Youri, thank you in
particular for the fieldwork in 2015, for your help in making sense of all the different carbonates that
Haiti has on offer, and thank you for offering me your water shoes for the aforementioned river hike
(even though they only limited the blisters to around 15). Daniel, thank you for your relaxed and calm
attitude in the field.

Laurent Emmanuel and Herman Ravelojaona. Laurent, thank you for the work on my stable oxygen
and isotope samples, the XRD, and how to make sense of all that data. Herman, thank you for the quick
turnaround times on the large number of thin sections, and for QC-ing their thickness with me.

381
Acknowledgements

Abdellatif Lahfid and Damien Deldicque. Abdellatif, thank you for the three days at the BRGM in
Orléans and the effort to try and find a useful Raman signal from the vein and host-rock data, and
thank you for the warm welcome at your place. Damien, thank you for the opportunity to test the fluid
inclusions samples with the Raman at ENS in Paris.

Daniel, Frednel, Oscar and Ronald, the drivers and local guides. Thank you all for driving us safely
around Haiti, for your no-nonsense mindset, and for the great in-car music.

Camesuze, Marceau, Newdeskarl and Rosemund, my companions in the field. Thank you for spending
time with me in the field, and I hope we’ve learned something from one another back and forth.
Rosemund, thank you especially for your optimism (which only decreased once while sailing to
Gonâve) and for continuously pulverizing large quantities of pillow basalts and other rocks to allow for
easier sampling.

Vasilis and Nikolaos, you two malakas. What a time it’s been. You both mean a lot to me, and I’m glad
that we shared this PhD-ride together. Happy hunting in Cyprus and I’ll see you soon.

Anouk Beniest. Dankjewel. Dankjewel dat je me mee hebt genomen op weer een avontuur! Dank je
voor al je support en vertrouwen, voor je liefde, en voor je geduld. Ik kan niet wachten om het volgende
avontuur aan te gaan, waar dat ook mag zijn. Ik heb er zin in lief!

382

You might also like