<<

Quantitative palinspastic tectonic reconstruction of northwestern South America and Caribbean since Late times

Presented by: Andrés Felipe Rodríguez Corcho

In Partial Fulfilment of the Requirements for Undergraduate Thesis is Geosciences

Department of Geosciences Universidad de los

May 6, 2016 Quantitative palinspastic tectonic reconstruction of northwestern South America and Caribbean since Late Cretaceous times

------Student

------Advisor

Universidad de los Andes Sciences Faculty Department de Geosciences May of 2016 Contents

Introduction: ...... 1 Methods:...... 2 Regional Geological Setting: ...... 3 : ...... 4 Caribbean plate displacement: ...... 5 Mantle structure evidence: ...... 6 Caribeana: ...... 7 Central America (Chortis, Siuna, Panama Blocks): ...... 7 Cayman Trough: ...... 10 Great Arc of the Caribbean: ...... 10 Aves Ridge, Lesser Antilles and Eastern Caribbean Basins: ...... 10 Cuba and Hispaniola: ...... 11 An Intraoceanic Arc: ...... 12 Leeward Antilles and Venezuela sedimentary basins: ...... 12 Guajira and Sierra Nevada de Santa Marta: ...... 13 Western Cordillera, Ecuadorian Andes and Tumaco Fore-Arc: ...... 17 Central Cordillera, Arquía and Quebradagrande complexes: ...... 19 Eastern Cordillera: ...... 19 Piercing Points: ...... 24 Eocene times (~50 Ma): ...... 24 Early (~21 Ma): ...... 25 Middle Miocene (15-13 Ma): ...... 25 Tectonic reconstruction of Andean blocks: ...... 27 Tectonic Reconstruction and Discussion: ...... 30 Caribbean plate and related blocks: ...... 30 Caribbean plate geometry and tectonic interactions: ...... 30 Caribbean plate interior: ...... 31 CV Intra-Oceanic arc, Oblique Collision and subduction flip related events: ...... 31

i

Collision of CV intra-oceanic arc and fore-arc basin opening: ...... 33 Tectonic predictions: ...... 35 Propagation of deformation: ...... 35 Modeled depositional events: ...... 37 Central Cordillera and Cretaceous metamorphosed passive margin sequences: ...... 37 Quebradagrande: ...... 38 Continuity of Colombian Andes and Colombian Massifts: ...... 38 Caribeana and Greater Antilles: ...... 38 Central America blocks: ...... 39 Panama arc block:...... 40 Sediment accumulation along Caribbean trailing edge: ...... 42 Western South America deformation front: ...... 43 Magdalena Valley basin: ...... 44 Eastern Cordillera:...... 45 Conclusions: ...... 46 Acknowledgments: ...... 47 References: ...... 48 Appendix 1. Supplementary Data: ...... 55

ii

Abstract: A quantitative kinematic model of northwestern Andes and Caribbean for the last 90 Ma illustrates how diachronous collision of Cabo de la Vela (CV) intra-oceanic arc triggers dextral transpressional deformation and extensive block rotation along the entire South America margin. Main implications of the reconstructed kinematics allow to constrain a ~65 Ma timing for subduction initiation of Caribbean plate beneath South America plate, and a gradual slab flattening process initiated since middle Eocene times (~47 Ma). Orogenic collapse of CV intra-oceanic arc and subduction flip process developed for Caribbean plate is interpreted to start at least since ~65 Ma in conjunction with subduction initiation. In addition, gradual orthogonalization of Caribbean plate slab starts at ~54 Ma, evolves in conjunction with Eocene ephemeral magmatism until ~47 Ma, and ends with Caribbean plate and Bahamas platform collision at ~45 Ma. Reconstructed geometry of Caribbean plate allows to infer an inner strike-slip lateral motion that evolves since ~79 Ma, and is prolonged to the east since ~58 Ma defining a displacement of > 700 km. The boundary established by this strike-slip motion segments completely Caribbean plate from Cuban segment at ~49 Ma, controls the opening of Cayman Trough pull- apart basin, and evolves progressively to the east until the end of the reconstruction, defining a final lateral displacement of ~1000 km.

iii

Introduction: The tectonic setting of northwestern South America and Caribbean is defined by the interaction of multiple tectonic plates: Caribbean, Atlantic, proto-Caribbean, Farallon (Cocos and Nazca), and South America. These interactions have controlled the regional tectonics, and paleogeography in the last 90 Ma. The understanding of these regional tectonics requires the integration of all the geological data in a consistent temporal, and kinematic context. However, this presents a complex scenario because part of data is diffuse, contradictory or is not available. In the Andes, events include the accretion of allochthonous terrains (Cardona et al., 2012; Weber et al., 2010), and the development of traspressional systems (Montes et al., 2003). In Caribbean, events include the evolution and closure of the Panama Isthmus (Montes et al., 2012), and the still contentious displacement of the Caribbean plate (Acton et al., 2000; Boschman et al., 2014; James, 2009; Pindell & Kennan, 2001) due to its incomplete record of magnetic anomalies (Ghosh et al., 1984). Kinematic tectonic reconstructions are a powerful tool to constrain geological data. These allow us to evaluate the spatial/temporal coherence of different geological elements, and can be used to establish predictions where there is no available data. This is achieved using plate motion circuits, based in spreading-rate calculations from oceanic magnetic anomalies (Müller, Sdrolias, Gaina, & Roest, 2008), and using hotspot absolute reference frames (Doubrovine et al., 2012; O’Neill et al., 2005; Torsvik et al., 2008). In this project, I present a palinspastic quntitative plate tectonic model that shows the evolution of Colombian Andes, and a part of Caribbean for the last 90 Ma, using the same absolute reference frame integrated by the EarthByte Group (Müller et al., 2008; O’Neill et al., 2005). This model honors all the data from available kinematic markers and regional piercing points (Bayona et al., 2008, 2012, 2013; García- Casco et al., 2008; Jiménez et al., 2014; Lamus Ochoa et al., 2013; Montes et al., 2005; Montes et al., 2003; Mora et al., 2006, 2010; Parra et al., 2009; Priem H.N.A et al., 1986; van Benthem et al., 2013; Zapata et al., 2014). This model follows Montes et al., 2010a kinematic where three main continental blocks were established: Maracaibo, Central Cordillera, and Eastern Cordillera blocks. This work is refined by updating the cartographic database, by using state-of-the-art tools such as GPlates (EarthByte group), and by using a technique known as "continuously closing polygons". This technique allows to dynamically evaluate kinematically- plausible scenarios, while tectonic blocks are allowed to change shape through time, somehow bending one of the foundations of , that plates are rigid (Cox & Hart, 1985). First, I review relevant literature for the Colombian Andes, the Caribbean. This is followed by the construction of a temporal schema that correlates tectonic events. Structural, stratigraphic, paleomagnetic and thermochronological data from the

1

Andes, are used to date and define the style of deformation adopted in the model. A palinspastic reconstruction of northwestern South America and the Caribbean is presented with snapshots of variable intervals to quantify deformation events, block displacements, and changing paleogeographic configurations. Methods: Euler rotation theorem (1776) establish that any motion over the surface of a sphere, can de described by a rotation about a pole that passes through the center of it. This pole is known as Euler pole, and it is the basis for tectonic reconstructions because movement of tectonic plates can be modeled as rotations in certain intervals of time. A consistent plate tectonic model has to include an adequate tree of hierarchies because movement of tectonic plates occur relative to a fixed reference frames (other plate, a hotspot, etc.). A rotation developed in a specific interval of time, and respect to a fixed frame of reference is defined as a finite rotation (Greiner, 1999). In consequence, a tectonic reconstruction is the result of the combination of multiple finite rotations developed in a consistent absolute reference frame that follows a tree of hierarchies. The software used to perform the model presented in this work is Gplates v.1.5. This free software was developed by the EarthByte Group, and is a powerful tool for developing palinspastic tectonic reconstructions. This because it facilitates the calculation of Euler poles that adjust to any movement desired. This is the possibility to calculate finite rotations that fit paleomagnetic, and other geological data. Also, Gplates let us to establish detailed absolute reference frames structuring consistent, and dynamic hierarchies between plates modeled. This let us rotate a plate respect to other during a defined interval of time, and change its reference in another interval. The adopted absolute reference frame in this model is a combination of two models proposed by Torsvik et al., 2008 for North America, and Müller et al., 2008; O’Neill et al., 2005 for South America. In this model, all Andean blocks were rotated respect to South American craton, Central America blocks respect to North America, and Caribbean plate according to Boschman et al., 2014. Finally, Caribbean and Panama blocks were rotated respect to Caribbean plate interior, and Caribbean plate interior was rotated according to kinematic coherence established by the model, and geological data. The closed dynamic polygons technique offers great advantages with respect to traditional rigid polygons because it makes it possible to adjust block overlapping with internal deformation. In practice, plate tectonics do not behave as rigid bodies because tectonic plates interior present very dynamic processes that include broad zones of diffuse tectonic activity, as the case of transpressional deformation developed in Colombian Andes (Molnar, 1988; Montes et al., 2003).This possibility allows reconstructions to show evolving tectonic blocks, as they grow (such an accretionary prism) shrinks (as occurs in a collisional margin), or get dismembered and stretched (as in pull-apart basins).The introduction of closed dynamic polygons

2 in this model was possible using the Topology Utility integrated in GPlates 1.5, which let us to create a closed polygon from the intersection of at least four lines, which are rotated independently. Regional Geological Setting: Major tectonic events in the evolution of the Colombian Andes, and Caribbean region since Late Cretaceous are reviewed in the following section. A comprehensive review of Caribbean-northern Andean tectonics is beyond the scope of this study, therefore this review is focused on key kinematic, and paleogeographic aspects of this margin (see Figure 1 for considered tectonic blocks). This review is followed by the description of piercing-points, or fundamental observations that help constrain the tectonic reconstruction.

Figure 1. Simplified Tectonic Map of Northwestern Andes and Caribbean with key tectonic blocks discussed below. (a) ECA: Eastern Cordillera (Axial), ECE: EC Eastern flank, ECW: EC Western flank, CC: Central Cordillera, MMB: Middle Magdalena basin, LMB: Lower Magdalena basin, UMB: Upper Magdalena basin, QG: Quebradagrande, CA: Central America, AT: Accreted terrains (Guajira allochthon and LMB allochthon from North to South), NPDB: North Panama deformed belt, GA: Guajira autochthonous, PR: Perijá Range, SNR: South Nicaraguan Rise, GM: Garzón Massif.

3

Caribbean Plate: Caribbean plate is composed by the Colombian and Venezuela basins, Beata ridge, and island arcs developed since Early and Late Cretaceous (Boschman et al., 2014). The evolution of Caribbean plate is dynamic because this plate seems to have travelled from a Pacific location to its current setting thousands of kilometers away (Boschman et al., 2014; Pindell et al., 1982, 1988, 1993, 2001, 2005, 2006, 2009). This results in a continuous change in Caribbean plate geometry related to intense deformation processes derived from its large displacement (Pindell & Kennan, 2001), and this is evidenced in the multiple structural elements mapped in Figure 2 (Burke et al., 1978). Shape and size changes are not only conditioned by deformation events, but also by accretion of sections of Caribbean plate in North and South America plates. The oceanic crust that composes Caribbean plate is anomalously thick ranging 15- 20 km, when an average expected value near of ~6-10 km (Burke et al., 1978). Deep sea drilling into this crust shows that it consist of basaltic rocks and fine-grained intrusives occurred over almost the entire Venezuelan basin and northern Colombian basin during Late Cretaceous (91-88Ma) in a large basalt flooding event (Donnelly, 1973). This flooding event gave rise to thick and buoyant oceanic plateau known as the Caribbean Large Igneous Province (CLIP) (Burke, 1978, 1978; Kerr et al., 1997). The emplacement of CLIP at 91-88 Ma establishes a temporal division in the components of Caribbean plate separating primary terrains formed in Early Cretaceous from those developed since Late Cretaceous (Boschman et al., 2014). CLIP formation obscures almost all the magnetic anomalies record, and with that the possibility to trace an absolute, and quantifiable displacement trajectory for this plate (van Benthem et al., 2013).

4

Figure 2. Structural evidence for intense deformation of Caribbean plate interior (taken from Burke et al., 1978). Caribbean plate displacement: The almost complete absence of oceanic magnetic anomalies in Caribbean plate makes it difficult to reconstruct its motion relative to neighboring plates (van Benthem et al., 2013). This presents a complex scenario because the estimations of past positions, relative motions, and plate velocities have to be done in an indirect way. As a consequence, these parameters have been derived from the data available for the surrounding plates (North and South America plates; van Benthem et al., 2013), and using kinematic constraints. Caribbean plate evolution is assumed to be defined by a displacement of thousands of kilometers (Pindell & Kennan, 2001), and a marked west to east strike slip motion since Eocene times. This motion is evidenced in the trastensional process that marks the opening and formation of the Cayman Trough, and the developing of Gonave microplate in conjunction with paleomagnetic data (Acton et al., 2000; Rosencrantz & Mann, 1991). Although Cayman Trough opening is related to Caribbean plate displacement, its opening rate is a measure of plate motion between Gonave microplate and North America (Cuban segment since Paleocene). This means that it is not a measure of relative motion between Caribbean and North America plate, and that it is not is not correct to derive a displacement rate of Caribbean plate respect to North America based in its opening (Rosencrantz & Mann, 1991). Thus, displacement rates have to be derived from kinematic models, motions of surrounding plates, and other

5 analytic techniques. Paleomagnetic data suggest that Caribbean plate was 5°-15° south of its current position at 80 Ma, it means that possibly was placed just over the Equator in Late Cretaceous (Acton et al., 2000). Mantle structure evidence: Tectonic reconstructions are useful in the understanding of mantle structure processes because let us to link crustal events to mantle dynamics. Many predictions and interpretations about mantle structure have been proposed in different places including the Caribbean region. In consequence, different authors have proposed diverse tectonic models that adjust to available data, but in some cases these ideas were not derived from quantitative data. Thus, mantle structure data derived from quantitative real sources is useful to quantify, and validate crustal models inferred from geological data or kinematic coherence (van Benthem et al., 2013). Recently, authors as van Benthem 2013 have contrasted these interpretations deriving mantle structure models from tomographic images using seismic wave velocities. In the Caribbean region, positive seismic anomalies are measured under the Lesser Antilles, Greater Antilles, and Puerto Rico suggesting that these anomalies are remnants of proto-Caribbean subducted plate lithosphere (van Benthem et al., 2013). These results shows that indeed the Atlantic lithosphere subduction model is consistent in time, and that inferred kinematic tectonic setting of Caribbean plate proposed by authors as (Pindell & Kennan, 2001) agree with quantitative data. In Figure 3, the slabs associated to Caribbean plate tectonic activity are shown as evidence for the developing of arc volcanism along the plate. To the northwest the subducted slab is responsible of the formation of Greater Antilles (Great Arc of the Caribbean (nGAC)). To the east and northeast the slab associated to Lesser Antilles (nLA, sLA), and in the southwest the slabs are associated with the consumption of Atlantic crust beneath South America (sGAC), and the change in subduction polarity (subduction flip) with the developing of SC slab.

6

Figure 3. Schematic Mantle structure of Caribbean Region (taken from van Benthem et al., 2013). Black arrows denote tectonic limits. Abbreviations for the subducted slabs: sLA: Southern Lesser Antilles; nLA: northern Lesser Antilles; Puerto Rico; sGAC: southern Great Arc of the Caribbean; nGAC: northern Great Arc of the Caribbean; SC: South Caribbean. Caribeana: Proto-Caribbean subduction beneath Caribbean plate triggered the evolution of the Great Arc of the Caribbean. In Greater Antilles arc material intrudes many ophiolitic complexes, which overlie accretionary wedges of deformed sediments related to the collision of Caribbean plate with Bahamas platform (Boschman et al., 2014). These ophiolitic complexes are related to a single Mesozoic paleogeographic terrain named as Caribeana by García-Casco et al., 2008. Caribbeana was a sedimentary prism extended southeastward of the southeastern part of Maya block (García-Casco et al., 2008). This prism was subducted in Late Cretaceous-Early Paleocene beneath Caribbean plate (~76-70 Ma), and is the precursor of many metamorphic complexes as Samaná (Cuba), Puerto Rico, East Yucatán, etc (García-Casco et al., 2008). Caribeana is a marker for the attenuation or interruption of Cretaceous volcanism in Caribbean plate, derived from Pro-Caribbean plate consumption. (García-Casco et al., 2008). Central America (Chortis, Siuna, Panama Blocks): Central America is a region composed of different tectonic blocks with a diverse origin and history. Since Pangea break-up in Jurassic times (~200 Ma), the development of Proto-Caribbean, and the appearance of Caribbean plate in Early Cretaceous the Caribbean region has suffered intense tectonic processes resulting in a complex and dynamic paleogeography (James Pindell & Kennan, 2001). Central America blocks are a principal key in the understanding of Caribbean history by its

7 role in the final separation of Pacific and Atlantic oceans, and the connection between North America and South America lands after 200 Ma of isolation. Central America Land Bridge can be divided in (north to south) Chortis, south Chortis, Siuna, and Panamá blocks (Chorotega, Panama, Chocó). Below, details of Central America blocks will be discussed in order establish the considerations relevant for the model. Chortis block has a crystalline Paleozoic basement that possibly was a part of North America prior to the Cenozoic (Boschman et al., 2014) before Proto-Caribbean openning. Chortis block is delimited to the north by the strike-slip Motagua zone (Rogers et al., 2007) that can be prolonged to the east in the Cayman Trough. South Chortis block in contrast with northern Chortis block has no pre-paleozoic basement, and it is interpreted to be a part of an accreted island arc (Rogers et al., 2007). Following the model of Boschman et al., 2014 southern Chortis will be fixed to north Chortis in this model because uncertainty timing of accretion. The Siuna block is composed of volcanics, serpentinized peridotite, ultramafic cumulates, and carbonate-rich sediments (Boschman et al., 2014). This block is interpreted to be an Early Cretaceous island arc, therefore a part of the Great Arc of Caribbean accreted to Chortis block in Late Cretaceous times (~85 Ma; Rogers et al., 2007). The kinematics associated to Siuna block will be not detailed in this model because are restricted almost completely to the Early and Late Cretaceous interval. The last part of Central America is the Panama blocks. These are interpreted as volcanic arc bodies developed since the Late Cretaceous to Neogene onto CLIP basalts (Montes et al., 2012). Geochemical data shows that at ~75 Ma a proto-arc was formed, represented by dykes and lava flows similar to CLIP, indicating initiation of subduction process (Boschman et al., 2014). The kinematics associated to Panama blocks are fundamental in the understanding of Caribbean tectonics because define the age of the final closure of the Central American seaway with the global biotic, and environmental repercussions related to it (Montes et al., 2012). Traditionally, the accepted model for Panama Isthmus Closure was sustained by molecular phylogenies, stratigraphic relations, paleobathymetry, marine paleosalinity, and paleotemperatures. In the basis of the initiation of the Great American Biological Interchange (GABI) was dated in 4-3 Ma (Coates & Stallard, 2013). Paleomagnetic analysis in the Panamá blocks shows that the original arc underwent a segmentation process triggered by individual vertical axis rotation of blocks. Rotations correspond to a counterclockwise vergence of high (70.9°±6.7°) and moderate magnitude (40°±4.1°, 56.2°±11.1°) during 38-28 Ma along an east- west trending fault (Rio Gatun fault), and clockwise rotations o central blocks during 28- 25 Ma (Figure 4; Montes et al., 2012).

8

Paleomagnetic data presented by Montes et al., 2012 is integrated in this model, to show temporal deformation and geometry of Panama blocks as a whole. Also the initiation of subduction and appearance of Panama arc in the model is presented at 67 Ma following the model of Montes et al., 2010a, and radiometric ages along the entire isthmus presented by Montes et al., 2012. Fore-Arc sedimentary units developed in western Chortis and Siuna blocks are a part of a fore-arc sliver that contains the Nicaragua basin, where extension and quantitative data was recorded by volcanics from 25 Ma to present (Boschman et al., 2014; Garza et al., 2012). Using this data this block was integrated in the reconstruction following the geometries proposed by Boschman et al., 2014.

Figure 4. Tectonic reconstruction of Panamá blocks since 38 Ma. In each one of the stages events as counterclockwise rotation in 38-28 is evidence, and counterclock rotation since 28. Paleomagnetic components and rotations magnitudes are presented. (taken from (Montes et al., 2012)).

9

Cayman Trough: The Cayman Trough is defined as a pull-apart basin resulting from strike slip motion between Caribbean plate and Cuban segment since ~49.4 Ma (Leroy et al., 2000). The Cayman Trough is the only place in Caribbean plate realm that contains magnetic anomalies; this means that quantitative measurements suppose valuable data for the development of tectonic models. It is thought that Cayman Through aperture is a measurement of North America and Caribbean plate relative displacement, but it really is a measurement of the relative displacement between North America plate and Gonave microplate (Rosencrantz & Mann, 1991). Caribbean plate displacements rates derived from Cayman Trough aperture are not valid, but allow to constrain consistent values based on the fact that Caribbean-North America relative motion is distributed across faults located close to the spreading axis. Thus, Caribbean-North America motion value must be greater than Cayman Trough spreading rate (Rosencrantz & Mann, 1991). Great Arc of the Caribbean: The concept of Great Arc of the Caribbean was introduced by (Burke, 1988), and refers to the volcanic arc material derived from the subduction of proto-Caribbean oceanic crust in all the perimeter of Caribbean plate, and developed in Early Cretaceous times (Boschman et al., 2014). These volcanic bodies are the precursors of Cuba segment, Chortis, and Siuna blocks. Aves Ridge, Lesser Antilles and Eastern Caribbean Basins: The Aves Ridge is a remnant island arc thought to be a part of the Great Arc of the Caribbean (Burke, 1988) with magmatic ages ranging from 88 to 59 Ma (Neill et al., 2011). This volcanic ridge consist of light rare earth element-enriched granitoids, calc-alkaline island arc basaltic andesites and metabasalts (Neill et al., 2011). Ages of active volcanism from Aves Ridge arc suggest that it was developed apart from the primary terrains grouped as the Great Arc of the Caribbean during Early Cretaceous times. The Grenada basin is a much debated oceanic basin. There are many models that explain its origin and formation. The most accepted hypothesis is that both Tobago and Grenada basins were a fore-arc associated to Aves Ridge (Aitken et al., 2011). This model is based in seismic, and on-land geological studies. The single fore-arc proposed continued to extend and widen during the Paleocene and middle Eocene (~66-45 Ma). Roll-back of the Proto-Caribbean slab beneath Caribbean plate caused eastward migration of volcanism 50-250 km from Aves ridge to Lesser Antilles arc, dividing the former fore-arc into Tobago and Grenada Basins (~38 Ma; Boschman et al., 2014). The Lesser Antilles arc is a magmatic body developed as a result of Atlantic oceanic crust subduction beneath Caribbean Plate. Ages of volcanic activity for this arc range from 38 Ma to present (Briden et al., 1979). Initiation of Lesser

10

Antilles arc volcanism marks the division of the Aves Ridge fore-arc into Tobago and Grenada basins. Cuba and Hispaniola: Cuba and Hispaniola islands are a part of the Greater Antilles and were derived from the Great Arc of the Caribbean (Boschman et al., 2014). Kinematics of both islands are linked to Caribbean plate until the collision with Bahamas platform in early Eocene (~45 Ma; Gordon et al., 1997; Meyerhoff & Hatten, 1968). The Greater Antilles exhibit a complex geological history derived from its multiple interactions with terrains of distinct origins and ages of formation that includes the Caribeana Mesozoic prism in Late Cretaceous times (~79-70 Ma), and the Bahamas platform (García-Casco et al., 2008). It means that the Cuban segment records a great part of Caribbean evolution history since the aperture of Proto-Caribbean Ocean, the appearance of Caribbean plate (~135 Ma; Boschman et al., 2014), and the subsequent subduction process of Proto-Caribbean crust beneath Caribbean plate resulting in the formation of the Great Arc of the Caribbean in Early Cretaceous times. Detailed data about Cuban segment (Cuba island, and Puerto Rico) is found in the paper of (Boschman et al., 2014), and other works from García-Casco et al., 2008, García-Casco et al., 2002, and many authors that discuss the geological constrains, and implications of the distinct heterogeneous units found in Greater Antilles. Therefore, the data presented below will discuss only the relevant data for this tectonic model, including some generalities that validate accepted hypothesis. Cuba exposes a fold-thrust belt derived from North America/Proto-Caribbean overlain by a Caribbean plate ophiolitic complex, and a volcanic arc sequence related to Proto-Caribbean subduction (Boschman et al., 2014). U/Pb ages date volcanic arc magmatism in Cuba in the interval 133-80 Ma, which coincides with the collision of Cuba and Caribeana (García-Casco et al., 2008; Stanek et al., 2009). Cuba fore-arc is almost missing as it is expected to have at least 166±60 km wide, (van Hinsbergen et al., 2009) and may be the result of a subduction erosion in the Cretaceous to Paleocene interval (~90-65 Ma). Hispaniola Island is located on the transform plate boundary between North America and Caribbean plates (Boschman et al., 2014). This boundary defines the limit in which Caribbean plate was segmented, and a part of it assimilated to North American plate after its collision with Bahamas Platform. The northern part of the island contains metasedimentary rocks suggested to be a part of the Bahamas carbonate platform overlain by an ophiolitic complex and arc material from Lower Cretaceous (Great Arc of the Caribbean) to Middle Eocene ages (Boschman et al., 2014). Above a subduction erosion process was discussed for Cuban island based in the fact of almost a complete inexistence of Cuban fore-arc, the same occurred in Hispaniola, and the evidence is found in a mélange. This contains blocks originated from Caribbean and Proto-Caribbean plate (Escuder-Viruete et al., 2011). In the Hispaniola island is found the Samaná peninsula, which means that exist also

11

Caribeana fragments there (García-Casco et al., 2008). Collision of this island with Bahamas Platform is interpreted to be Late-Middle Eocene, similar to Cuba (García- Casco et al., 2008). An Intraoceanic Arc: The history of the northwestern Andes is dominated by intense tectonic activity controlled by marked and continuous deformational events since Late Cretaceous time. These events are developed in a compressional environment, and are triggered by the accretion of allochthonous oceanic terranes. Accretionary terrains possess volcanic arc affinity, and were derived from the magmatism related to the consumption of Proto-Caribbean crust. It is thought that oceanic terrains accreted to South America during Late Cretaceous were CLIP fragments (plateau basalts; Vallejo et al., 2006; Villagómez et al., 2011). However, provenance studies and detrital zircons age populations from San Jacinto deformed belt presented by (Cardona et al., 2012) show that these terrains belong to an intra-oceanic arc active during 88 and 73 Ma (Figure 5). Provenance studies in the northwestern Andes shows that intra-oceanic arc remnants are present along the entire northwestern South American margin. This suggests that volcano-plutonic sequences found in the Western Cordillera, SNSM, Guajira (Cabo de la Vela), Venezuela, and Leeward Antilles have a common source of origin, at least for the allochthonous components of SNSM and Guajira (Cardona et al., 2012; Priem H.N.A et al., 1986; Weber et al., 2009; Zapata et al., 2014). Once collision took place, subduction flip followed, caused orogenic collapse, and basin opening (57-40 Ma). This results in the Caribbean plate subduction beneath South America Plate (Figure 5; Cardona et al., 2010, 2012). Below, terrains assumed to have a contribution of Late Cretaceous intra-oceanic arc will be discussed. Leeward Antilles and Venezuela sedimentary basins: The Leeward Antilles (Bonaire, Aruba, and Curacao) are a group of islands located in the northern part of Venezuela that include a series of juxtaposed Upper Cretaceous intra-oceanic arc and plateau remnants (van der Lelij et al., 2010; Zapata et al., 2014). Arc material present in Leeward Antilles has been related to Late Cretaceous intra-oceanic arc found in Ecuadorian Andes, Colombian Western Cordillera, Cabo de la Vela (Guajira, ), and SNSM suggesting a distinct origin from Great Arc of Caribbean and CLIP magmatic event (Cardona et al., 2010; Vallejo et al., 2009; Villagómez et al., 2011; Weber et al., 2009; Wright & Wyld, 2011). Paleomagnetic data from arc volcanics suggest a clockwise vertical-axis rotation of 90° for Aruba and Bonaire blocks respect to South America based in samples from Washikemba Formation on Bonaire. This data is agrees with the tectonic model because calculated poles are similar to those of Guajira, and northern Venezuela blocks suggesting the migration of the Late Cretaceous intra-oceanic arc along

12

South America-Caribbean margin (Stearns et al., 1982). Additionally, the earliest time of start of sedimentation is ~50 Ma, estimated using U-Th/He thermochronology data implying that a land connection between South American massifs and Leeward Antilles existed in late Eocene times (Zapata et al., 2014). Displacement between South America and Caribbean plate resulted in the formation of an extensive pull- apart basin system in northern South America since Eocene times. This shearing process created the space of accommodation for northern Venezuela sedimentary basins (Paraguana, Falcon, Los Monjes). Guajira and Sierra Nevada de Santa Marta: The northern margin of is formed by a group of accreted metamorphosed and non-metamorphosed volcano-plutonic Mesozoic oceanic and sedimentary rocks (Weber et al., 2010). The continuity of Andean Cordillera is interrupted to a series of isolated uplifted crystalline fragments limited by major strike-slip systems. These uplifted massifs include the Santa Marta massif and the Guajira serranías (Simaura, Jarara and Macuira), and are a part of Maracaibo microplate in conjunction with Serranía del Perija and Merida Andes (Cardona et al., 2011; Duque-Caro, 1979; Montes et al., 2010b). Also, both massifs can be divided in three lithological belts that in general terms represent an allochthonous, and autochthonous section (Figure 6; Cardona et al., 2010, 2011). There are several metamorphic and crystalline bodies found in Guajira and Sierra Nevada massifs, and are interpreted to record the tectonic interactions between South America and Caribbean plates since the Late Cretaceous. They include the Cretaceous Etpana- Jarara Formations, the Cabo de la Vela ultramafic complex (CVC), the Eocene Parashi granitoid, and the Eocene Santa Marta batolith (Cardona et al., 2010, 2011, 2014; Weber et al., 2009, 2010). U/Pb detrital zircon analysis carried out in Etpana formation and in metamorphic bulders from Cenozoic conglomerates present similar age populations, suggesting a common source and paleogeography. Ages of 1624 and 1315 Ma (Meso- Paleoproterozoic) and 630, 580, 547 Ma (Neoproterozoic) indicate that South American margin is a major source for sedimentary protoliths in conjunction with 90- 70 Ma ages from an allochthonous intra-oceanic arc (Weber et al., 2010). This means that Eptana Formation was developed adjacent to South American craton and not part of Late Cretaceous allochthonous intra-oceanic arc (Figure 7). The Jarara Formation was developed in the same sedimentary basin as Eptana formation, meaning that they are paleographically linked (Weber et al., 2009).

13

Figure 5. Tectonic stages recorded by the San Jacinto deformed belt. A. 76–73 intra‐oceanic arc growth and approach to the continental margin. B. 70–65 arc-continent collision. C. Basin formation due to orogen collapse. D. Basin filling and northeastern along strike transport. (taken from (Cardona et al., 2012a)

14

The Parashi stock, derived from ephemeral Eocene magmatism, records the final oblique convergence stage between South America and Caribbean plates (Weber et al., 2010). This granitoid includes a major quartzdiorite body with andesite to dacite dykes and mafic enclaves (Cardona et al., 2014). Zircon U/Pb and K/Ar geochronology analysis estimates and age of 51-47 Ma. Figure 6C shows that the Stock of Parashi (in red) intrudes Late Cretaceous sequences of intra-oceanic arc proposed by Cardona et al., 2012 and Weber et al., 2010. U-Th/(He) zircon and apatite thermochronology from granitoids in Santa Marta massif shows a major exhumation event in late Eocene times (45-40 Ma), late Oligocene (25 Ma), and Miocene (15 Ma). In contrast, Guajira only records Late Eocene to Early Oligocene (35-25 Ma) exhumation events suggesting that Guajira was isolated from Sierra Nevada de Santa Marta and Andean Chain by transtensional tectonics since post- Eocene times (Cardona et al., 2011).

Figure 6. Schematic division of allochthonous and autochtonous sections in SNSM and Guajira. Green part correspond to Late Cretaceous intra-oceanic arc accreted terranes. Pink and Orange part correspond to Precambrian and Mesozoic cratonic units respectively .(taken from (Cardona et al., 2011). Eocene plutonism in the Colombian Andes, derived from Late Cretaceous oblique convergence between Caribbean and South America plates, is exposed in the Central Cordillera, SNSM, and Guajira. However, this magmatic activity had a short duration (<10 Ma), and was followed by a hiatus associated to block rotation and extensive deformation in all the South American margin that resulted in an orthogonal plate convergence, and strike-slip motion between Caribbean and South America (Bayona et al., 2012 ; Cardona et al., 2014; Lamus et al., 2013; Macellari, 1995; Montes et al., 2005, 2010). Cardona et al., 2014 proposes that oblique

15 convergence triggered transpressional (Montes et al., 2003) and transtensional events in northern Andes causing exhumation, and disruption of Guajira region and an eastward tectonic transport of blocks along the South American margin (Montes et al., 2005).

Figure 7. Four stage model for the evolution of the Caribbean South American margin for Late Crecaceous to Eocene times. 1) In Late Crecaceous a Caribbean intra-oceanic arc approaches (obliquely) to South American margin. 2) Cabo de la Vela Complex (CVC) is formed in a back-arc system. Large arrows indicate plate movement, small arrows indicate the transpressive component in the fore-arc due to the oblique approach of the arc”. In 83-71 Ma and post-71 Ma, 3) an arc- continent collision between Caribbean and the South America plates takes place, and sediments are incorporated into the subduction system. 4) Intrusion Parashi stock defines the final stage of the model (arc pluton in 47-51 Ma; figure and interpretation taken from Weber et al., 2010).

16

Palinspastic restorations shows that Guajira and Santa Marta massifs once were a part of a continuous Andean chain during Late Cretaceous-Early Paleogene interval, but Cenozoic tectonics fragmented this unity (Cardona et al., 2011; Montes et al., 2005, 2010b). SNSM current position implies an extensional process in Plato-San Jorge basins, a shortening event in Cesar-Rancheria basin (between Santa Marta massif and Serrania del Perijá), and the emplacement of Cerrejón Thrust sheet that results in the fragmentation of Cordillera Central-Santa Marta massif original unity. Cesar-Ranchería basin preserves an Upper Cretaceous to Paleogene clastic wedge, that records a Paleocene unroofing process triggered by collision with Caribbean deformation front induced by Late Cretaceous intra-oceanic arc (~40 Ma; (Bayona et al., 2007; Cardona et al., 2012). This collision caused the opening of accommodation space for the Cerrejon Formation (Montes, 2010). Early Eocene, middle Miocene and late Oligocene unconformities have been reported in the western part of Serranía del Perijá, suggest that deformation started in early Eocene- late Oligocene implying shortening in Cesar-Rancheria basin (Montes et al., 2010b). Basement of Plato-San Jorge basin in extended continental crust, and is demonstrated by geophysical data (Montes et al., 2010b). Basement grabens in Plato-San Jorge basin were mapped by Cerón et al., 2007. A few paleomagnetic studies have been performed in the Sierra Nevada massif and Maracaibo realm, most of them reporting a clockwise rotation (Figure 8). With this in mind Montes et al., 2010b proposed a model that solves the situation presented before, and consist of a clockwise rotation of the Santa Marta massif to best fit the aperture of graben areas defined in Plato-San Jorge blocks. Results of this model present two possible scenarios: a rotation of 30°, and a rotation of 23°, starting in middle to late Eocene times (Montes et al., 2010b). The model of 30° clockwise rotation respect to fixed South America is adopted in this reconstruction. Western Cordillera, Ecuadorian Andes and Tumaco Fore-Arc: Western Cordillera is interpreted to be an allochthonous terrain that belong to a Late Cretaceous intra-oceanic arc (Cardona et al., 2012; Weber et al., 2010). Accreted terrains consist of middle to Late Cretaceous picrites, basalts and dolerites. The geochemistry of these bodies is consistent with an oceanic plateau environment, and yield radiometric ages of ~90 Ma (Kerr et al., 1997). Basaltic lavas and hyaloclastites of the Pallatanga unit, and ultramafic cumulates and gabbros from San Juan unit represent fragments of Late Cretaceous intra-oceanic arc accreted to South America in Ecuadorian Western Cordillera. Radiometric ages from these rocks are in the interval 92-88 Ma, and with ages of 85.5 ± 1.4 Ma for Pujilí granite, make possible a correlation with Western Cordillera rocks (Vallejo et al., 2006, 2009). In this tectonic model, previously discussed intra-oceanic arc related terrains that include: Leeward Antilles, Western Cordillera, Ecuadorian Andes allochthonous Guajira, and SNSM are assumed to be a unique original unit in Late Cretaceous

17 times that records continuous deformation, and is fragmented in the afore mentioned individual blocks during its accretion with South America margin. Cenozoic fore-arc and intra-arc sedimentary basins developed in southwestern Colombia over Late Cretaceous oceanic terrains (Tumaco and Cauca-Patía respectively, (Echeverri et al., 2015; Kerr et al., 1997). The Neogene stratigraphic record from Tumaco fore-arc basin consists of Eocene to middle Miocene sequence of ~1000 m of deltaic sandstones and minor sandstones deposited on the top of a volcano-sedimentary substrate (Borrero et al., 2012).This sequence is overlain by sequences resulting from other sedimentation processes, but these are no relevant to the reconstruction because its spatial distribution is defined by Eocene to middle Eocene Tumaco fore-arc.

Figure 8. Paleomagnetic data in SNSM and Maracaibo suggesting clockwise rotation of tectonic blocks. Data from many authors is presented. Abbreviations for faults are taken from Montes et al., 2010b. Cif, Cimitarra fault; Pf, ; Tf, Tigre fault; Sbf, Santa Marta-Bucaramanga fault; Rf, Romeral fault system; Mf, Morón fault system. Figure taken from (Montes et al., 2010b). Contemporaneous intra-arc basin as mentioned by (Echeverri et al., 2015), in Cauca-Patía area includes a ~2500 m Miocene record controlled by transitional to dominantly continental siliciclastic sedimentation. The spatial distribution of this basin is not illustrated in the reconstruction because the intention of model is to

18 present the spatial distribution of accreted oceanic terrains, and displaced tectonic blocks. Central Cordillera, Arquía and Quebradagrande complexes: Autochthonous continental crust of Central Cordillera is exposed to the east of Romeral fault system and west of Otú-Pericos fault (Villagómez et al., 2011). The Tahami terrain groups these crustal fragments, and consist of Paleozoic gneisses of the Puqui and La Miel units that are in non-conformable contact with overlying metasedimentary and metamorphosed igneous rocks of the Cajamarca complex (Restrepo & Toussaint, 1988). Cajamarca sequences are intruded by Jurassic, calc- alkaline, I-type granitoids of the Ibagué Batolith (150-140 Ma), and Tahami rocks by calc-alkaline, dioritic-granitic Late Cretaceous Antioqueño Batolith (88-83 Ma; Ibañez-Mejia et al., 2007). Also, continental arc granites from Sonsón batolith cross- cut Antioqueño batolith. Late Cretaceous magmatic rocks are exposed in the Central Cordillera in the Antioqueño batolith yielding ages of 94-72 Ma, and intruding Late Permian to Triassic deformed granitoids, additional to volcanic arc magmatism associated to allochthonous accretionary terrains (Villagómez et al., 2010). Also, a Late Cretaceous contribution from Central Cordillera is not excluded in San Jacinto deformed belt conglomerates (Cardona et al., 2012). This means that even when intra-oceanic arc derived terrains were the major source for San Jacinto deformed belt Eocene conglomerates, Late Cretaceous magmatic units from the Central Cordillera probably were as well. Cardona et al., 2012 suggests that this limited input of continental sediments input is related to a south paleogeographic position or the Antioqueño batolith as suggested provenance studies. This implies a more meridional paleogeographic position of all Central Cordillera blocks (Bayona et al., 2012; Lamus et al., 2012). Tahami terrain rocks are separated from Quebradagrande complex igneous and sedimentary rocks by the San Jerónimo fault (Villagómez et al., 2011). Quebradagrande complex is in faulted contact with lawsonite-glaucophane schists of Arquía Complex (Restrepo & Toussaint, 1988) across the Silvia-Pijao fault dated at 120-60 Ma (Bustamante, 2008). Kinematics of Quebradagrande and Arquía complexes are linked, and are assumed to be related to the accretion of a Late Cretaceous intra-oceanic arc (Cardona et al., 2012; Villagómez et al., 2011; Weber et al., 2010) in conjunction with Central Cordillera terrains. Eastern Cordillera: The Eastern Cordillera is a bivergent fold and thrust orogen that overthrust Magdalena valley and Llanos basin. The deformation style that defines contractional structures in the Eastern Cordillera was derived from anisotropies developed in an ancient Mesozoic extensional structure (Mora et al., 2006). The inversion of Mesozoic extensional structures in Northern Andes has controlled the dispersal of

19 detritus since ~72 Ma. This age is consistent with the accretion of allochthonous oceanic terrains to South America (Bayona et al., 2013; Cardona et al., 2012). Additionally, tilting of the Central Cordillera (~72-66 Ma) favored reverse reactivation of the western border of Cretaceous extensional basin, deposited in Mesozoic structures, along La Salina fault (Bayona et al., 2013; Parra et al., 2009). Bayona et al., 2013 discusses the role of Caribbean plate oblique convergence in this complex scenario, and shows how this is related to eastward migration of deformation and dynamic evolution of Colombian Andes depocenters. The Piedras-Girardot fold belt, located between Central and Eastern Cordilleras, is a dextral transpressional system that accommodates oblique contraction through north and northwest-trending segments of faults, dextral strike-slip along northeast- trending segments, and extension along east-trending faults (Montes et al., 2003). This is an example of deformation styles that dominate the Eastern Cordillera, and the geometry of structures associated to transpressional deformation. Montes et al., 2003 exposes that oblique convergence vectors imposed by Caribbean plate have to be distributed in a diffuse zone of deformation, and not in focalized spots. This means that deformation style and geometry that rules Piedras-Girardot fold belt is an expression of a regional transpresional system that governs tectonic block interactions in Eastern and Central Cordilleras (see Figure 9). Contractional forces are derived from Central Cordillera that acts like an indenter propagating deformation to Eastern Cordillera (Montes et al., 2003). Provenance from Guaduas Syncline presents three deformational events since Maastrichtian to middle Miocene. Central Cordillera tilting and fault reactivation in Eastern Cordillera (70-49 Ma). Intra-basin blocks deformation (52-40 Ma). Increasing of deformation in Eastern Cordillera western flank (Lamus et al., 2013). Stratigraphic studies along syntectonic basins of Eastern Cordillera are in agreement with Guadas Syncline results (Bayona et al., 2008). Thermochronological data from Eastern Cordillera reveal diachronous exhumation associated with Cenozoic contractional deformation (Parra et al., 2009). Low- temperature data over apatite-zircon crystals, and inverse modeling of fission-tracks shows a close correspondence in the timing and style of deformation along western and eastern flanks of the Eastern Cordillera (Mora et al., 2010). Post- middle Eocene to Holocene deformation progressively shifted toward east and western flanks, affecting marginal thrust systems in Magdalena Valley and Llanos foreland basins with a subsequent inversion of Mesozoic structures and exhumation of other structures (Mora et al., 2010).

20

Figure 9. Tectonic map of part of Eastern and Central Cordilleras. Light stippled pattern indicates elevations greater than 500 m above sea level. Heavy stippled pattern indicates elevations greater than 1000 m above sea level. AF = Ataco fault; BF = Bituima fault; CaF =Calarma fault; CF = Cambao fault; GS = Guaduas syncline; HF = Honda fault; IF = Ibague fault; MF = Magdalena fault; PF = Palestina fault; RFZ =Romeral fault zone. Square in map shows location and geometry of Piedras-Girardot fold belt. Analoguos structures can be seen in all Eastern Cordillera (taken from Montes et al., 2003).

Paleomagnetic studies in Cretaceous to Miocene sediments of Eastern Cordillera present non-systematic tectonic rotations, including non-rotated areas respect to fixed South America plate, and local clockwise rotations in specific areas due to right- lateral shear (Cúcuta-Merida Andes; Jiménez et al., 2014). It is found that Eastern Cordillera inverted a NNE oriented Mesozoic rift, in the Miocene-Recent deformation event, that defined the structural trends followed by posterior contractional structures (Bayona et al., 2013; Jiménez et al., 2014; Mora et al., 2010). Magnetic lineation trends NNE to NW, meaning that its oblique respect to Eastern Cordillera orogenic trend. A 35°±9° clockwise rotation is observed in Cucuta zone that reflects late Cenozoic dextral strike-slip motion occurring in faults parallel to Bocono fault system that involves kinematics of Merida Andes (Jiménez et al., 2014). Jiménez et al., 2014 present two possible scenarios for the evolution of the contractional structures of Eastern Cordillera: oblique slip/transpressive deformation and strain partitioning. The first suggest an oblique reactivation of previous extensional faults, and the second imply dip-slip motion on both flanks and strike-slip motion in the axial zone (see Figure 10; Figure 11).

21

Figure 10. Fold belt schema for Eastern Cordillera. Valid for two models proposed by Jiménez et al., 2014 are valid here. (Taken and modified from Jiménez et al., 2014). Seismicity in the Colombian Andes is associated to Nazca plate subduction, and is diffused over a wide area that includes a high seismicity area known as the Bucaramanga nest (Chiarabba et al., 2016). Chiarabba et al., 2016 relocate about 5000 earthqueakes to produce a 3-D velocity model of the region to define the geometry of subducted Nazca slab beneath Colombian Andes (Figure 12). Anomalous clustered seismicity found in Bucaramanga Nest, and seismicity in Eastern Cordillera is associated to a flat subduction geometry. Flat subducted slab is related to a massive dehydration and eclogitization of thickened oceanic crust, but this hypothesis cannot be demonstrated with seismic tomography due to problems in data resolution (Chiarabba et al., 2016; Prieto et al., 2012). Therefore, additional evidence from the spatial distribution of volcanic bodies is required to demonstrate the effect of an eventual flat slab. Chiarabba et al., 2016 exposes that flat subduction is consistent with abrupt absence of volcanism in latitudes >5°. Flattening of Nazca subducted slab was dated in ~10 Ma based in its average convergence rate of ~80 mm/yr. This time is consistent with the last deformational event of Eastern Cordillera in Miocene to Recent times (Chiarabba et al., 2016; Jiménez et al., 2014). This means that flattening of subducion slab correlates with Eastern Cordillera compressive event in 15-10 Ma, favored by pre-existing extensional structures (Figure 12; Espurt et al., 2008). This correlation was reported as well during the collision of Panama arc-northern South America (Vargas & Mann, 2013), establishing a relationship between basin inversion and crustal thickening that results in flat slab patterns (Chiarabba et al., 2016).

22

Figure 11. Two scenarios proposed by Jiménez et al., 2014.IF = Ibague Fault, BSF = Bucaramanga- Santa Marta Fault, S-PF = Soapága Pesca Fault, BF = Boyacá Fault, and SBF = Salina-Bituima Fault. (Taken from Jiménez et al., 2014). All presented data from Eastern Cordillera is in agreement that oblique convergence of Caribbean plate triggered transpressional deformation in northern Andes, favored by reactivation of Mesozoic extensional structures. This data includes structural, seismic and paleomagnetic derived models that agree with the dextral transpressional model. Stratigraphic studies in Guaduas Syncline are in agreement with termochronology results that dates initiation of deformation in both flanks of Eastern Cordillera since middle Eocene to Holocene. Also, information is consistent with indenter-role of Central Cordillera in propagation of deformation to Eastern Cordillera.

23

Figure 12. Evolution of Nazca-South America plates subduction system. (taken from Chiarabba et al., 2016). Piercing Points: Caribbean plate displacement constrains were discussed before, including the problematic absence of magnetic anomalies (van Benthem et al., 2013) and the inferred motion derived from adjacent plates: North and South America. Here, two snap-shots at: 50, 21, 15 Ma are presented as reference for paleo-positions of the Caribbean plate, inferred from provenance analysis, geochronology, paleomagnetic, and biological constrains of Panama arc. These piercing points allow to place fixed points in the total reconstructed trajectory for Caribbean plate, conferring kinematic spatial/temporal coherence for intermediate and final stages of the model. Eocene times (~50 Ma): There are two main piercing points to constrain the tectonic configuration for Eocene times: 1) the Parashi granitoid, that intrudes metamorphosed Cretaceous passive margin sequences; and 2) stranded Precambrian clasts in the island of Bonaire. The first piercing point is the Parashi granitoid. This granitoid developed as a result of ephemeral Eocene magmatism and constitutes a marker from Caribbean plate oblique subduction under northernmost South America. This granitoid indicates that the leading edge of the Caribbean plate was located at the current location of the Alta Guajira ranges during Eocene times (Figure 13; Cardona et al., 2014). The string of granitoids shown in Figure 13 (from Santa Barbara to Parashi) suggest that the tectonic configuration along this margin was similar at this time. Paleo-latitudes from Cardona et al., 2014 were preserved in this reconstruction but paleo-longitudes changed a little due to kinematic constrains derived from northern Caribbean data that includes fragmentation of Cuban segment, and aperture of Cayman Trough

24

(Boschman et al., 2014; Leroy et al., 2000). Additionally, this piercing point establishes the gradual initiation of west-east strike-slip displacement of Caribbean plate respect to South America that will result in the opening of multiple pull-apart basins developed since Oligocene times (Falcon, Bonaire; Macellari, 1995). The second piercing point are high-grade metamorphic clasts found in conglomerates of the Soebi Blanco Formation (Bonaire basin; Priem et al., 1986). The presence of these clasts in Bonaire has been interpreted as evidence of Eocene fans that were sourced in nearby basement massifs of northern South America. Therefore, the Leeward Antilles arc were dextrally displaced from northern Colombia ~300 km to its current position. Evidence of this are detrital zircons recovered from the sandy matrix of the Soebi Formation that show populations of 1200-1000 Ma, 950-750 Ma, and 300-200 Ma, which confirm that Boinaire and the Lesser Antilles were attached to South American basement massifs (SABM). These massifs are composed by Paraguana, Falcon, Maracaibo, Guajira, Perija, and Santa Marta blocks (Zapata et al., 2014). Early Miocene (~21 Ma): An early Miocene land connection between Chortis block and Panama arc block is necessary for terrestrial mammal communication between the Canal Basin in the Istmus of Panama and North America. There is ample fossil evidence (Kirby & MacFadden, 2005; Rincon et al., 2013, 2015) that there was free genetic flow between North American faunas and Canal Basin faunas in late Oligocene and early Miocene times, including the first reported North America monkey fossil found in Las Cascadas Fm. (Canal basin; Bloch et al., 2016). There is, however, no record of any North American terrestrial mammals in South America, also suggesting that there was no communication between southern Central America and South America at this time. A land connection implies tectonic closure, so that there is no longer any piece of oceanic crust that would geodynamically be isostatically compensated, and thus hundreds or thousands of meters below sea-level. Middle Miocene (15-13 Ma): Closure of Panama Isthmus is dated in 15-13 Ma from geochronological and provenance studies. Evidence comes from Panamanian Eocene (59-42 Ma) detrital zircons, found in middle Miocene-Oligocene fluvial and shallow marine strata from northern Andes (Montes et al., 2015). Also, Middle Miocene magmatism is absent in this area, and additional zircons, derived from old Andean terrains, are found yielding Late Precambrian, Permo-Triassic and Late Cretaceous ages (Montes et al., 2015).

25

Figure 13. Tectonic reconstruction of northern Andes in Eocene times (~50 Ma). Location of Caribbean plate and Panama arc respect to South America is illustrated. Additionally Parashi stock, and distinct sedimentary formations are shown. (taken from Cardona et al., 2014). This evidence allowed to Montes et al., 2015 infer a fluvial land connection between Panama and northern Andes, and propose a complete closure of Central American seaway, considering that the only possible source for Eocene detrital zircons founded in northern Andes fluvial systems are Late Cretaceous-Eocene magmatic rocks from Panama arc (see Figure 14 for details about sampling area and fluvial system geometry). This fluvial communication seems to be corroborated by analysis of multiple clade migration times (Bacon et al., 2015), suggesting that main events of migration between South and North America took place in waves starting at approximately 20 Ma. The reconstruction, therefore, closes any deep-water interoceanic (Pacific-Caribbean) gaps at 15 Ma.

26

Figure 14. Fluvial systems containing Eocene detrital zircon populations. (taken from Montes et al., 2015). Tectonic reconstruction of Andean blocks: Tectonic blocks defined for this model are based in the rigid blocks reconstruction of Montes et al., 2010a as previously mentioned. Also, kinematics derived from Piedras-Girardot fold belt are presented in Figure 15. Division of reconstructed tectonic blocks was made following fault systems exposed in northern Andes (Figure 16).

27

Figure 15. Tectonic reconstruction of Andean blocks from (Montes et al, 2005) in six different stages of geological time. (taken from Montes et al., 2005). Non-Palinspastic.

28

Figure 16. (a) Tectonic blocks by major tectonic domains at 0 Ma. EC: Eastern Cordillera, CC: Central Cordillera, WC: Western Cordillera, MMB: Middle Magdalena basin, LMB: Lower Magdalena basin, UMB: Upper Magdalena basin, QG: Quebradagrande, CA: Central America, Accreted Terrains: Guajira allochthon and LMB allochthon from North to South. (b) Dividing faults used in the model. GF: Guaicaramó fault, PF: Palestina fault, RF: Romeral fault system, YP: Yopal fault, BSF: Bucaramanga-Santa Marta fault, OF: Oca fault, CF: Cusiana fault, UF: Uribe fault, CAF: Caguan fault, BAF: Borde Amazónico fault, AF: Algeciras fault, LSF: La Salina fault, IB: Ibagué fault, PAF: Pajarito fault, HF: Honda fault, DPF: De Prado fault, ALF: Altamizal fault, SPF: Santiago Perez fault, MF: Macama fault, SJF: San Jerónimo fault, OTF: Otu fault, CPF: Cali-Patia fault, UF: Uramita fault, EF: Encarnacion fault, BF: Bocono fault system, SSF: San Sebastian fault, EPF: El Pilar fault, URF: Urica fault, LAF: La Victoria fault, AMF: Amurrapa fault, DSF: De Sardinata fault, RIF: Riecito fault, ABF: Arenas Blancas fault, BNF: Burro Negro fault, RGF: Rio Gatun fault, PTF: Perija-El Tigre faults, LPF: , SF: Simarúa fault.

29

Tectonic Reconstruction and Discussion: The presented palinspastic tectonic reconstruction was based in the non-dynamic closed blocks tectonic reconstruction of Montes et al., 2005, 2010a, and Boschman et al., 2014. Rotation file (Gplates) from Boschman et al., 2014 was used and adapted to the hybrid absolute reference frame adopted in this model. Below, snapshots in different time intervals will be presented to discuss model considerations, kinematic inferences and to show critical tectonic and deformation events that shaped the geography and paleography of northern Andes since Late Cretaceous times (90-0 Ma). During the development of the model and following the philosophy of this reconstruction, blocks that initiate tectonic activity will be segmented from South American craton, and will be treated as independent blocks. Segmented blocks will be integrated to principal tectonic domains (divided by faults) established by Montes et al., 2005, 2010a. Discussion of model is presented in five major tectonic domains: 1) Caribbean plate; 2) Cabo de la Vela arc; 3) Central Cordillera; 4) Caribbeana and Greater Antilles; 5) Caribbean plate trailing edge; 6) Western South America deformation front. Caribbean plate and related blocks: Caribbean Plate is treated in this kinematic model as the entity that drives deformation along the northwestern margin of South America. Propagation of deformation triggered by Caribbean plate oblique convergence is modeled since the beginning of the reconstruction at 90 Ma until its gradual orthogonalization in ~54- 45 Ma. Geometry of Caribbean plate is dynamic according to model philosophy, and was derived from paleogeographic positions and geometries of Greater Antilles (Boschman et al., 2014), Caribbean interior (Montes, 2010a; Boschman et al., 2014), Siuna block (Boschman et al., 2014) and Late Cretaceous intra-oceanic arc (Montes, 2010a). Pacific (Boschman et al., 2014; Pindell & Kennan, 2001) or Atlantic (James, 2009) origin of Caribbean plate cannot be discussed in this model because paleomagnetic data from Late Cretaceous is valid for both models (Acton et al., 2000).  Caribbean plate geometry and tectonic interactions: In the initial stages of the reconstruction internal deformation of Caribbean plate is not appreciable, but during early Paleocene-early Eocene (65-50 Ma; Figures 17 and 18) interval deformation of Caribbean interior is intense, resulting in a narrow displacement of developing Panama arc block respect to South American margin (Figure 18), according to paleomagnetic constrains (Cardona et al., 2014; Montes, 2010). In addition, extension inside Caribbean plate starts at 58 Ma between Aves Ridge and Hispaniola block, as a consequence of intense block rotation and displacement of accreted allochthonous terrains (Stearns et al., 1982). This results in a major segmentation of original Caribbean plate into Cuban segment, and Cayman Trough at 49 Ma (Leroy et al., 2000). Also, subduction of Caribbean plate beneath South America starts at least from 65 Ma (Cardona et al., 2010, 2012; Leal-

30

Mejía, 2011) originating Eocene magmatism, last magmatic pulse in Antioqueño batolith and a gradual change in convergence pattern from oblique to orthogonal (Bayona et al., 2012) that ends with the collision of Cuban segment with Bahamas platform in 45 Ma (Boschman et al., 2014; García-Casco et al., 2008). Later deformation event of Caribbean plate interior since ~44-21 Ma (Figures 19, 20, 21) is modeled to adjust the model to previously discussed piercing points.  Caribbean plate interior: Additional blocks attached to Caribbean plate as Nicaraguan rise, Aves Ridge volcanic arc, Aves Ridge fore-arc and Lesser Antilles arc were modeled using the geometries proposed by Boschman et al., 2014. Timing for magmatism in Aves Ridge is dated in the 88-60 Ma interval (Neill et al., 2011), and for Lesser Antilles arc at the 38-0 Ma interval (Briden et al., 1979). Growing of both Lesser Antilles and Aves Ridge arcs pattern is uniform, and in contrast to Panama arc evolution (Figure 19, 20, 21), is not developed through different volcanic centers that interconnects each other in time. Aves Ridge fore-arc block grows at a constant rate, and is segmented into Grenada and Tobago basins at 38 Ma, when Lesser Antilles arc starts to grow by Atlantic slab roll-back, and subsequent magmatism migration (Aitken et al., 2011; Boschman et al., 2014). Growing of Lesser Antilles arc between segmented Aves Ridge fore-arc basin is illustrated by reduction of Tobago and Grenada basins area (Figure 20). CV Intra-Oceanic arc, Oblique Collision and subduction flip related events: The Cabo de la Vela intra-oceanic arc is perhaps the most important tectonic feature in this tectonic reconstruction. It sweeps along the northwestern margin of South America, from south to north, diachronously colliding with this margin, metamorphosing Lower Cretaceous passive-margin sequences, and propagating deformation along the western margin of the South American craton. Where it encounters rift-sequences, those get reactivated and inverted (Bayona et al., 2013; Cardona et al., 2012; Montes et al., 2003; Weber et al., 2009). To the north, its collision causes vertical-axis block rotation, orogen-parallel extension and basin opening (Macellari, 1995; Stearns et al., 1982; Zapata et al., 2014). As the arc collides with the South American passive margin, a subduction flip takes place where ephemeral subduction of Caribbean flares up a short-lived magmatic episode that is soon put off and replaced by widespread deformation, exhumation and erosion in the northern Andes (Bayona et al., 2012; Cardona et al., 2012). This deformation- cooling-exhumation episode is the result of arrival of thick Caribbean crust to the subduction zone. The orogen that was built during this episode looses geodynamic support and collapses so that only sliver of it survive in a few places along the margin (Cardona et al., 2012; Dewey, 2005).

31

Figure 17. (a) Tectonic blocks by major tectonic domains at 65 Ma. EC: Eastern Cordillera, CC: Central Cordillera, WC: Western Cordillera, MMB: Middle Magdalena basin, LMB: Lower Magdalena basin, UMB: Upper Magdalena basin, QG: Quebradagrande, CA: Central America, Accreted Terrains: Guajira allochthon and LMB allochthon from North to South. (b) Dividing faults used in the model. GF: Guaicaramó fault, PF: Palestina fault, RF: Romeral fault system, YP: Yopal fault, BSF: Bucaramanga-Santa Marta fault, OF: Oca fault, CF: Cusiana fault, UF: Uribe fault, CAF: Caguan fault, BAF: Borde Amazónico fault, AF: Algeciras fault, LSF: La Salina fault, IB: Ibagué fault, PAF: Pajarito fault, HF: Honda fault, DPF: De Prado fault, ALF: Altamizal fault, SPF: Santiago Perez fault, MF: Macama fault, SJF: San Jerónimo fault, OTF: Otu fault, CPF: Cali-Patia fault, UF: Uramita fault, EF: Encarnacion fault, BF: Bocono fault system, SSF: San Sebastian fault, EPF: El Pilar fault, URF: Urica fault, LAF: La Victoria fault, AMF: Amurrapa fault, DSF: De Sardinata fault, RIF: Riecito fault, ABF: Arenas Blancas fault, BNF: Burro Negro fault, RGF: Rio Gatun fault, PTF: Perija-El Tigre faults, LPF: La Plata fault, SF: Simarúa fault.

32

 Collision of CV intra-oceanic arc and fore-arc basin opening: Reconstructed Cabo de la Vela (CV) intra-oceanic arc extends >1000 km along the major part of Caribbean plate eastern margin (Cardona et al., 2012; Weber et al., 2010; Montes, 2010a). This feature represents all the Cretaceous oceanic crust terrains found in the Western Cordillera, Leeward Antilles, Guajira, SNSM and Ecuadorian Andes (Figure 17; Cardona et al., 2011, 2012, 2014; Vallejo et al., 2009; Weber et al., 2009; Zapata et al., 2014). Original CV intra-oceanic arc is segmented diachronously in three different stages, due to extensional forces derived from diachronous collision and distinct rates of displacement between collided and non- collided sections. A first segmentation occurs at 67 Ma separating CV intra-oceanic arc in two segments, a second occurs at 57 Ma separating Western Cordillera block from Ecuadorian Andes block, and a last occurs at 33 Ma separating Leeward Antilles blocks from SNSM, Guajira and Lower Magdalena basin allochthonous part. Oblique collision of Cabo de la Vela intra-oceanic arc starts at 80 Ma for its southern part propagating deformation to Central Cordillera and Quebradagrande blocks. Diachronous collision of CV intra-oceanic arc results in northeastward migration and shortening in Andean blocks (transpression). Northern part of CV intra-oceanic arc collides with northern South America margin at 67 Ma and starts as well a dextral transpressional process followed by a west-east migration of ~300 km (Figures 19, 20 and 21, Priem et al., 1986; Zapata et al., 2014). CV northern part collision differs from its counterpart by triggering an intense clock-wise block rotation event developed at 40 Ma, with ~90° of magnitude (Stearns et al., 1982). The unstable and not isostatically compensated orogen resulted from CV intra- oceanic arc collision results in a subduction flip event that triggers Caribbean plate subduction beneath South America (Cardona et al., 2012; Dewey, 2005). Eocene ephemeral magmatism is derived from this subduction event that allows the evolution of Parashi stock (Cardona et al., 2014), and other granitic intrusives along the entire Eocene South American margin (~51-47 Ma; Bayona et al., 2012; Cardona et al., 2014). These events are recorded by the San Jacinto deformed belt block (Figure 18, 19), which grows over the collapsed orogen and deforms in the 57-28 Ma interval. In the model, sediment sources for San Jacinto block are limited to CV intra-oceanic arc, and a part from Central Cordillera (Cardona et al., 2012). Just over San Jacinto block, Sinú deformed belt block starts to grow as an accretionary wedge since 28 Ma (Kellog et al., 2005). This model divides these blocks to illustrate the two-sided wedge character of this basin developed by Cenozoic compressional shortening (Figure 21, 22; Kellog et al., 2005). The SNSM block is treated as a rigid indenter, and remains stable until 40 Ma, when starts to present a gradual clock-wise rotation that is propagated to adjacent Perija block and Maracaibo blocks, causing extension in subjacent Low Magdalena basin block. (Montes et al., 2010a; Stearns et al., 1982).

33

Figure 18. (a) Tectonic blocks by major tectonic domains at 57 Ma. EC: Eastern Cordillera, CC: Central Cordillera, WC: Western Cordillera, MMB: Middle Magdalena basin, LMB: Lower Magdalena basin, UMB: Upper Magdalena basin, QG: Quebradagrande, CA: Central America, Accreted Terrains: Guajira allochthon and LMB allochthon from North to South. (b) Dividing faults used in the model. GF: Guaicaramó fault, PF: Palestina fault, RF: Romeral fault system, YP: Yopal fault, BSF: Bucaramanga-Santa Marta fault, OF: Oca fault, CF: Cusiana fault, UF: Uribe fault, CAF: Caguan fault, BAF: Borde Amazónico fault, AF: Algeciras fault, LSF: La Salina fault, IB: Ibagué fault, PAF: Pajarito fault, HF: Honda fault, DPF: De Prado fault, ALF: Altamizal fault, SPF: Santiago Perez fault, MF: Macama fault, SJF: San Jerónimo fault, OTF: Otu fault, CPF: Cali-Patia fault, UF: Uramita fault, EF: Encarnacion fault, BF: Bocono fault system, SSF: San Sebastian fault, EPF: El Pilar fault, URF: Urica fault, LAF: La Victoria fault, AMF: Amurrapa fault, DSF: De Sardinata fault, RIF: Riecito fault, ABF: Arenas Blancas fault, BNF: Burro Negro fault, RGF: Rio Gatun fault, PTF: Perija-El Tigre faults, LPF: La Plata fault, SF: Simarúa fault.

34

This results in the reactivation of Bocono fault system and the segmentation- expulsion of Merida block from South American craton at 37 Ma (Figure 20). Guajira autochthonous blocks underwent a clock-wise rotation event as well, however, dextral transpressional deformation occurs along Oca fault since 40 Ma, triggering segmentation of Guajira blocks from Paraguana and Falcon basins at 28 Ma along Simarúa fault.  Tectonic predictions: Subduction of Caribbean plate beneath South America plate can be interpreted to be active at least during Paleocene times (~65Ma; See Figure 17). This constrain derives from the steady-state convergence of CV intra-oceanic allochthonous terrains, that limits Caribbean plate crust thrusting over South America margin, and forces it to be subducted. This is consistent with U/Pb zircon ages presented by Leal- Mejía, 2011 for Antioqueño Batolith that yields four magmatic pulses at 96-92 Ma, 89-82 Ma, 81-72 Ma and 60-58 Ma, and Paleocene-Eocene ages for Sonsón and El Bosque batholiths. This evidence suggest that there is a Paleocene magmatism. Additionally, primary magmatic age pulses of Antioqueño Batolith emplacement (post-Albian to pre-Paleocene ages; Leal-Mejía, 2011) present a complex scenario that is not solved in this model because northeastwards displacement of Central Cordillera block (that contain Antioqueño Batolith) would imply a northeast magmatism migration. Ephemeral Eocene magmatism is a key element in this model because subduction is not stopped (Figure 19). A hypothesis that explains this magmatic hiatus is flattening of Caribbean subduction maybe be related with Bucaramanga nest seismicity (Chiarabba et al., 2016; Prieto et al., 2012). Therefore, it is possible to think that this flattening process triggered by the change of convergence pattern between Caribbean and South America plates occurred gradually since the termination of short Eocene magmatism.  Propagation of deformation: Segmentation of Perija block from South American craton, and extension of Lower Magdalena basin block starts at 40 Ma in conjunction with César-Ranchería basin shortening (Cerón et al., 2007; Montes et al., 2010a). This derives in the emplacement of thrust sheets and the development of graben structures (Cerón et al., 2007; Montes et al., 2010a). The magnitude of these extensional and compressional processes is driven by model kinematics and the rotation rate of SNSM.

35

Figure 19. (a) Tectonic blocks by major tectonic domains at 45 Ma. EC: Eastern Cordillera, CC: Central Cordillera, WC: Western Cordillera, MMB: Middle Magdalena basin, LMB: Lower Magdalena basin, UMB: Upper Magdalena basin, QG: Quebradagrande, CA: Central America, Accreted Terrains: Guajira allochthon and LMB allochthon from North to South. (b) Dividing faults used in the model. GF: Guaicaramó fault, PF: Palestina fault, RF: Romeral fault system, YP: Yopal fault, BSF: Bucaramanga-Santa Marta fault, OF: Oca fault, CF: Cusiana fault, UF: Uribe fault, CAF: Caguan fault, BAF: Borde Amazónico fault, AF: Algeciras fault, LSF: La Salina fault, IB: Ibagué fault, PAF: Pajarito fault, HF: Honda fault, DPF: De Prado fault, ALF: Altamizal fault, SPF: Santiago Perez fault, MF: Macama fault, SJF: San Jerónimo fault, OTF: Otu fault, CPF: Cali-Patia fault, UF: Uramita fault, EF: Encarnacion fault, BF: Bocono fault system, SSF: San Sebastian fault, EPF: El Pilar fault, URF: Urica fault, LAF: La Victoria fault, AMF: Amurrapa fault, DSF: De Sardinata fault, RIF: Riecito fault, ABF: Arenas Blancas fault, BNF: Burro Negro fault, RGF: Rio Gatun fault, PTF: Perija-El Tigre faults, LPF: La Plata fault, SF: Simarúa fault.

36

Strike-slip motion between Caribbean and South America plates is initiated since 45 Ma (Figure 19), and results in basin opening by extension between CV intra-oceanic arc northern part and Guajira autochthonous blocks. As a result, Paraguana, Falcon, Los Monjes and South Caribbean deformed belt basins are developed in Caribbean- South America plate boundary. Paraguana and Falcon basins starts to be filled in Oligocene times (~33 Ma), but, this is not illustrated in this model because its space is already taken by subjacent cratonic sequences (Macellari, 1995). Segmentation of these developing basins and Guajira block occurs at 28 Ma, and transpressional deformation continues as basin filling occurs. Timing for Los Mojes and South Caribbean deformed belt blocks growing is derived from kinematic constrains and correspond to 19 Ma and 23 Ma respectively (Figure 21, 22).  Modeled depositional events: Bonaire basin is thought to be developed as a result of basin extension, and to be an offshore extension of Falcon basin (Macellari, 1995), however, this interpretation is not consistent with zircon chronology and time of deposition reported by Zapata et al., 2014. Therefore, Bonaire basin is modeled as an accretionary wedge developed over South American margin since 50 Ma (see Figure 19 for geometry), and continuously transpressed by adjacent blocks (Zapata et al., 2014). Tumaco fore-arc is developed during the 35-12 Ma interval (Echeverri et al., 2015). Posterior depositional events that overlain Tumaco block are not illustrated in this model because maximum extension of these sedimentary sequences is reached at 12 Ma (Figure 20, 21, 22; Echeverri et al., 2015). Central Cordillera and Cretaceous metamorphosed passive margin sequences: Central Cordillera block include autochthonous Paleozoic basement is represented by Puqui and La Miel units (Restrepo & Toussaint, 1988), and Jurassic to Late Cretaceous intrusives represented by Ibague and Antioqueño batoliths respectively. The paleogeographic position of Central Cordillera is modeled based in interpretations presented before (Bayona et al., 2012; Lamus Ochoa et al., 2012). Central Cordillera changes is size as transpressional deformation occurs integrating Ibague batolithand Palestina region at 79 Ma (Montes et al., 2010a). Since 79 Ma, Central Cordillera acts as an indenter that propagates deformation to South America margin. Transpressed Central Cordillera block deforms following a sigmoidal geometries (Figure 17) in its leading edge until its collision with Middle Magdalena basin and Lower Magdalena basin at 57, 50 Ma (Figure 18; Montes et al., 2003).

37

 Quebradagrande: Quebradagrande and Arquía blocks were linked together in a same tectonic because they shared a common paleogeography and deformational styles (Montes et al., 2010a). Kinematics of Quebradagrande block was derived from inferred displacement of Central Cordillera blocks, spatial distribution of strike-slip faults (San Jerónimo and Romeral fault system), and oblique convergence of Western Cordillera blocks based in Caribbean plate displacement until Upper Eocene. Dextral transpressional deformation of Quebradagrande block starts at 79 Ma with an associated northeastwards migration as occurs with Central Cordillera block, and remains until the end of the reconstruction, exposing lesser intense deformation since 57 Ma (Figure 18).  Continuity of Colombian Andes and Colombian Massifs: Continuity of Central and Eastern Cordilleras are interrupted to a series of isolated uplifted crystalline fragments limited by major strike-slip systems. These uplifted massifs correspond to SNSM, autochthonous Guajira, Perija and Merida blocks that were diachronously segmented from South American stable craton (Cardona et al., 2011; Duque-Caro, 1979; Montes et al., 2010b). Three sedimentary basin units limit presented blocks: 1) Cesar-Ranchería basin, 2) Lower Magdalena basin and, 3) Baja, Alta Guajira basins. Eptana and Jarara formations were developed just adjacent to South American margin, and are exposed in conglomeratic sequences found at Serrania de Jarara foothills (Cardona et al., 2011; Weber et al., 2010). These two formations were developed in a same sedimentary basin (Weber et al., 2009), and are treated as a single paleogeographic domain included in the Guajira block since 68 Ma (Figure 17). CV intra-oceanic arc allochthonous contribution to SNSM, Guajira and Lower Magdalena basin blocks is compressed in a single tectonic block that exposes how lateral continuity of original fragmented CV intra-oceanic arc extends along South America margin in conjunction with Leeward Antilles, Western Cordillera and Ecuadorian Andes (Figure 20 to see lateral continuity). Caribeana and Greater Antilles: Caribeana Mesozoic prism and Greater Antilles (Cuba and Puerto Rico) geometry, paleogeography, piercing points, and motion vectors are adopted from Boschman et al., 2014 model. Major piercings points are established at 45 Ma with the collision of Greater Antilles (Cuba) with Bahamas platform at 45 Ma, and the final subduction of Caribbeana prism beneath Caribbean plate at 76-70 Ma, considering the paleogeographic position for Caribeana at the southern part of Yucatán (García- Casco et al., 2008). Bahamas platform collision attaches Greater Antilles to North America plate (Figure 19). Additional features as Cuba fore-arc were adapted exactly from Boschman et al., 2014 and do not present any change of shape.

38

Caribbean plate trailing edge: Panama and Central America blocks: Caribbean plate trailing edge compress blocks diverse in its origin, age, and kinematics, triggered by multiple tectonic events that back from Mesozoic times with the Pangea breakthrough (~200 Ma; Boschman et al., 2014). Central America blocks refers to Mexico, Guerrero, Chortis and Siuna blocks. Kinematics of these blocks are not detailed because the interval of time covered by this model is restricted to 90-0 Ma, and the data for Mexico and Guerrero blocks is standard, and was derived from EarthByte group (Müller et al., 2008; O’Neill et al., 2005). In this reconstruction, these blocks have a secondary role until early Miocene representing a very important piercing point involved in mammal species exchange between North America and Canal basin (Figure 21; Bloch et al., 2016; Kirby & MacFadden, 2005; Rincon et al., 2013, 2015). Panama blocks are modeled as isolated volcanic centers developed since 65 Ma (Figure 17; Montes et al., 2012), changing its shape and size trough the time. Panama block establish the more important piercing point in this reconstruction because defines de Central America seaway closure at ~15 Ma, and the final separation of Pacific from Caribbean (Figure 22).  Central America blocks: Chortis and Siuna blocks geometries are adapted from Boschman et al., 2014. Siuna offshore and onshore blocks are modeled to be a part of Caribbean plate until 85 Ma, when are accreted to Chortis block and are integrated to Central America blocks behaving as a single tectonic block since this time. Therefore, distinction of Siuna and Chortis blocks is no longer useful, and this is illustrated since 69 Ma. Posterior interactions of theses blocks follow Boschman et al., 2014 model until during early Oligocene to early Miocene interval (30-21 Ma), when displacement rate of Chortis and Siuna blocks along dextral strike-slip Motagua fault zone increases to adjust the model to early Miocene piercing point that establishes a land connection between Panama arc block and Chortis & Siuna block (Figure 21; Bloch et al., 2016).

39

Figure 20. (a) Tectonic blocks by major tectonic domains at 38 Ma. EC: Eastern Cordillera, CC: Central Cordillera, WC: Western Cordillera, MMB: Middle Magdalena basin, LMB: Lower Magdalena basin, UMB: Upper Magdalena basin, QG: Quebradagrande, CA: Central America, Accreted Terrains: Guajira allochthon and LMB allochthon from North to South. (b) Dividing faults used in the model. GF: Guaicaramó fault, PF: Palestina fault, RF: Romeral fault system, YP: Yopal fault, BSF: Bucaramanga-Santa Marta fault, OF: Oca fault, CF: Cusiana fault, UF: Uribe fault, CAF: Caguan fault, BAF: Borde Amazónico fault, AF: Algeciras fault, LSF: La Salina fault, IB: Ibagué fault, PAF: Pajarito fault, HF: Honda fault, DPF: De Prado fault, ALF: Altamizal fault, SPF: Santiago Perez fault, MF: Macama fault, SJF: San Jerónimo fault, OTF: Otu fault, CPF: Cali-Patia fault, UF: Uramita fault, EF: Encarnacion fault, BF: Bocono fault system, SSF: San Sebastian fault, EPF: El Pilar fault, URF: Urica fault, LAF: La Victoria fault, AMF: Amurrapa fault, DSF: De Sardinata fault, RIF: Riecito fault, ABF: Arenas Blancas fault, BNF: Burro Negro fault, RGF: Rio Gatun fault, PTF: Perija-El Tigre faults, LPF: La Plata fault, SF: Simarúa fault.

40

 Panama arc block: Panama arc block evolves since 65 Ma using two isolated volcanic centers that grow in time and reach major part of its maximum extension at 50 Ma (Figure 17, 18; Montes et al., 2012; Montes et al., 2015). Location of volcanic centers correspond to older ages reported by Montes et al., 2012, and correspond to Proto-arc dykes derived from an evolving subduction system (Boschman et al., 2014). Volcanic centers of Panama eventually are connected at ~39 Ma (see Figure 20), when it exposes a flat shaped geometry parallel to Farallon plate subduction margin. Volcanism in Panama arc remains active until late Miocene times, but it is restricted uniquely to its western part, after a magmatic hiatus in its eastern part since 38 Ma (Figure 20; Montes et al., 2012). Inner segmentation of Panama arc blocks occurs at 38 Ma, by Rio Gatun strike-slip fault activation, and a counter-clockwise block rotation. Segmentation of Panama arc block is not illustrated in this model, instead deformation associated to strike-slip motion is shown (Montes et al., 2012). Clockwise rotation of Panama arc central blocks develops between 28-25 Ma (Montes et al., 2012), and is evidenced with a mild left-lateral drift of Panama arc block that results in Chortis-Panama closure at 21 Ma (Figure 21). Finally, orthogonal convergence between Caribbean and South America plates results in Central America seaway closure at 15 Ma (Figure 22; Montes et al., 2015). The time for the complete closure is subject to interpretation because data presented by Montes et al., 2015 demonstrates a land connection based in the existence of fluvial channels, but its connection can be narrow and its extension can vary.

41

Figure 21. (a) Tectonic blocks by major tectonic domains at 21 Ma. EC: Eastern Cordillera, CC: Central Cordillera, WC: Western Cordillera, MMB: Middle Magdalena basin, LMB: Lower Magdalena basin, UMB: Upper Magdalena basin, QG: Quebradagrande, CA: Central America, Accreted Terrains: Guajira allochthon and LMB allochthon from North to South. (b) Dividing faults used in the model. GF: Guaicaramó fault, PF: Palestina fault, RF: Romeral fault system, YP: Yopal fault, BSF: Bucaramanga-Santa Marta fault, OF: Oca fault, CF: Cusiana fault, UF: Uribe fault, CAF: Caguan fault, BAF: Borde Amazónico fault, AF: Algeciras fault, LSF: La Salina fault, IB: Ibagué fault, PAF: Pajarito fault, HF: Honda fault, DPF: De Prado fault, ALF: Altamizal fault, SPF: Santiago Perez fault, MF: Macama fault, SJF: San Jerónimo fault, OTF: Otu fault, CPF: Cali-Patia fault, UF: Uramita fault, EF: Encarnacion fault, BF: Bocono fault system, SSF: San Sebastian fault, EPF: El Pilar fault, URF: Urica fault, LAF: La Victoria fault, AMF: Amurrapa fault, DSF: De Sardinata fault, RIF: Riecito fault, ABF: Arenas Blancas fault, BNF: Burro Negro fault, RGF: Rio Gatun fault, PTF: Perija-El Tigre faults, LPF: La Plata fault, SF: Simarúa fault.

42

 Sediment accumulation along Caribbean trailing edge: Caribbean plate trailing edge presents two sedimentation events since Miocene times: 1) North Panama deformed belt, and 2) Nicaraguan fore-arc sliver. Sedimentation above Panama arc starts at 19 Ma, in North Panama deformed belt fore-arc basin (Pindell & Kennan, 2009), between the available space Caribbean plate interior, Panama arc block, and Sinú accetionary wedge (Montes et al., 2010a, 2012). Adjacent to Chortis & Siuna onshore block, a fore-arc basin that contains Nicaragua basin start to develop at 25 Ma (Boschman et al., 2014; Garza et al., 2012). Western South America deformation front: Deformation is propagated to South America margin during diachronous accretion and segmentation of CV intra-oceanic arc at 80 (southern part) and 67 Ma (northern part). CV intra-oceanic arc collision derives in extensive vertical axis rotation of tectonic blocks (Figure 20, 21), and in continuous dextral transpressional deformation propagated from southwest to northeast, reactivating older Mesozoic extensional structures, and displacing blocks along faults. In this model, Western South America deformation front is restricted to deformation propagated by southern segment of CV intra-oceanic arc trough Central Cordillera indenter. Displaced blocks are contracted against the margin delimited by Bucaramanga-Santa Marta fault to northwest, and Cusiana-Yopal faults to northeast (Figure 17). Deformation front first reaches Central Cordillera, Quebradagrande, and Upper Magdalena basin blocks in the 80-75 Ma interval, resulting in the segmentation of these tectonic blocks. Next, major segmentation occurs at 74 Ma with the separation of Garzon Massif, Eastern Cordillera (axial and both flanks), and Middle Magdalena basin blocks. Last major segmentation event occurs at 65 Ma, propagating previous reactivated faults (74 Ma) to its maximum extension, and defines all the Andean blocks involved in the rest of the reconstruction, with the exception of eastern flank of Eastern Cordillera.

43

Figure 22. (a) Tectonic blocks by major tectonic domains at 15 Ma. EC: Eastern Cordillera, CC: Central Cordillera, WC: Western Cordillera, MMB: Middle Magdalena basin, LMB: Lower Magdalena basin, UMB: Upper Magdalena basin, QG: Quebradagrande, CA: Central America, Accreted Terrains: Guajira allochthon and LMB allochthon from North to South. (b) Dividing faults used in the model. GF: Guaicaramó fault, PF: Palestina fault, RF: Romeral fault system, YP: Yopal fault, BSF: Bucaramanga-Santa Marta fault, OF: Oca fault, CF: Cusiana fault, UF: Uribe fault, CAF: Caguan fault, BAF: Borde Amazónico fault, AF: Algeciras fault, LSF: La Salina fault, IB: Ibagué fault, PAF: Pajarito fault, HF: Honda fault, DPF: De Prado fault, ALF: Altamizal fault, SPF: Santiago Perez fault, MF: Macama fault, SJF: San Jerónimo fault, OTF: Otu fault, CPF: Cali-Patia fault, UF: Uramita fault, EF: Encarnacion fault, BF: Bocono fault system, SSF: San Sebastian fault, EPF: El Pilar fault, URF: Urica fault, LAF: La Victoria fault, AMF: Amurrapa fault, DSF: De Sardinata fault, RIF: Riecito fault, ABF: Arenas Blancas fault, BNF: Burro Negro fault, RGF: Rio Gatun fault, PTF: Perija-El Tigre faults, LPF: La Plata fault, SF: Simarúa fault.

44

 Magdalena Valley basin: After 74 Ma, Upper Magdalena basin block will not present further segmentations, but will present transpressional deformation along Ibague, Altamizal and Del Prado faults. The only massif illustrated in this model correspond to Garzón massif that exposes Paleozoic basement that since its segmentation will not present internal deformation by transpression, and its interaction with other tectonic blocks will be limited to strike-slip motion. In this reconstruction, block is divided in two distinct blocks due to they are delimitated by faults. Secondary Middle Magdalena valley block vanish completely by deformation induced by Central Cordillera indenter at 49 Ma.  Eastern Cordillera: Eastern Cordillera is divided in three different blocks: Western flank (Magdalena thrust front), axial zone, and Eastern flank. Crustal tilting of Central Cordillera is dated at ~72-66 Ma, and fault reactivation in the 70-49 Ma according to Guaduas syncline data (Lamus Ochoa et al., 2013). Therefore, reactivation of Mesozoic extensional structures coincides with a primary and secondary segmentation of Eastern Cordillera axial zone and western flank in 74 Ma, and 65 Ma proposed in this model. Reactivation of La Salina fault and Honda thrust marks the fragmentation of the mentioned cratonic segments, and separates western flank from axial zone (Eastern Cordillera) and Middle Magdalena Valley from western flank respectively (Parra et al., 2009). Collision of Central Cordillera indenter with Middle Magdalena basin occurred in 49 Ma results in a marked shortening of Eastern Cordillera Western flank and axial zone in the Nuevo Mundo syncline (La Paz and Esmeraldas Formations), and in Los Cobardes Anticline (Caballero et al., 2010). Bocono fault system starts at 44 Ma in Eastern Cordillera Axial zone, and is propagated to Merida block at 37 Ma. Thermochronological data shows that deformation in both flanks were initiated since post-Eocene times, but intense tectonic activity that occurred in Western flank since 4 Ma defines a conceptual segmentation of this block adopted in this model. Evidence of this can be found along Llanos foothills, in the middle Eocene to Pliocene Leon and Guayabo formations that corresponds to the major shortening phase of Eastern Cordillera (Bayona et al., 2013; Parra et al., 2010).

45

Conclusions: This work presents a detailed review of the history of northwestern Andes and Caribbean region that is integrated in a spatial/temporal consistent kinematic context that adjust to kinematic markers and available geological data for the past 90 Ma. This model consist of a combination of rigid blocks models proposed by Boschman et al., 2014 for Caribbean, and Montes et al., 2010a for northwestern Andes with the introduction of the closed dynamic polygons feature, that support changes in shape and geometry in bocks due to tectonic activity. The following predictions and conclusions are derived from the reconstruction: 1. Integration of Montes et al., 2010a and Boschman et al., 2014 models is contradictory for the 90-80 Ma interval because motion of Caribbean plate seems to have a double component to northwest for Caribbean plate leading edge (Greater Antilles) and northeast for CV intra-oceanic arc. 2. CV intra-oceanic arc evolution ages in conjunction with Antioqueño Batolith pulses present a spatial inconsistency for this reconstruction that would force a magmatism migration of Paleocene volcanism along with Caribbean plate. Primary pulses of Antioqueño Batolith prior to 90 Ma present a complex scenario in which Atlantic lithosphere have to be subducted beneath South America plate during approaching of Caribbean plate. 3. Strike-slip motion that defines the transform limit between Caribbean plate principal segment and Cayman Trough could be initiated in the interior of Caribbean plate since 79 Ma and prolonged to the east since ~58 Ma defining a lateral displacement >700 km and a final major segmentation of Caribbean plate. 4. Subduction of Caribbean plate beneath South America plate has to start at least from ~65 Ma because Caribbean plate colliding margin cannot stop its motion after CV intra-oceanic arc orogen emplacement. This also implies that posterior orogenic collapse and subduction flip process that trigger subduction could start prior to ~57 Ma and was gradual because San Jacinto deformed belt basin timing initiation for sedimentation is dated at ~57 Ma. 5. Caribbean plate convergence oblique pattern changes gradually to a pure orthogonal pattern since ~54 Ma until the final timing from Eocene magmatism (~47 Ma), and collision of Caribbean plate with Bahamas platform at 45 Ma. 6. Eocene ephemeral magmatism had a short duration of ~10 Ma, but subduction cannot be stopped in order to preserve spatial coherence of the model implying a gradual flattening process in Caribbean plate slab since middle Eocene times prior to reported ~15 Ma from seismic velocity models.

46

Acknowledgments: I thank gratefully to my thesis director Dr. Camilo Montes by his advisement during the development of this work, and his innumerable lessons about geology and its life style, also, his very valuable contributions that helped me to improve this manuscript. I thank immensely to my family, especially to my mother Dora for her accompaniment during all my educational, and thesis development process, for her aid in all these years of study and her continuous support and lessons during all my life. I am very grateful to my brother Julian by his constructive critics and his aid during all my educational process. I thank to my father Jarbin by his advices about life and his aid during my career. Also, I completely thank to my best friends Aura, Marco, and Manuel by the very good and invaluable moments shared during all these years, their continuous accompaniment and advising that made a better and happey person. Finally, I thank to all other good friends that contributed to my educational process and my personal improving as Camilo, Nataly, Helman, Carolina, Juliana, Andrea and Catalina.

47

References:  Acton, G., Galbrun, B., & King, J. (2000). 9 . PALEOLATITUDE OF THE CARIBBEAN PLATE SINCE THE LATE CRETACEOUS 1. North, 165.  Aitken, T., Mann, P., Escalona, A., & Christeson, G. L. (2011). Evolution of the Grenada and Tobago basins and implications for arc migration. Marine and Petroleum Geology, 28(1), 235–258.  Bacon, C. D., Silvestro, D., Jaramillo, C., Smith, B. T., Chakrabarty, P., & Antonelli, A. (2015). Biological evidence supports an early and complex emergence of the Isthmus of Panama. Proceedings of the National Academy of Sciences, 112(19), 6110–6115.  Bayona, G., Cardona, A., Jaramillo, C., Montes, C., & Mora, A. (2012). Effects of Paleogene Subduction on Basin Segmentation and Oil Migration in NW South America. In Nuevas Fronteras II.  Bayona, G., Cardona, A., Jaramillo, C., Mora, A., Montes, C., Caballero, V., … Valencia, V. (2013). Onset of fault reactivation in the Eastern Cordillera of Colombia and proximal Llanos Basin; response to Caribbean–South American convergence in early Palaeogene time. Geological Society, London, Special Publications, 377(1), 285–314.  Bayona, G., Cardona, A., Jaramillo, C., Mora, A., Montes, C., Valencia, V., … Ibañez-Mejia, M. (2012). Early Paleogene magmatism in the northern Andes: Insights on the effects of Oceanic Plateau–continent convergence. Earth and Planetary Science Letters, 331–332, 97– 111.  Bayona, G., Cortés, M., Jaramillo, C., Ojeda, G., Aristizabal, J. J., & Reyes-Harker, A. (2008). An integrated analysis of an orogen–sedimentary basin pair: Latest Cretaceous–Cenozoic evolution of the linked Eastern Cordillera orogen and the Llanos foreland basin of Colombia. Geological Society of America Bulletin, 120(9-10), 1171–1197.  Bayona, G., Lamus Ochoa, F., Cardona, A., Jaramillo, C., & Montes, C. (2007). Procesos orogénicos del Paleoceno para la cuenca de Ranchería (Guajira, Colombia) y áreas adyacentes definidos por análisis de procedencia.  Bloch, J. I., Woodruff, E. D., Wood, A. R., Rincon, A. F., Harrington, A. R., Morgan, G. S., … MacFadden, B. J. (2016). First North American fossil monkey and early Miocene tropical biotic interchange. Nature, advance online publication.  Borrero, C., Pardo, A., Jaramillo, C. M., Osorio, J. A., Cardona, A., Flores, A., … Castillo, H. (2012). Tectonostratigraphy of the Cenozoic Tumaco forearc basin (Colombian Pacific) and its relationship with the northern Andes orogenic build up. Journal of South American Earth Sciences, 39, 75–92.  Boschman, L. M., van Hinsbergen, D. J. J., Torsvik, T. H., Spakman, W., & Pindell, J. L. (2014). Kinematic reconstruction of the Caribbean region since the Early Jurassic. Earth-Science Reviews, 138, 102–136.  Briden, J. C., Rex, D. C., Faller, A. M., & Tomblin, J. F. (1979). K-Ar Geochronology and Palaeomagnetism of Volcanic Rocks in the Lesser Antilles Island Arc. Philosophical Transactions of the Royal Society of London A: Mathematical, Physical and Engineering Sciences, 291(1383), 485–528.  Burke, K. (1988). Tectonic Evolution of the Caribbean. Annual Review of Earth and Planetary Sciences, 16(1), 201–230.  Burke, K., Fox, P. J., & Şengör, A. M. C. (1978). Buoyant ocean floor and the evolution of the Caribbean. Journal of Geophysical Research: Solid Earth, 83(B8), 3949–3954.

48

 Bustamante, A. (2008). Geobarometria, geoquímica, geocronologia e evolução tectônica das rochas de fácies xisto azul nas áreas de Jambaló (Cauca) e Barragán (Valle del Cauca), Colômbia. Universidad de São Paulo, Colombia.  Caballero, V., Parra, M., Bohorquez, M., & Roberto, A. (2010). LEVANTAMIENTO DE LA CORDILLERA ORIENTAL DE COLOMBIA DURANTE EL EOCENO TARDÍO - OLIGOCENO TEMPRANO: PROVENIENCIA SEDIMENTARIA EN EL SINCLINAL DE NUEVO MUNDO, CUENCA VALLE MEDIO DEL MAGDALENA. Boletin de Geología, 32(1), 45–77.  Cardona, A., Montes, C., Ayala, C., Bustamante, C., Hoyos, N., Montenegro, O., … Zapata, S. (2012). From arc-continent collision to continuous convergence, clues from Paleogene conglomerates along the southern Caribbean–South America plate boundary. Tectonophysics, 580, 58–87.  Cardona, A., Valencia, V. A., Bayona, G., Duque, J., Ducea, M., Gehrels, G., … Ruiz, J. (2011). Early-subduction-related in the northern Andes: Turonian to Eocene magmatic and provenance record in the Santa Marta Massif and Rancheria Basin, northern Colombia. Terra Nova, 23(1), 26–34.  Cardona, A., Valencia, V., Bustamante, C., García-Casco, A., Ojeda, G., Ruiz, J., … Weber, M. (2010). Tectonomagmatic setting and provenance of the Santa Marta Schists, northern Colombia: Insights on the growth and approach of Cretaceous Caribbean oceanic terranes to the South American continent. Journal of South American Earth Sciences, 29(4), 784–804.  Cardona, A., Weber, M., Valencia, V., Bustamante, C., Montes, C., Cordani, U., & Muñoz, C. M. (2014). Geochronology and geochemistry of the Parashi granitoid, NE Colombia: Tectonic implication of short-lived Early Eocene plutonism along the SE Caribbean margin. Journal of South American Earth Sciences, 50, 75–92.  Cerón, J., F., Kellog, J., & Ojeda, G. (2007). BASEMENT CONFIGURATION OF THE NORTHWESTERN SOUTH AMERICA - CARIBBEAN MARGIN FROM RECENT GEOPHYSICAL DATA.  Chiarabba, C., De Gori, P., Faccenna, C., Speranza, F., Seccia, D., Dionicio, V., & Prieto, G. A. (2016). Subduction system and flat slab beneath the Eastern Cordillera of Colombia. Geochemistry, Geophysics, Geosystems, 17(1), 16–27.  Coates, A. G., & Stallard, R. F. (2013). How old is the Isthmus of Panama? Bulletin of Marine Science, 89(4), 801–813.  Cox, A., & Hart, R. . (1985). Plate Tectonics: How it Works. Retrieved April 6, 2016, from https://www.researchgate.net/publication/241942759_Plate_Tectonics_How_it_Works Dewey, J. F. (2005). Orogeny can be very short. Proceedings of the National Academy of Sciences, 102(43), 15286–15293.  Donnelly, T. W. (1973). BASALTS AND DOLERITES OF LATE CRETACEOUS AGE FROM THE CENTRAL CARIBBEAN.  Doubrovine, P. V., Steinberger, B., & Torsvik, T. H. (2012). Absolute plate motions in a reference frame defined by moving hot spots in the Pacific, Atlantic, and Indian oceans. Journal of Geophysical Research: Solid Earth, 117(B9), B09101.  Duque-Caro, H. (1979). Major Structural Elements and Evolution of Northwestern Colombia: Small Basin Margins.  xEcheverri, S., Cardona, A., Pardo, A., Monsalve, G., Valencia, V. A., Borrero, C., … López, S. (2015). Regional provenance from southwestern Colombia fore-arc and intra-arc basins: implications for Middle to Late Miocene orogeny in the Northern Andes. Terra Nova, 27(5), 356–363.

49

 Escuder-Viruete, J., Friedman, R., Castillo-Carrión, M., Jabites, J., & Pérez-Estaún, A. (2011). 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(3–4), 483–504.  Espurt, N., Funiciello, F., Martinod, J., Guillaume, B., Regard, V., Faccenna, C., & Brusset, S. (2008). Flat subduction dynamics and deformation of the South American plate: Insights from analog modeling. Tectonics, 27(3), TC3011.  García-Casco, A., Iturralde-Vinent, M. A., & 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.  Garza, R. S. M., van Hinsbergen, D. J. J., Rogers, R. D., Ganerød, M., & Dekkers, M. J. (2012). The Padre Miguel Ignimbrite Suite, central Honduras: Paleomagnetism, geochronology, and tectonic implications. Tectonophysics, 574–575, 144–157.  Ghosh, N., Hall, S. A., & Casey, J. F. (1984). Seafloor spreading magnetic anomalies in the Venezuelan Basin. Geological Society of America Memoirs, 162, 65–80.  Gordon, M., Cáceres, D., & Flores, R. (1997). Cenozoic tectonic history of the North America- Caribbean plate boundary zone in western Cuba. JOURNAL OF GEOPHYSICAL RESEARCH, 102, 10–055.  Greiner, B. (1999). Euler rotations in plate-tectonic reconstructions. Computers & Geosciences, 25(3), 209–216.  Ibañez-Mejia, M., Tassinari, C., & Jaramillo-Mejia, J. . (2007). U–Pb zircon ages of the “Antioquian Batholith”: geochronological constraints of Late Cretaceous magmatism in the Central Andes of Colombia (Vol. 11). Presented at the 11th Colombian Geological Congress, extended abstracts.  James, K. H. (2009). In situ origin of the Caribbean: discussion of data. Geological Society, London, Special Publications, 328(1), 77–125.  Jiménez, G., Speranza, F., Faccenna, C., Bayona, G., & Mora, A. (2014). Paleomagnetism and magnetic fabric of the Eastern Cordillera of Colombia: Evidence for oblique convergence and nonrotational reactivation of a Mesozoic intracontinental rift. Tectonics, 33(11), 2014TC003532.  Kellog, J., Toto, R., & Cerón, J., F. (2005). Structure and tectonics of the Sinu-San Jacinto accretionary prism in northern Colombia. (pp. 1–10). Presented at the In X Congreso Colombiano de Geologia.  Kerr, A. C., Marriner, G. F., Tarney, J., Nivia, A., Saunders, A. D., Thirlwall, M. F., & Sinton, C. W. (1997). Cretaceous Basaltic Terranes in Western Columbia: Elemental, Chronological and Sr–Nd Isotopic Constraints on Petrogenesis. Journal of Petrology, 38(6), 677–702.  Kerr, A. C., Tarney, J., Marriner, G. F., Nivia, A., & Saunders, A. D. (1997). The Caribbean- Colombian Cretaceous Igneous Province: The Internal Anatomy of an Oceanic Plateau. In J. J.honey & M. F. Coffin (Eds.), Large Igneous Provinces: Continental, Oceanic, and Planetary Flood Volcanism (pp. 123–144). American Geophysical Union.  Kirby, M. X., & MacFadden, B. (2005). Was southern Central America an archipelago or a peninsula in the middle Miocene? A test using land-mammal body size. Palaeogeography, Palaeoclimatology, Palaeoecology, 228(3–4), 193–202.  Lamus Ochoa, F., Bayona, G., Cardona, A., Montenegro, O., Mora, A., & Caballero, V. (2012). Aproximación a la proveniencia sedimentaria de unidades Paleógenas de la Cordillera Oriental mediante el uso de minerales pesados. In Yacimientos I (poster).

50

 Lamus Ochoa, F., Bayona, G., Cardona, A., & Mora, A. (2013). PROCEDENCIA DE LAS UNIDADES CENOZOICAS DEL SINCLINAL DE GUADUAS: IMPLICACIÓN EN LA EVOLUCIÓN TECTÓNICA DEL SUR DEL VALLE MEDIO DEL MAGDALENA Y ORÓGENOS ADYACENTES. Boletin de Geología, 35(1), 17–42.  Leal-Mejía, H. (2011). Phanerozoic Gold Metallogeny in the Colombian Andes: A tectono- magmatic approach (Doctoral dissertation). Universidad de Barcelona.  Leroy, S., Mauffret, A., Patriat, P., & 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.  Macellari, C. E. (1995). Cenozoic Sedimentation and Tectonics of theSouthwestern Caribbean Pull-Apart Basin, Venezuela and Colombia, 757–780.  Meyerhoff, A. A., & Hatten, C. W. (1968). Diapiric Structures in Central Cuba, 153, 315–357.  Molnar, P. (1988). Continental tectonics in the aftermath of plate tectonics, 335(6186), 131– 137.  Montes, C. (2005). Uplift of the Sierra Nevada de Santa Marta and subsidence in the CesarRancheria valley: Rigid-beam pivot model. 6th International Symposium on Andean Geodynamics (ISAG 2005, Barcelona).  Montes, C., Bayona, G., Cardona, A., Buchs, D. M., Silva, C. A., Morón, S., … Valencia, V. (2012). Arc-continent collision and orocline formation: Closing of the Central American seaway. Journal of Geophysical Research: Solid Earth, 117(B4), B04105.  Montes, C., Cardona, A., Jaramillo, C., Pardo, A., Silva, J. C., Valencia, V., … Niño, H. (2015). Middle Miocene closure of the Central American Seaway. Science, 348(6231), 226–229.  Montes, C. (2010a). Palinspastic reconstruction of northwestern south America during Cenozoic times. In prep.  Montes, C., Guzman, G., Bayona, G., Cardona, A., Valencia, V., & Jaramillo, C. (2010b). Clockwise rotation of the Santa Marta massif and simultaneous Paleogene to Neogene deformation of the Plato-San Jorge and Cesar-Ranchería basins. Journal of South American Earth Sciences, 29(4), 832–848.  Montes, C., Hatcher Jr., R. D., & Restrepo-Pace, P. A. (2005). Tectonic reconstruction of the northern Andean blocks: Oblique convergence and rotations derived from the kinematics of the Piedras–Girardot area, Colombia. Tectonophysics, 399(1–4), 221–250.  Montes, C., Restrepo-Pace, P. A., & Jr, R. D. H. (2003). Three-dimensional Structure and Kinematics of the Piedras-Girardot Fold Belt: Surface Expression of Transpressional Deformation in the Northern Andes, 849–873.  Mora, A., Horton, B. K., Mesa, A., Rubiano, J., Ketcham, R. A., Parra, M., … Stockli, D. F. (2010). Migration of Cenozoic deformation in the Eastern Cordillera of Colombia interpreted from fission track results and structural relationships: Implications for petroleum systems. AAPG Bulletin, 94(10), 1543–1580.  Mora, A., Parra, M., Strecker, M. R., Kammer, A., Dimaté, C., & Rodríguez, F. (2006). Cenozoic contractional reactivation of Mesozoic extensional structures in the Eastern Cordillera of Colombia. Tectonics, 25(2). Retrieved from  Müller, R. D., Sdrolias, M., Gaina, C., & Roest, W. R. (2008). Age, spreading rates, and spreading asymmetry of the world’s ocean crust: DIGITAL MODELS OF THE WORLD’S OCEAN CRUST. Geochemistry, Geophysics, Geosystems, 9(4), n/a–n/a.  Neill, I., Kerr, A. C., Hastie, A. R., Stanek, K.-P., & Millar, I. L. (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, 168(2), 333–348. O’Neill, C.,

51

Müller, D., & Steinberger, B. (2005). On the uncertainties in hot spot reconstructions and the significance of moving hot spot reference frames. Geochemistry, Geophysics, Geosystems, 6(4), Q04003.  Parra, M., Mora, A., Jaramillo, C., Strecker, M. R., Sobel, E. R., Quiroz, L., … Torres, V. (2009). Orogenic wedge advance in the northern Andes: Evidence from the Oligocene-Miocene sedimentary record of the Medina Basin, Eastern Cordillera, Colombia. Geological Society of America Bulletin, 121(5-6), 780–800.  Parra, M., Mora, A., Sobel, E. R., Strecker, M. R., & González, R. (2009). Episodic orogenic front migration in the northern Andes: Constraints from low-temperature thermochronology in the Eastern Cordillera, Colombia. Tectonics, 28(4), TC4004.  Parra, M., Mora, A., Jaramillo, C., Torres, V., Zeilinger, G., & Strecker, M. R. (2010). Tectonic controls on Cenozoic foreland basin development in the north-eastern Andes, Colombia. Basin Research, 22(6), 874–903.  Pindell, J. (1993). Regional synopsis of Gulf of Mexico and Caribbean evolution (pp. 251– 274). Presented at the GCSSEOM Foundation 13th Annual Research Conference.  Pindell, J., & 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., & Kennan, L. (2001). Kinematic evolution of the Gulf of Mexico and Caribbean. Perkins Research Conference, Petroleum Systems of Deep-Water Basins, Houston, Texas, 2– 5.  Pindell, J., Kennan, L., Draper, G., Maresch, W. V., & Stanek, K. P. (2006). Foundations of Gulf of Mexico and Caribbean evolution: eight controversies resolved. Geologica Acta: An International Earth Science Journal, 4(1), 303–341.  Pindell, J., Kennan, L., Maresch, W. V., Stanek, K.-P., Draper, G., & Higgs, R. (2005). Plate- kinematics and crustal dynamics of circum-Caribbean arc-continent interactions: Tectonic controls on basin development in Proto-Caribbean margins. Geological Society of America Special Papers, 394, 7–52.  Pindell, J. L., Cande, S. C., Pitman, W. C., Rowley, D. B., Dewey, J. F., Labrecque, J., & Haxby, W. (1988). 8th Geodynamics Symposium, “Mesozoic and Cenozoic Plate Reconstructions”A plate-kinematic framework for models of Caribbean evolution. Tectonophysics, 155(1), 121– 138.  Pindell, J. L., & Kennan, L. (2009). Tectonic evolution of the Gulf of Mexico, Caribbean and northern South America in the mantle reference frame: an update. Geological Society, London, Special Publications, 328(1), 1–55.  Priem H.N.A, Beets, D. ., & Vendurmen, E. A. . (1986). Precambrian rocks in an Early Tertiary conglomerate on Bonaire, Netherlands Antilles (southern Caribbean borderland): evidence for a 300km eastward displacement relative to the South American mainland? 1986, 35–40.  Prieto, G. A., Beroza, G. C., Barrett, S. A., López, G. A., & Florez, M. (2012). Earthquake nests as natural laboratories for the study of intermediate-depth earthquake mechanics. Tectonophysics, 570–571, 42–56.  Restrepo, J. ., & Toussaint, J. . (1988). Terranes and continental accretion in the Colombian Andes. Episodes.  Rincon, A. F., Bloch, J. I., Macfadden, B. J., & Jaramillo, C. A. (2013). First Central American record of Anthracotheriidae (Mammalia, Bothriodontinae) from the early Miocene of Panama. Journal of Vertebrate Paleontology, 33(2), 421–433.

52

 Rincon, A. F., Bloch, J. I., Macfadden, B. J., & Jaramillo, C. A. (2015). New early Miocene protoceratids (Mammalia, Artiodactyla) from Panama. Journal of Vertebrate Paleontology, 35(5), e970688.  Rogers, R. D., Mann, P., & Emmet, P. A. (2007). Tectonic terranes of the Chortis block based on integration of regional aeromagnetic and geologic data. Geological Society of America Special Papers, 428, 65–88.  Rosencrantz, E., & Mann, P. (1991). SeaMARC II mapping of transform faults in the Cayman Trough, Caribbean Sea. Geology, 19(7), 690–693.  Stanek, K. P., Maresch, W. V., & Pindell, J. L. (2009). The geotectonic story of the northwestern branch of the Caribbean Arc: implications from structural and geochronological data of Cuba. Geological Society, London, Special Publications, 328(1), 361–398.  Stearns, C., Mauk, F. J., & Van der Voo, R. (1982). Late Cretaceous-Early Tertiary paleomagnetism of Aruba and Bonaire (Netherlands Leeward Antilles). Journal of Geophysical Research: Solid Earth, 87(B2), 1127–1141.  Torsvik, T. H., Müller, R. D., Van der Voo, R., Steinberger, B., & Gaina, C. (2008). Global plate motion frames: Toward a unified model. Reviews of Geophysics, 46(3), RG3004.  Vallejo, C., Spikings, R. A., Luzieux, L., Winkler, W., Chew, D., & Page, L. (2006). The early interaction between the Caribbean Plateau and the NW South American Plate. Terra Nova, 18(4), 264–269.  Vallejo, C., Winkler, W., Spikings, R. A., Luzieux, L., Heller, F., & Bussy, F. (2009). Mode and timing of terrane accretion in the forearc of the Andes in Ecuador. Geological Society of America Memoirs, 204, 197–216.  van Benthem, S., Govers, R., Spakman, W., & Wortel, R. (2013). Tectonic evolution and mantle structure of the Caribbean. Journal of Geophysical Research: Solid Earth, 118(6), 3019–3036.  van der Lelij, R., Spikings, R. A., Kerr, A. C., Kounov, A., Cosca, M., Chew, D., & Villagomez, D. (2010). Thermochronology and tectonics of the Leeward Antilles: Evolution of the southern Caribbean Plate boundary zone. Tectonics, 29(6), TC6003.  van Hinsbergen, D. J. J., Iturralde-Vinent, M. A., van Geffen, P. W. G., García-Casco, A., & 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.  Vargas, C. A., & Mann, P. (2013). Tearing and Breaking Off of Subducted Slabs as the Result of Collision of the Panama Arc‐Indenter with Northwestern South America. Bulletin of the Seismological Society of America, 103(3), 2025–2046.  Villagómez Diaz, D. (2010). Thermochronology, geochronology and geochemistry of the Western and Central cordilleras and Sierra Nevada de Santa Marta, Colombia: the tectonic evolution of NW South America. University of Geneva.  Villagómez, D., Spikings, R., Magna, T., Kammer, A., Winkler, W., & Beltrán, A. (2011). Geochronology, geochemistry and tectonic evolution of the Western and Central cordilleras of Colombia. Lithos, 125(3–4), 875–896.  Weber, M. B. I., Cardona, A., Paniagua, F., Cordani, U., Sepúlveda, L., & Wilson, R. (2009). The Cabo de la Vela Mafic-Ultramafic Complex, Northeastern Colombian Caribbean region: a record of multistage evolution of a Late Cretaceous intra-oceanic arc. Geological Society, London, Special Publications, 328(1), 549–568.

53

 Weber, M., Cardona, A., Valencia, V., García-Casco, A., Tobón, M., & Zapata, S. (2010a). U/Pb detrital zircon provenance from late cretaceous metamorphic units of the Guajira Peninsula, Colombia: Tectonic implications on the collision between the Caribbean arc and the South American margin. Journal of South American Earth Sciences, 29(4), 805–816.  Weber, M., Cardona, A., Valencia, V., García-Casco, A., Tobón, M., & Zapata, S. (2010b). U/Pb detrital zircon provenance from late cretaceous metamorphic units of the Guajira Peninsula, Colombia: Tectonic implications on the collision between the Caribbean arc and the South American margin. Journal of South American Earth Sciences, 29(4), 805–816.  Wright, J. E., & Wyld, S. J. (2011). Late Cretaceous subduction initiation on the eastern margin of the Caribbean-Colombian Oceanic Plateau: One Great Arc of the Caribbean (?). Geosphere, 7(2), 468–493.  Zapata, S., Cardona, A., Montes, C., Valencia, V., Vervoort, J., & Reiners, P. (2014a). Provenance of the Eocene Soebi Blanco formation, Bonaire, Leeward Antilles: Correlations with post-Eocene tectonic evolution of northern South America. Journal of South American Earth Sciences, 52, 179–193.

54

Appendix 1. Supplementary Data: 1) Additional reconstruction Snapshots at 90 Ma, 79 Ma, 74 Ma, 28 Ma, 6 Ma.

55

Figure 1. (a) Tectonic blocks by major tectonic domains at 90 Ma. EC: Eastern Cordillera, CC: Central Cordillera, WC: Western Cordillera, MMB: Middle Magdalena basin, LMB: Lower Magdalena basin, UMB: Upper Magdalena basin, QG: Quebradagrande, CA: Central America, Accreted Terrains: Guajira allochthon and LMB allochthon from North to South. (b) Dividing faults used in the model. GF: Guaicaramó fault, PF: Palestina fault, RF: Romeral fault system, YP: Yopal fault, BSF: Bucaramanga-Santa Marta fault, OF: Oca fault, CF: Cusiana fault, UF: Uribe fault, CAF: Caguan fault, BAF: Borde Amazónico fault, AF: Algeciras fault, LSF: La Salina fault, IB: Ibagué fault, PAF: Pajarito fault, HF: Honda fault, DPF: De Prado fault, ALF: Altamizal fault, SPF: Santiago Perez fault, MF: Macama fault, SJF: San Jerónimo fault, OTF: Otu fault, CPF: Cali-Patia fault, UF: Uramita fault, EF: Encarnacion fault, BF: Bocono fault system, SSF: San Sebastian fault, EPF: El Pilar fault, URF: Urica fault, LAF: La Victoria fault, AMF: Amurrapa fault, DSF: De Sardinata fault, RIF: Riecito fault, ABF: Arenas Blancas fault, BNF: Burro Negro fault, RGF: Rio Gatun fault, PTF: Perija-El Tigre faults, LPF: La Plata fault, SF: Simarúa fault.

56

Figure 2. (a) Tectonic blocks by major tectonic domains at 79 Ma. EC: Eastern Cordillera, CC: Central Cordillera, WC: Western Cordillera, MMB: Middle Magdalena basin, LMB: Lower Magdalena basin, UMB: Upper Magdalena basin, QG: Quebradagrande, CA: Central America, Accreted Terrains: Guajira allochthon and LMB allochthon from North to South. (b) Dividing faults used in the model. GF: Guaicaramó fault, PF: Palestina fault, RF: Romeral fault system, YP: Yopal fault, BSF: Bucaramanga-Santa Marta fault, OF: Oca fault, CF: Cusiana fault, UF: Uribe fault, CAF: Caguan fault, BAF: Borde Amazónico fault, AF: Algeciras fault, LSF: La Salina fault, IB: Ibagué fault, PAF: Pajarito fault, HF: Honda fault, DPF: De Prado fault, ALF: Altamizal fault, SPF: Santiago Perez fault, MF: Macama fault, SJF: San Jerónimo fault, OTF: Otu fault, CPF: Cali-Patia fault, UF: Uramita fault, EF: Encarnacion fault, BF: Bocono fault system, SSF: San Sebastian fault, EPF: El Pilar fault, URF: Urica fault, LAF: La Victoria fault, AMF: Amurrapa fault, DSF: De Sardinata fault, RIF: Riecito fault, ABF: Arenas Blancas fault, BNF: Burro Negro fault, RGF: Rio Gatun fault, PTF: Perija-El Tigre faults, LPF: La Plata fault, SF: Simarúa fault.

57

Figure 3. (a) Tectonic blocks by major tectonic domains at 74 Ma. EC: Eastern Cordillera, CC: Central Cordillera, WC: Western Cordillera, MMB: Middle Magdalena basin, LMB: Lower Magdalena basin, UMB: Upper Magdalena basin, QG: Quebradagrande, CA: Central America, Accreted Terrains: Guajira allochthon and LMB allochthon from North to South. (b) Dividing faults used in the model. GF: Guaicaramó fault, PF: Palestina fault, RF: Romeral fault system, YP: Yopal fault, BSF: Bucaramanga-Santa Marta fault, OF: Oca fault, CF: Cusiana fault, UF: Uribe fault, CAF: Caguan fault, BAF: Borde Amazónico fault, AF: Algeciras fault, LSF: La Salina fault, IB: Ibagué fault, PAF: Pajarito fault, HF: Honda fault, DPF: De Prado fault, ALF: Altamizal fault, SPF: Santiago Perez fault, MF: Macama fault, SJF: San Jerónimo fault, OTF: Otu fault, CPF: Cali-Patia fault, UF: Uramita fault, EF: Encarnacion fault, BF: Bocono fault system, SSF: San Sebastian fault, EPF: El Pilar fault, URF: Urica fault, LAF: La Victoria fault, AMF: Amurrapa fault, DSF: De Sardinata fault, RIF: Riecito fault, ABF: Arenas Blancas fault, BNF: Burro Negro fault, RGF: Rio Gatun fault, PTF: Perija-El Tigre faults, LPF: La Plata fault, SF: Simarúa fault.

58

Figure 4. (a) Tectonic blocks by major tectonic domains at 28 Ma. EC: Eastern Cordillera, CC: Central Cordillera, WC: Western Cordillera, MMB: Middle Magdalena basin, LMB: Lower Magdalena basin, UMB: Upper Magdalena basin, QG: Quebradagrande, CA: Central America, Accreted Terrains: Guajira allochthon and LMB allochthon from North to South. (b) Dividing faults used in the model. GF: Guaicaramó fault, PF: Palestina fault, RF: Romeral fault system, YP: Yopal fault, BSF: Bucaramanga-Santa Marta fault, OF: Oca fault, CF: Cusiana fault, UF: Uribe fault, CAF: Caguan fault, BAF: Borde Amazónico fault, AF: Algeciras fault, LSF: La Salina fault, IB: Ibagué fault, PAF: Pajarito fault, HF: Honda fault, DPF: De Prado fault, ALF: Altamizal fault, SPF: Santiago Perez fault, MF: Macama fault, SJF: San Jerónimo fault, OTF: Otu fault, CPF: Cali-Patia fault, UF: Uramita fault, EF: Encarnacion fault, BF: Bocono fault system, SSF: San Sebastian fault, EPF: El Pilar fault, URF: Urica fault, LAF: La Victoria fault, AMF: Amurrapa fault, DSF: De Sardinata fault, RIF: Riecito fault, ABF: Arenas Blancas fault, BNF: Burro Negro fault, RGF: Rio Gatun fault, PTF: Perija-El Tigre faults, LPF: La Plata fault, SF: Simarúa fault.

59

Figure 5. (a) Tectonic blocks by major tectonic domains at 6 Ma. EC: Eastern Cordillera, CC: Central Cordillera, WC: Western Cordillera, MMB: Middle Magdalena basin, LMB: Lower Magdalena basin, UMB: Upper Magdalena basin, QG: Quebradagrande, CA: Central America, Accreted Terrains: Guajira allochthon and LMB allochthon from North to South. (b) Dividing faults used in the model. GF: Guaicaramó fault, PF: Palestina fault, RF: Romeral fault system, YP: Yopal fault, BSF: Bucaramanga-Santa Marta fault, OF: Oca fault, CF: Cusiana fault, UF: Uribe fault, CAF: Caguan fault, BAF: Borde Amazónico fault, AF: Algeciras fault, LSF: La Salina fault, IB: Ibagué fault, PAF: Pajarito fault, HF: Honda fault, DPF: De Prado fault, ALF: Altamizal fault, SPF: Santiago Perez fault, MF: Macama fault, SJF: San Jerónimo fault, OTF: Otu fault, CPF: Cali-Patia fault, UF: Uramita fault, EF: Encarnacion fault, BF: Bocono fault system, SSF: San Sebastian fault, EPF: El Pilar fault, URF: Urica fault, LAF: La Victoria fault, AMF: Amurrapa fault, DSF: De Sardinata fault, RIF: Riecito fault, ABF: Arenas Blancas fault, BNF: Burro Negro fault, RGF: Rio Gatun fault, PTF: Perija-El Tigre faults, LPF: La Plata fault, SF: Simarúa fault.

60

61