1 The biostratigraphic record of Early Cretaceous to Paleogene relative sea-level

2 change in 3 David P. Gold a, * James P. G. Fentona, Manuel Casas-Gallegoa, Vibor Novaka, Irene Pérez-Rodrígueza, 4 Claudia Ceteana, Richard Pricea, Nicole Nembhardb, Herona Thompsonb 5 6 7 a,* CGG Robertson, Llandudno, North Wales LL30 1SA, United Kingdom 8 9 b Petroleum Corporation of Jamaica Ltd, 5th Floor PCJ Building, 36 Trafalgar Road, Kingston 10, 10 Jamaica 11 12 13 14

* Corresponding author at: CGG Robertson, part of CGG GeoConsulting, Llandudno, North Wales LL30 1SA, United Kingdom. Tel: +44 1492 581111 E-Mail address: [email protected] (David Gold)

Page 1 of 54 15 ABSTRACT

16 The island of Jamaica forms the northern extent of the Nicaraguan Rise, an elongate linear tectonic 17 feature stretching as far as Honduras and to the south. Uplift and subaerial exposure of 18 Jamaica during the Neogene has made the island rare within the Caribbean region, as it is the only area 19 where rocks of the Nicaraguan Rise are exposed on land. Biostratigraphic dating and 20 palaeoenvironmental interpretations using larger benthic foraminifera, supplemented by planktonic 21 foraminifera, nannopalaeontology and palynology of outcrop, well and corehole samples has enabled 22 the creation of a regional relative sea-level curve through identification of several depositional 23 sequences. This study recognises ten unconformity-bounded transgressive-regressive sequences which 24 record a complete cycle of relative sea level rise and fall. Sequences are recognised in the Early to 25 ‘Middle’ Cretaceous (EKTR1), Coniacian-Santonian (STR1), Campanian (CTR1), Maastrichtian 26 (MTR1-2), Paleocene-Early Eocene (PETR1), Eocene (YTR1-3) and Late Eocene-Oligocene (WTR1). 27 These transgressive-regressive cycles represent second to fourth order sequences, although most tie 28 with globally recognised third order sequences. Comparisons of the Jamaican relative sea-level curve 29 with other published global mean sea-level curves show that local tectonics exerts a strong control on 30 the deposition of sedimentary sequences in Jamaica. Large unconformities (duration >1 Ma) are 31 related to significant regional tectonic events, with minor overprint of a global eustatic signal, while 32 smaller unconformities (duration <1 Ma) are produced by global eustatic trends. The relatively low 33 rates of relative sea-level rise calculated from the regional relative sea-level curve indicate that 34 carbonate production rates were able to keep pace with the rate of relative sea-level rise accounting for 35 the thick successions of Maastrichtian carbonates and those of the Yellow and White Limestone 36 Groups. Carbonate platform drowning within the White Limestone Group during the Oligocene to 37 Miocene is attributed to environmental deterioration given the low rates of relative sea-level rise. 38 39 Keywords: Carbonates, biostratigraphy, foraminifera, relative sea-level, tectonics

40

Page 2 of 54 41 1. Introduction

42 Jamaica is the third largest island in the Caribbean, behind Cuba and Hispaniola. It is situated in 43 the western portion of the and to the west of the Greater Antilles arc. The island of 44 Jamaica forms the northern extent of the Nicaraguan Rise (Fig. 1), a major NE-SW trending 45 bathymetric and structural feature extending as far as Honduras and Nicaragua to the southwest (Hine 46 et al., 1992; Mutti et al., 2005). The Nicaraguan Rise is the north-eastern submarine continuation of 47 the Chortis Block (James, 2007; Carvajal-Arenas et al., 2015; Sanchez et al., 2016; Bunge et al., 2017) 48 and comprises upper and lower parts, divided by the Pedro Bank Fault Zone (Mutti et al., 2005; James, 49 2007; Ott et al., 2013). Sanchez et al. (2016) infer that the upper part of the Nicaraguan Rise, which 50 includes western and central Jamaica, is underlain by crystalline rocks of island-arc affinity, whereas 51 the lower part has an oceanic plateau (Caribbean Large Igneous Province) origin. 52 The Nicaraguan Rise is bound to the northwest by the Cayman Trough and to the southeast by 53 the remarkably straight Hess escarpment (Fig. 1), which separates it from the Colombian Basin (Hine 54 et al., 1992; Mutti et al., 2005; James, 2007; James-Williamson et al., 2014; Carvajal-Arenas et al., 55 2015; Bunge et al., 2017). The Cayman Ridge runs parallel to the northern margin of the Nicaraguan 56 Rise to the north of the Cayman Trough (James, 2007). The upper part of the Nicaraguan Rise tapers 57 towards the northeast and is segmented into banks such as the Gorda, Rosalind and Pedro Banks, and 58 the island of Jamaica (James, 2007). The maximum crustal thickness of the Nicaraguan Rise is 59 calculated between 22 and 28km based on seismic refraction data (Ewing et al. 1960; Arden 1969;

60 Holcombe et al. 1990; Sanchez et al., 2016). 61 The present day configuration of Nicaraguan Rise comprises a shallow water carbonate 62 ‘megabank’ (Droxler, 1991; 1993; Sigurdsson et al., 1997; Mutti et al., 2005), it is up to 540 km in 63 width and 1,350 km in length, dropping steeply in to the Cayman Trough on its north-western margin, 64 with a gentler incline into the Colombian Basin on its south-eastern side (Arden Jr., 1975; Hine et al., 65 1992). The island of Jamaica is rare within the Caribbean region, as it is the only area where rocks of 66 the Nicaraguan Rise are exposed on land. 67 Jamaica and its offshore sedimentary basins are sparsely explored, with only 2 wells drilled 68 offshore, which are not in a present day basin setting, and 9 wells onshore along with several shallow 69 coreholes (Fig. 2). No wells have been drilled since the mid-1980s. There have been oil or gas shows 70 in 10 of the 11 wells drilled in the onshore and offshore to date. The two offshore wells are Arawak-1 71 and Pedro Bank-1; the following nine wells are onshore: Cockpit-1, Content A/Content-1, Hertford-1, 72 Negril Spots-1 (formerly known as Jamaica-1), Portland Ridge-1, Retrieve-1, Santa Cruz-1, West 73 Negril-1 and Windsor-1. The distribution of all 11 wells across on- and offshore Jamaica is depicted in 74 Figure 2.

Page 3 of 54 75 The objective of this study is to describe new data from Jamaica which comprises new analyses 76 of over 800 outcrop and well/corehole samples. In addition, these new biostratigraphic and 77 sedimentological analyses revise and update the sequence-stratigraphic framework of Jamaica and tie 78 depositional cycles to third-order sequences of Hardenbol et al. (1998) recalibrated to the 2012 79 Geologic time scale (Gradstein et al., 2012). 80 81 2. Geological history

82 2.1. Early to Late Cretaceous

83 Formation of the Caribbean oceanic plateau or large igneous province (CLIP), which forms the 84 interior of the Caribbean Plate, occurred during the Early Jurassic (Burke et al., 1978; Burke, 1988). 85 By the Early Cretaceous, the Yucatán and Chortis Blocks had reached their present position (Rogers et 86 al., 2007; Carvajal-Arenas et al., 2015) and the crustal fragments that later assembled to form Jamaica 87 and the Nicaraguan Rise were in a period of relative tectonic quiescence. During this time shallow 88 marine limestones of the Benbow Inlier, and similar rocks known from Puerto Rico and Trinidad, were 89 deposited in vast shallow seas close to a volcanic arc (Kauffman and Sohl, 1974; Draper, 1987; 1998; 90 James-Williamson et al., 2014). These shallow seas were connected by the proto- Caribbean seaway 91 which had opened as North and South America began to separate during the Late Jurassic to Early 92 Cretaceous continuation of the break-up of Pangaea (Burke 1988; Pindell et al. 1988; Pindell 1993; 93 Mann et al. 2006; Rogers et al., 2007; Carvajal-Arenas et al., 2015; Sanchez et al., 2016). Strata of the 94 Benbow Inlier are the oldest exposed in Jamaica, where there are no sediments older than Valanginian 95 age (Brown and Mitchell, 2010). Early to ‘Middle’ Cretaceous volcanics in the Central Inlier and the 96 succession of the Benbow Inlier are interpreted to represent a volcanic island arc and/or ophiolite suite 97 formed either as part of the CLIP, or a back/fore-arc spreading centre (Wadge et al., 1982; Jackson, 98 1987; Kerr et al., 2004; Mitchell, 2006). The Turonian-Coniacian unconformity is interpreted to have 99 formed during the assembly of CLIP fragments and subduction of normal ocean floor at convergent 100 margins on the northern and southern peripheries of unusually buoyant CLIP ocean floor (Burke et al., 101 1978; Burke, 1988; Mitchell, 2006; Hastie et al., 2010a; 2010b). 102 103 2.2. Late Cretaceous

104 Following the formation of the Turonian-Coniacian unconformity, deep water sedimentation 105 was prevalent over Jamaica during the Coniacian and Santonian. This deep water setting was 106 responsible for the deposition of the Windsor, Clamstead, Middlesex and Dias Formations. The 107 transition to deep-water siliciclastics of these units suggests rapid subsidence and increase in the rate 108 of relative sea-level rise associated with continued intra-arc rifting (Mitchell, 2003). The primary

Page 4 of 54 109 control on the Caribbean region during the Late Cretaceous was the entry of the Great Arc of the 110 Caribbean into the proto-Caribbean seaway (Burke 1988; Pindell et al. 1988; Pindell 1993; Mann et al. 111 2006; Carvajal-Arenas et al., 2015; Sanchez et al., 2016). A jump in the position of the magmatic arc 112 during the latest Santonian, possibly related to the jarring of buoyant CLIP arriving at the trench of the 113 volcanic arc (Burke et al., 1978; Hastie et al., 2010a; 2010b), was responsible for the formation of the 114 early Campanian unconformity at the base of the Crofts Synthem (Mitchell, 2003). 115 Mitchell (2003; 2006) attributed the late Campanian to early Maastrichtian unconformity to 116 uplift in the northern Nicaraguan Rise related to collision in the south between the Chortis Block and 117 arc, and oceanic plateau rocks of the CLIP (Rogers et al., 2007; Carvajal-Arenas et al., 2015; Sanchez 118 et al., 2016). This caused NW-SE shortening of the Colon fold and thrust belt in Honduras and 119 Nicaragua which extended as far as the northern Nicaraguan Rise (Rogers et al., 2007; Carvajal- 120 Arenas et al., 2015; Sanchez et al., 2016). This event is associated with the formation of presently 121 southward-verging imbricate thrust sheets onshore Jamaica as island arc material of the upper 122 Nicaraguan Rise, including western and central Jamaica, were thrust over CLIP rocks of the lower 123 Nicaraguan Rise (Burke, 1988; Pindell and Barrett, 1990; Pindell, 1994; Pindell and Kennan, 2001; 124 2009; Mitchell, 2003; 2006; Kerr et al., 2004; Pindell et al., 2005; Rogers et al., 2007; Carvajal- 125 Arenas et al., 2015; Sanchez et al., 2016). These remnant thrust sheets are now expressed as a series of 126 E-W trending faults, which are the oldest structural features found on Jamaica. 127 The late Maastrichtian ‘Kellits Synthem’ of Mitchell (2003) is reported to represent a single 128 major transgressive-regressive cycle (Mitchell and Blissett, 2001; Mitchell, 2003; 2006). This study 129 recognises two T-R cycles within the Maastrichtian. In the east of Jamaica, a suture is recorded in the 130 John Crow Mountain area between characteristic island arc material that comprises the basement of 131 western and central Jamaica with crust similar to that of Haiti east of the suture (Lewis et al., 2011). 132 The suturing of western/central and eastern Jamaica occurred in response to the collision of the 133 Caribbean Plate with the Bahamas during the Maastrichtian to Paleocene (Arden Jr, 1969; 1975; Mann 134 et al., 1995; Calais, 1999; McCann, 1999; James-Williamson et al., 2014). This resulted in the 135 subduction of CLIP crust beneath Jamaica and culminated in the formation of the Cretaceous-Tertiary 136 unconformity. This suture is expressed by the Plantain Garden Fault Zone which represents the 137 southern margin of Gonâve Microplate (Mann et al., 1995; Calais, 1999; McCann, 1999; James- 138 Williamson et al., 2014). 139 140 2.3. Paleogene

141 Continued subduction and melting of the CLIP beneath Jamaica during the early Paleogene was 142 responsible for the formation of the proto- and extrusion of exotic volcanics, such as 143 adakites, within the Early Eocene age Halberstadt Volcanics (Hastie et al., 2010a; 2010b; 2011). This

Page 5 of 54 144 may have been responsible for the formation of an unconformity recorded in eastern Jamaica at this 145 time. 146 The unconformity between the Kellits Synthem and Middle Eocene strata of the Yellow 147 Limestone Group over much of western Jamaica (Clarendon and Hanover Blocks) is interpreted to 148 have formed during Late Paleocene to Early Eocene NE-SW directed extension (Mitchell, 2003). This 149 extension was related to sinistral strike-slip movement between the North American Plate and the 150 northern margin of the Caribbean Plate (Pindell, 1994; Mitchell, 2003). Extension occurred over much 151 of the Nicaraguan Rise which resulted in the formation of NW-SE trending horst and graben structures 152 such as the onshore Montpelier-Newmarket and Wagwater Troughs, and the offshore Walton Basin 153 graben and the Pedro Bank horst (Draper, 1987; Mann and Burke, 1990; James-Williamson et al., 154 2014; Carvajal-Arenas et al., 2015). 155 The grabens were filled with sediments which were shed from adjacent highs, most notably the 156 Paleocene strata of the Wagwater Trough in the east of the island (Arden Jr., 1969). During the Middle 157 Eocene, collision of the Great Arc of the Caribbean in Cuba with the Bahamas caused movement of 158 the Caribbean Plate to change from north-eastward to an eastward direction (Mann et al., 1995; 159 Carvajal-Arenas et al., 2015). Reactivation of NW-SE striking extensional faults to reverse faults 160 during the Late Eocene resulted in the denudation of the tops of sedimentary sections and may have 161 been responsible for the unconformity between the White and Yellow Limestone units. 162 163 2.4. Neogene

164 During the Neogene, sinistral strike-slip motion along the boundary between northern Jamaica 165 and Cayman Trough was transferred to the southern boundary of the Gonâve Plate by fault 166 displacement through the Wagwater Trough (Mann et al., 1985; 1995; DeMets and Wiggins- 167 Grandison, 2007; Abbott Jr. et al., 2013; James-Williamson et al., 2014). The collision and locking of 168 the South Hispaniola and Gonâve Plates, and the Nicaraguan Rise, is believed to have caused this 169 switch (DeMets and Wiggins-Grandison, 2007). Consequently, sinistral displacement was transferred 170 through a restraining bend in the area of the Blue Mountains which resulted in major transpression and 171 significant uplift (Mann and Burke, 1984; Mann et al., 2007; James-Williamson et al., 2014; Carvajal- 172 Arenas et al., 2015). 173 The youngest tectonic event in Jamaica began in the Late Miocene, approximately 8-9 Ma, and 174 is continuing to the present day (Wright, 1971; Steineck, 1974; 1981; Katz and Miller, 1993). 175 Convergence of the North American Plate along the NW-SE oriented northern margins of Cuba and 176 Hispaniola resulted in a stress regime which is expressed by the Neogene uplift of Jamaica. This final 177 stage of tectonic deformation caused Plio-Pleistocene age carbonate terraces to be uplifted at a rate of 178 1.5mkyr-1 (Steineck, 1974; Katz and Miller, 1993; Cochran et al., 2017) and regional NW directed

Page 6 of 54 179 extension forming a large-scale structural trend that is broadly NE-SW. Within this stress regime the 180 former Campanian E-W oriented thrust sheets are reactivated as strike-slip faults to further 181 accommodate sinistral displacement and folding of the Oligo-Miocene age Montpelier Formation 182 produced spectacular NE verging folds that can be observed close to the north coast of Jamaica. 183 184 3. Carbonate depositional settings

185 During the Cretaceous, many of the rudist-bearing limestones of Jamaica were deposited in vast 186 shallow seas. It is interpreted that during this time Jamaica comprised a series of island platforms that 187 were punctuated by isolated rudist build-ups. These rudist build-ups consisted of overgrown thickets 188 and coppices that never developed into more complex rudist banks or barrier reefs (Kauffman and 189 Sohl, 1974). These extensive carbonate platforms had well-developed margins that passed basinward 190 from platform margin slope to continental shelf and distal bathyal settings. Consequently, the 191 terminology of the Cretaceous sequences is described in carbonate platform and shelf terminology. 192 The term carbonate 'platform' is generally applied to any thick, relatively flat- topped 193 carbonate depositional system (Wright and Burchette, 1998). Carbonate ramps are carbonate platforms 194 which have a low-gradient submarine depositional slope which is dominated by carbonate 195 sedimentation (Burchette and Wright, 1992; Wright and Burchette, 1998). Read (1985) further 196 subdivided carbonate ramps to distally steepened and homoclinal, based on the presence or absence of 197 outer ramp change in slope angle, respectively. The Nicaraguan Rise is described to comprise a regular 198 gradient for approximately 200km to the southeast, before a steeper decline into the Basin 199 with a well-defined slope break (Arden Jr., 1975; Hine et al., 1992). The bathymetric expression of the 200 Nicaraguan Rise is therefore consistent with distally steepened carbonate ramps (Fig. 1). Although 201 Hine et al. (1992) suggest the Nicaraguan Rise is not a carbonate ramp, they concede it does not 202 comprise rimmed platforms, due to the absence of reefal, rock or sediment barriers. A carbonate 203 ‘megabank’ has also been described to make up most of the modern northern Nicaraguan Rise since 204 the Paleogene (Droxler, 1991; 1993; Sigurdsson et al., 1997; Mutti et al., 2005; James, 2007), 205 although siliciclastic sediments are also known from Jamaica from this time. An analogue for the 206 Nicaraguan Rise megabank is the nearby Great Bahama Bank which seismic data indicates formed a 207 modern megabank through a change in geometry from a distally steepened carbonate ramp to a flat 208 topped platform during the Pliocene (Eberli and Ginsburg, 1987; 1989; Wilber et al., 1990; Betzler et 209 al., 1999). It has also been noted that on satellite-derived bathymetry the Nicaraguan Rise repeats the 210 highly extended appearance of the Bahamas Bank (James, 2007). Based on these observations, 211 carbonate ramp terminology is used within this paper to describe the gross carbonate depositional 212 environments of the Nicaraguan Rise since its development the Paleogene. Although Paleogene 213 extension across the Nicaraguan Rise formed a series of fault blocks in Jamaica, where shallow marine

Page 7 of 54 214 carbonates and siliciclastics were deposited around localised block crests and marginal half-grabens, 215 general carbonate ramp terminology is still preferred to describe the broad depositional setting of the 216 northern Nicaraguan Rise. There are many examples of extended carbonate ramps that are dissected by 217 horst and graben structures, or tilted fault block ramps, including Cenozoic ramps from Sulawesi, 218 Indonesia (Wilson et al., 2000), Mesozoic ramps of the Iberian Peninsula (Gómez-Pérez et al., 1998; 219 Azeredo et al., 2002; Mercedes-Martín et al., 2013; 2014), Late Paleozoic ramps of the Barents Sea 220 (Di Lucia et al., 2017) and intra-ramp basins within the Arabian ramp through much of the 221 Phanerozoic (Murris et al., 1981). 222 223 4. Material and methods

224 This research forms part of a wider study evaluating the petroleum potential of Jamaica. Fieldwork in 225 Jamaica was conducted over 25 days between October and November 2016. During the fieldwork, 60 226 formations were described and sampled, with over 200 samples collected from over 200 localities (Fig. 227 2). Field samples were collected using a spot sampling strategy to cover a broad area that incorporates 228 most of the Jamaican stratigraphy relevant to petroleum history owing to few good continuous sections 229 not overgrown by dense vegetation. An additional 600 samples were collected from 10 Jamaican wells 230 and 10 shallow coreholes (Table 1) provided by the Petroleum Corporation of Jamaica (Fig. 2). In 231 addition, data from 11 DSDP and ODP wells were reviewed (Table 1). 232

Subsurface data source Type Arawak-1 Well Cockpit-1 Well Content A/Content-1 Well Hertford-1 Well Negril Spots-1 (formerly known as Jamaica-1) Well Portland Ridge-1 Well Retrieve-1 Well Santa Cruz-1 Well West Negril-1 Well Windsor-1 Well Blowfire Hill Corehole Ecclesdown Corehole Elderslie Corehole Jerusalem Mountain Corehole Marchmont Corehole Pindars River-3 Corehole Pindars River-10 Corehole Sunderland Blackshop-1 Corehole Sunderland Potosi-2 Corehole

Page 8 of 54 Windsor Corehole DSDP 151 Drilling programme site DSDP 152 Drilling programme site DSDP 153 Drilling programme site DSDP 154 Drilling programme site DSDP 154A Drilling programme site ODP 999A Drilling programme site ODP 999B Drilling programme site ODP 1000A Drilling programme site ODP 1000B Drilling programme site ODP 1001A Drilling programme site ODP 1001B Drilling programme site 233 Table 1. Names of wells, shallow coreholes and drilling programme sites used in this study 234 235 4.1. Relative sea-level curve

236 The interpretation of sea-level change in Jamaica was made using traditional sequence 237 stratigraphic concepts. Depositional sequences were identified as relatively conformable successions 238 of genetically related strata bounded by unconformities, or their correlative conformities (Mitchum et 239 al., 1977; Van Wagoner et al., 1988; Haq et al., 1988), that were deposited during a complete cycle of 240 sea-level rise and fall (Haq et al., 1988). Detailed description of sequence stratigraphy and its related 241 surfaces (e.g. sequence boundaries and maximum flooding surfaces) is beyond the scope of this work, 242 however, key references can be found within Vail et al., (1987), Van Wagoner et al. (1987), Haq et al. 243 (1988), Posamentier et al. (1993), Hardenbol et al. (1998) and Catuneanu (2006; 2011), amongst 244 others. 245 The first step in the creation of the relative sea-level (RSL) curve was to create a robust 246 chronostratigraphic framework (e.g. Haq et al., 1988) for Jamaica. Outcrop and well samples were 247 analysed for biostratigraphic dating using standard industrial biostratigraphic techniques incorporating 248 analyses of planktonic and larger benthic foraminifera, palynomorphs and nannofossils. Ages were 249 established, wherever possible, from the published ranges of regionally or globally recognised marker 250 fossils. Important global references for planktonic foraminifera include Sliter (1989) and Premoli Silva 251 and Verga (2004) for the Cretaceous, Olsson et al. (1999) for the Paleocene and Pearson et al. (2006) 252 for the Eocene. The ranges of larger benthic foraminifera were taken from BouDagher-Fadel (2008), 253 local references (Hanzawa, 1962; Jiang and Robinson, 1987; Robinson and Wright, 1993; Krijnen et 254 al., 1993; Robinson, 2004) and new interpretations of this study (Fig. 3). The planktonic foraminiferal 255 zones used in the Paleogene are those of Berggren et al. (1995) and Berggren and Pearson (2005), 256 recalibrated as sub-tropical foraminiferal Zones for the Cenozoic by Wade et al. (2011). Within the 257 Cretaceous, the zones are mainly a compilation by Gradstein et al. (2012) based on Robaszynski (in 258 Hardenbol et al., 1998). The nannofossil zonations used in the Tertiary section are the NP Zones of

Page 9 of 54 259 Martini (1971), with the global zonations of Sissingh (1977) and Perch-Nielsen (1979, 1985a, 1985b) 260 being applied in the Cretaceous. The time scale utilized was primarily taken from TSCreator (version 261 6.1.2, www.tscreator.org) using the 2012 Time scale (Gradstein et al., 2012). 262 The final step in the construction of the relative sea-level curve involved the integration of sea- 263 level events within the stratigraphic framework which was constrained by biostratigraphic analyses. 264 Where other sea-level curves use ‘realistic’ estimates of the magnitudes of sea-level rise and fall from 265 seismic and sequence stratigraphic data (e.g. Haq et al., 1988), and backstripping methods (Horowitz, 266 1976), the curve presented by this study uses palaeontological data that follows the ‘palaeobathymetric 267 resolution’ method of Hardenbol et al. (1981). This technique integrates stratigraphic, lithological, 268 environmental and palaeobathymetric data to determine the magnitude of sea-level change. 269 Paleobathymetric and environmental interpretations foraminiferal assemblages were based on Bandy 270 (1953), Tipsword et al. (1966), Steineck (1974), Eva (1976), Pflum and Frerichs (1976), Bé (1977), 271 Hallock and Glenn (1986), Murray (1991), Beavington-Penney and Racey (2004), Robinson (2004), 272 Hohenegger (2005), Murray (2006), BouDagher-Fadel (2008), Baker et al. (2009) and Mitchell 273 (2013). These studies use the modern and ancient depth distribution of mostly foraminiferal taxa as 274 potential water depth indicators. Hardenbol et al. (1991) suggest that the accuracy of the 275 ‘palaeobathymetric resolution’ method is acceptable for shallow-water deposits of up to 200 m, which 276 is suitable from the predominantly shallow-water limestones of Jamaica. The use of shallow benthic 277 and planktonic foraminifera as depth indicators have been used successfully to determine relative sea- 278 level histories and sequence stratigraphy in many other similar studies (e.g. Reiss and Hottinger, 1984; 279 Olsson and Nyong, 1994; Banner and Simmons, 1994; Arnaud-Vanneau, 1994; Gary et al., 1996; 280 Filipescu and Gîrbacea, 1997; Naish and Kamp, 1997; Cabioch et al., 1999; Simmons et al., 1999; 281 2000; Sharland, 2001; Leckie and Olson, 2003; Jones et al., 2004; amongst others). In addition to 282 foraminifera, palynological assemblages were used to determine depositional sequences (e.g. Morley, 283 1996; Simmons et al., 1999). 284 Seven depositional settings were defined for the purposes of this study and were assigned a 285 range of water depths below mean sea-level including: terrestrial (above mean sea-level), paralic (0m- 286 5m), proximal inner ramp/platform (5m-10m), distal inner ramp/platform (10m-15m), middle 287 ramp/shelf (15m-30m), outer ramp/shelf (30m-50m) and basinal (50m-100+m). During the Paleogene, 288 the bathymetric preferences of organisms were calibrated to the hydrodynamic boundaries of a 289 carbonate ramp depositional profile as defined by Burchette and Wright (1992), whereby fair-weather 290 and storm wave base occur at 10m-20m and 30-50m, respectively (Tucker and Wright, 1990). 291 Each sample was assigned a water depth within the range of their interpreted depositional 292 setting based on relative abundances of taxa or morphological characteristics. Through biostratigraphic 293 analyses the samples were also assigned an age in Ma, based on the biozones they were allocated. 294 Samples of the same age from different locations and basins across Jamaica were grouped together and

Page 10 of 54 295 an average water depth was calculated for the time interval. Average water depths were plotted against 296 Ma to create a relative sea-level curve. Standard error bars for each data point on the sea-level curve 297 incorporate the full range of environments interpreted for that period of time by calculating the 298 maximum, minimum and average bathymetry of all biostratigraphic control points. 299 In addition to palaeobathymetric estimates from foraminifera, and other palaeontological 300 evidence, the RSL curve was enhanced using the methodology of Heckel (1986) and Sahagian et al. 301 (1996). This involved identification of progradational and retrogradational sequences, calibrated by 302 palaeontological data, which allowed attribution of sea-level rise and fall (e.g. Sahagian et al., 1996; 303 Simmons et al., 2007). Several transgressive-regressive sequences were identified by lithological and 304 facies changes (e.g. normal and reverse grading) through new sedimentological analyses of well or 305 outcrop successions and compared to similar transgressive-regressive trends already documented in 306 Jamaica including the Kellits and Crofts synthems of Mitchell (2003). 307 Unconformities that bound depositional sequences may correlate to global sequence 308 boundaries and were recognised through the presence of hiatuses quantified using biostratigraphic 309 techniques. Sequence boundaries are classified as major, medium, and minor (Haq et al., 1988). 310 Sequence stratigraphic bounding surfaces of Hardenbol et al. (1998) were assigned to Jamaican 311 depositional sequences based on the most significant (major or medium) sequence boundary that 312 occurs within the biozone assigned to the sample through biostratigraphic analyses. Therefore 313 assignation of sequence stratigraphic surfaces to those of Hardenbol et al. (1998) was secondary to the 314 calibration of inflections of the sea-level curve to the biozones in which they occur. 315 In total 266 samples were assigned a depositional setting, palaeobathymetry and 316 biostratigraphic age and were used to create the sea-level curve (see supplementary data). This 317 regional RSL curve was then compared to global sea-level curves (Haq and Al-Qahtani, 2005; Kominz 318 et al., 2008; Snedden and Liu, 2010; Miller et al., 2005; 2011) to assess potential timing of tectonic 319 events. If the same sequence-stratigraphic surfaces were recognised in multiple locations or basins it 320 was interpreted that these depositional sequences relate to synchronous eustatic sea-level change (e.g. 321 Simmons et al., 2007) or regionally significant tectonic events. Calculations of rates of relative sea- 322 level change were also compared to calculations of subsidence rates (uncorrected for compaction) 323 from age-depth curves of well successions using StrataBugs v.2.1 to further constrain the RSL curve. 324 The RSL curve presented in this study is a composite curve using average water depths of 325 samples collected from different basins and locations across the island of Jamaica designed to 326 highlight broad trends in changing RSL. In the same way as the depositional sequences of Sharland et 327 al. (2001; 2004), correlated to the RSL curves of Haq et al. (1988), are composites of various basins 328 of the Arabian Plate, the depositional sequences and RSL curve presented here display average water 329 depths across the northern Nicaraguan Rise and is not specific to individual Jamaican basins. This 330 curve constitutes a simplified first attempt at reconstructing RSL change in Jamaica. This curve may

Page 11 of 54 331 be refined in the future by higher resolution field sampling, well analyses, localised basin-specific 332 curves and numerical forward modelling. This will enable the identification of further higher-order 333 depositional cycles and better determine, potentially multiple, controls on deposition of these 334 sequences. 335 336 5. Depositional sequences

337 Environmental preferences of organisms identified during biostratigraphic analyses permit the 338 construction of a RSL curve for the island of Jamaica (Figs. 4-8). The highest resolution of data points 339 occur during the Maastrichtian and Middle to Late Eocene. The RSL curve reveals ten unconformity- 340 bounded transgressive-regressive (T-R) depositional cycles which record a complete cycle of relative 341 sea level rise and fall in the stratigraphy of Jamaica. Commonly, only the transgressive deposits are 342 preserved, as strata deposited in the regressive phase of the previous sequence are eroded during 343 transgressive reworking and production of transgressive ravinement surfaces. Maximum transgressive 344 and regressive inflections in the RSL curve are correlated with third-order sequences. However, the 345 deposition of each sequence may be controlled by interplay between regional tectonism and global 346 eustatic trends. 347 Single T-R cycles are identified within the Early Cretaceous (EKTR1), Coniacian-Santonian 348 (STR1) and Campanian (CTR1), two T-R cycles are identified in the Maastrichtian (MTR1, MTR2) 349 and a further single T-R cycle is recognised within the Paleocene-Early Eocene (PETR1). Previous 350 studies have suggested the Middle Eocene age Yellow Limestone Group can be subdivided into two 351 depositional cycles; the lower or Stettin cycle and the upper or Chapelton-Troy cycle (Maharaj and 352 Mitchell, 2000; Matchette-Downes and Mitchell, 2005). However, this study recognises three T-R 353 cycles within the Yellow Limestone Group (YTR1, YTR2, YTR3), and a single transgressive 354 sequence within the Late Eocene to Miocene age White Limestone Group (WTR1). Transgressions 355 and regressions that are recorded within the Yellow Limestone Group represent changes in RSL of 356 only several tens of metres. The White Limestone Group sequence WTR1 indicates a longer term, 357 lower-order, transgressive sequence than the shorter time-span Yellow Limestone Group sequences. 358 359 5.1. Early Cretaceous (EKTR1)

360 The Early Cretaceous T-R cycle (EKTR1) is identified through biostratigraphic and 361 sedimentological analyses to reach average water depths of approximately 80m at the maximum 362 transgressive inflection of the RSL curve, interpreted to coincide with the Al8 (103.94 Ma) maximum 363 flooding surface (MFS) of Hardenbol et al. (1998). The regressive phase of the RSL curve from the 364 late Albian to late Turonian attains average water depths of less than 10m coinciding with the Tu4 365 (90.25 Ma) sequence boundary (SB) of Hardenbol et al. (1998). The EKTR1 sequence is represented

Page 12 of 54 366 by the Formation and Seafield Limestone. The EKTR sequence is observed within the 367 Retrieve-1 and Windsor-1 wells, and in outcrop in the Benbow Inlier. 368 The Rio Nuevo Formation contains a basal shale unit containing carinate planktonic 369 foraminifera, including Thalmanninella appenninica spp. (Fig. 4), and globular morphologies. In 370 addition, radiolaria and calcareous nannofossils are also abundant. This fossil assemblage and fine- 371 grained, laminated nature of the groundmass suggests deposition within a distal, low energy setting 372 such as an outer shelf or basinal environment. A certain degree of terrigenous input is interpreted 373 based on the occurrence of land-derived and water-transported fern spores in the palynological 374 assemblages, such as Deltoidospora spp. and Echinatisporis spp. 375 Interbedded shallow water carbonates and organic-rich mudstones of the Seafield Limestone are 376 observed in outcrop towards the top of the Rio Nuevo Formation suggesting periodic shallowing 377 sequences associated with minor oscillations within the RSL trend. Although assigned formation status 378 by Brown and Mitchell (2010), the Seafield Limestone is regarded as a member of the Rio Nuevo 379 Formation by this study due to its interbedded nature within the Rio Nuevo Formation in outcrop, 380 following the previous interpretation of Burke et al. (1969). Similar interbedded limestones and 381 organic-rich shales are observed in Albian to Cenomanian age strata in the Windsor-1 and Retrieve-1 382 wells and are interpreted to be coeval with Rio Nuevo and Seafield Limestone. Limestone interbeds 383 contain an assemblage of shallow water rudist bivalves (Fig. 4) and predominantly agglutinated 384 foraminifera indicating deposition within an oligotrophic, inner platform setting. The interbedded 385 organic-rich black shales are interpreted to correlate to highly restricted, anoxic, estuarine or lagoonal 386 sediments. The occurrence of superabundant amorphous organic matter within the black shales 387 suggests oxygen-deficient bottom conditions. 388 The overall shallowing upwards sequence of outer shelf to basinal shales, replaced by inner 389 platform Seafield Limestone interbedded with marginal marine organic-rich shales throughout the Rio 390 Nuevo Formation represents the regressive phase of the EKTR1 sequence (Fig. 4). An unconformity 391 separating Benbow Inlier strata from strata of the St. Ann’s and Lucea Inliers (Grippi, 392 1978; Mitchell et al., 2011) is identified from the Turonian and Coniacian. This unconformity lasted a 393 duration of 2.39 Ma and occurs between the Tu4 to Co1 sequence boundaries (90.25 to 87.86 Ma) 394 where part of the Dicarinella concavata micropalaeontological Zone is absent. 395 396 5.2. Coniacian-Santonian (STR1)

397 The Late Cretaceous, Coniacian-Santonian, T-R cycle (STR1) initiates at the Co1 SB of 398 Hardenbol et al. (1998), reaching average water depths of approximately 80m at the maximum 399 transgressive inflection of the RSL curve correlated to the Co1 MFS (86.92 Ma). The regressive phase 400 of the STR1 sequence culminates with the Sa3 SB (84.08 Ma) where subaerial exposure of Jamaica

Page 13 of 54 401 resulted in the formation of the early Campanian unconformity. The STR1 sequence is represented by 402 the Windsor and Clamstead Formations of the St. Ann’s Great River Inlier, the Middlesex and Dias 403 Formations of the Lucea Inlier, Retrieve-1 and Windsor-1 wells, and Sunderland Blackshop-1 and 404 Windsor coreholes. 405 Interbedded sandstone/conglomerate and shale units within the Windsor Formation are 406 interpreted as fining and deepening upwards sequences. Thick sandstone and conglomerate beds at the 407 base of the Windsor Formation are interpreted to originate as shelf sands, with possible submarine 408 volcanic material transported as sediment gravity flows causing emplacement of mudstone rip-up 409 clasts (Mitchell et al., 2011). The homogeneous nature of thick fine-grained sediments of the overlying 410 Clamstead Formation suggests continued deposition under persistent low energy conditions. The 411 abundance of carinate foraminifera including Contusotruncana fornicata, C. morozovae, Dicarinella 412 asymetrica, D. concavata, Marginotruncana marginata, M. pseudolinneiana, M. renzi, M. 413 schneegansi, M. sigali, M. sinuosa, M. tarfayaensis, M. undulata, Whiteinella aumalensis and W. 414 baltica observed within fine-grained units of the Windsor and Clamstead Formations suggest 415 deposition occurred in a distal marine, outer shelf to basinal setting associated with the Co1 MFS (Fig. 416 5). 417 Sedimentary structures such as groove and flute casts, and repetitive fining upwards Bouma 418 sequences, within the Dias and Middlesex Formations suggest these units were deposited as deep- 419 water turbidites (e.g. Grippi, 1978). The deposition of these deep-water turbidites is interpreted to be 420 coeval with sediment gravity flows within the Windsor Formation and is also interpreted to correspond 421 to the Co1 MFS within the STR1 sequence (Fig. 5). Samples of the Dias and Middlesex Formations of 422 the Lucea Inlier analysed by this study contain the nannofossils Marthasterites furcatus and Micula 423 staurophora, which indicate a Coniacian to Santonian age based on the interpretation that strata of 424 earliest Campanian age may be absent in Jamaica. 425 An unconformity at the base of the Campanian is reported by Arden Jr. (1969) and Mitchell 426 (2003). This early Campanian unconformity (Sa3 to Cam2, ca. 84.08 to 81.53 Ma) is recorded in 427 outcrop in the Lucea, Sunderland, Marchmont and St Ann’s Great River inliers where no strata of 428 earliest Campanian age were identified by this study. The presence of nannofossils restricted to Zones 429 CC18-CC19 in the Sunderland Blackshop-1 corehole suggests that nannofossil Zone CC17 is absent in 430 Jamaica (Fig. 5). It is, therefore, interpreted that the formation of the early Campanian unconformity, 431 which separates STR1 and CTR1 sequences in the Sunderland Blackshop-1 corehole, initiated at the 432 Sa3 depositional sequence boundary and lasted a duration of at least 2.55 Ma, with sedimentation not 433 resuming until the Cam2 depositional sequence. This unconformity is also recorded offshore at DSDP 434 Sites 151 and 152 where the presence of early Campanian strata remains questionable, due to scarcity 435 of both foraminifera and nannofossil marker taxa, between definitive middle Campanian and 436 Santonian age strata.

Page 14 of 54 437 438 5.3. Campanian (CTR1)

439 The Campanian T-R cycle (CTR1) initiates at the Cam2 SB (81.53 Ma), reaching average water 440 depths of 35m at the maximum transgressive inflection of the RSL curve correlated to the Cam5 MFS 441 (79.28 Ma). These water depths appear low for what has been previously interpreted as a deep-marine, 442 intra-arc basin in Jamaica at this time (Grippi, 1980; Grippi and Burke, 1980; Schmidt, 1988). It is 443 possible that bathyal settings may have been prevalent at this time although samples analysed by this 444 study lack the faunal recovery to assign greater water depths. The regressive phase of the CTR1 445 sequence culminates with the Cam9 SB (73.91 Ma), reaching average water depths of 15m before 446 further subaerial exposure of Jamaica resulted in the formation of the late Campanian – early 447 Maastrichtian unconformity (Fig. 5). The CTR1 sequence is represented by the Campanian age 448 successions of the St. Ann’s Great River, Sunderland and Marchmont Inliers, Retrieve-1, West Negril- 449 1 and Windsor-1 wells, and Marchmont, Sunderland Blackshop-1, Sunderland Potosi-2 and Windsor 450 coreholes. 451 The T-R cycle of the CTR1 sequence is best displayed within the Sunderland and Marchmont 452 Inliers. Here, the base of the John’s Hall Formation comprises a basal conglomerate that contains large 453 clasts and a foraminiferal assemblage which suggests a proximal, marginal marine depositional 454 environment. This unit contains occasional Pseudorbitoides chubbi which restricts its age to the early 455 Campanian (Jiang and Robinson, 1987; Krijnen et al., 1993) and may represent a basal conglomerate 456 lag at the beginning of the CTR1 transgressive sequence which passes upwards to deep-water shales 457 within the Johns Hall Formation. 458 The deposition of the Sunderland Formation continues the deepening upwards trend from the 459 proximal Johns Hall Formation. The presence of Planolites ichnotaxa observed within the Sunderland 460 Formation suggests deposition within the mid to distal continental shelf (Benton and Harper, 1997). 461 Globular planktonic foraminifera including Heterohelix spp. and Muricohedbergella spp. and lack of 462 carinate morphologies within this unit also indicate moderate water depths (50m-100m) of the middle 463 to outer shelf (e.g. Bé, 1977; BouDagher-Fadel, 2015). This formation represents the maximum 464 transgressive point within the CTR1 sequence, corresponding to the Cam5 MFS, and is interpreted to 465 be contemporaneous with the Drax Hall Formation of the St. Ann’s Great River Inlier (Fig. 5). 466 Planktonic foraminifera, such as Globotruncana spp. (Fig. 4), and spherical radiolaria are common 467 within the Drax Hall Formation. The presence of Globotruncana neotricarinata in the Drax Hall 468 Formation (Fig. 5) constrains the formation to a Campanian age, ranging from the Globotruncanita 469 elevata Zone to either the Radotruncana calcarata or Gansserina gansseri Zones in stratigraphic 470 sections from the Exmouth Plateau and Italy, respectively (Petrizzo et al., 2011).

Page 15 of 54 471 In the Sunderland and Marchmont Inliers, the Sunderland Formation is overlain by the 472 Newman’s Hall Formation. The Newman’s Hall Formation is described as comprising transitional 473 shelfal sands deposited between the deep water shales of the underlying Sunderland Formation and 474 overlying shallow water Stapleton Limestone (Krijnen et al., 1993). This formation is interpreted to 475 mark the start of the regressive phase of the CTR1 sequence. Based on the Newman’s Hall 476 Formation’s stratigraphic position above the Sunderland Formation, it is interpreted to be of middle 477 Campanian age. 478 The Stapleton Formation overlies the Newman’s Hall Formation to the south of the Sunderland 479 Inlier. This formation comprises fossiliferous limestone which contains a diverse assemblage of large 480 rudist bivalves including radiolitids, caprinids, plagioptychids and hippuritids, particularly specimens 481 of Whitfieldiella gigas, echinoids and larger benthic foraminifera. Limestones of the Stapleton 482 Formation contain occasional specimens of Pseudorbitoides trechmanni (Fig. 5). The presence of this 483 taxon restricts the age of the Stapleton Formation to the middle Campanian (Jiang and Robinson, 484 1987; Krijnen et al., 1993). The microfossil assemblage suggests deposition within a shallow marine 485 setting. The fragmented nature of the bioclasts indicates a moderate degree of hydrodynamic energy. 486 Therefore, it is interpreted that the Stapleton Formation was deposited in a platform margin setting 487 approaching fair-weather wave base. 488 The Stapleton Formation is interpreted to be a diachronous equivalent of the Lime Hall 489 Formation of the St. Ann’s Great River Inlier. The Lime Hall Formation contains a similar fossil 490 assemblage including rudist bivalves such as Whitfieldiella gigas and common larger benthic 491 foraminifera. The presence of Pseudorbitoides spp. observed within the Lime Hall Formation indicates 492 it is no younger than late Campanian age (Mitchell and Ramsook, 2009). Mitchell et al. (2011) note an 493 absence of Pseudorbitoides within the Lime Hall Formation which is interpreted here to be due to 494 sampling, rather than stratigraphic, reasons as this taxon is observed by this study at 18.397447°N, 495 77.198839°W. Nevertheless, a late Campanian age for the age of the Lime Hall Formation is 496 consistent with Mitchell et al. (2011). The Stapleton and Lime Hall Formations are interpreted to be 497 diachronous equivalents of the Back Formation of the Blue Mountain Block (Fig. 5). 498 Samples within the Back Rio Grande Formation contain Orbitoides megaloformis and Lepidorbitoides 499 minima (Fig. 4), the latter reported as an index fossil for the Campanian found associated with late 500 Campanian planktonic foraminifera in Europe (Aguilar et al., 2002). Therefore, the Back Rio Grande 501 Formation is interpreted to be late Campanian age. This is consistent with the age of the formation 502 assigned by Mitchell and Ramsook (2009). 503 The Stapleton, Lime Hall and Back Rio Grande Formations represent a shallow water facies at 504 the top of the St. Ann’s Great River, Sunderland and Marchmont Inliers’ regressive phase of the CTR1 505 sequence. This sequence shallows upwards from the deep water Sunderland and Drax Hall 506 Formations, whose deposition is associated with the Cam5 MFS. The late Campanian to early

Page 16 of 54 507 Maastrichtian unconformity is interpreted to occur between sequence boundaries Cam9 and Ma1 508 (73.91 to 72.05 Ma), lasting a duration of at least 1.86 Ma, where parts of the Globotruncana 509 aegyptiaca to Gansserina gansseri micropalaeontological zones are absent. 510 511 5.4. Maastrichtian 1 (MTR1)

512 The first Maastrichtian T-R cycle (MTR1) initiates at the Ma1 SB (72.05 Ma), reaching average 513 water depths of 85m at the maximum transgressive inflection of the RSL curve correlated to the Ma1 514 MFS (70.66 Ma). The regressive phase of the MTR1 sequence culminates with the Ma2 SB (69.27 515 Ma) reaching average water depths of 5m before the transgression of the second Maastrichtian T-R 516 cycle (Fig. 6). The MTR1 sequence is represented by the early to middle Maastrichtian age 517 successions of the Jerusalem Mountain, Sunderland, Marchmont, Carlton Hill and Maldon Inliers, 518 Blue Mountain Block, Cockpit-1, Content-1/A, Hertford-1, Negril Spots-1, Retrieve-1 and West 519 Negril-1 wells, and Jerusalem Mountain corehole. 520 In the Sunderland and Marchmont Inliers, the MTR1 sequence commences with the deposition 521 of the Shepherds Hall Formation. The Shepherds Hall Formation comprises a polymict, matrix- 522 supported conglomerate consisting of medium-grained sandstone supporting centimetre scale 523 subangular clasts. The Shepherds Hall Formation fines upwards towards a fine-grained, well sorted 524 sandstone with internal millimetre scale ripple cross laminations. Based on the presence of 525 conglomerates and ripple laminations, this formation is interpreted to have been deposited in a fluvio- 526 deltaic setting where conglomerates represent a basal lag at the base of the transgressive phase of the 527 MTR1 sequence. The Shepherds Hall Formation may also be a direct equivalent of the marginal 528 marine Mint Formation of the Jerusalem Mountain Inlier (Mitchell and Edwards, 2016). 529 Limestones of the Thicket River Formation and Kensington Limestone overlie the Mint and 530 Shepherds Hall Formations in the Jerusalem Mountain and Sunderland/Marchmont Inliers, 531 respectively. These limestones contain a diverse fossil assemblage including rudist bivalves, 532 particularly Trechmannites rudissimus, dasycladacean algae, miliolid foraminifera and gastropods 533 which indicate photic, oligotrophic conditions with normal marine salinities. The highly fragmented 534 nature of the bioclasts indicates high levels of hydrodynamic energy. Therefore, it is interpreted that 535 the Thicket River Formation and Kensington Limestone was deposited in an inner platform to platform 536 margin setting approaching fair-weather wave base, marking a deepening upwards trend from the 537 underlying Mint and Shepherds Hall Formations. In the Blue Mountain Block, the limestones of the 538 Thicket River Formation and Kensington Limestone are coeval with the Rio Grande Formation which 539 contains a benthic foraminiferal assemblage that includes Orbitoides megaloformis (Fig. 6), 540 Fissoelphidium operculiferum, Rotalia skourensis and Montcharmontia appenninica, indicative of a 541 Maastrichtian age.

Page 17 of 54 542 The transgressive phase of the MTR1 sequence culminates with the deposition of relatively 543 deeper water, middle to outer shelf, deposits of the Belleisle Formation in the Jerusalem Mountain 544 Inlier. The Belleisle Formation may be a lateral equivalent to planktonic foraminifera-rich Bath 545 Limestone of the Blue Mountain Block. The Bath Limestone sampled between 17.956547°N, 546 76.355358°W and 17.956318°N, 76.355317°W contains abundant globotruncanid planktonic 547 foraminifera including Abathomphalus intermedius, Contusotruncana contusa and Globotruncana 548 bulloides. This assemblage is restricted to the Pseudoguembelina palpebra - Abathomphalus 549 mayaroensis micropalaeontological Zones, confirming the absence of the Gansserina gansseri zone at 550 the Campanian-Maastrichtian unconformity. Both the Belleisle Formation and Bath Limestone mark 551 the maximum transgressive inflection of the MTR1 sequence, interpreted to correspond to the Ma1 552 MFS (Fig. 6). 553 The Belleisle Formation is overlain by the Jerusalem Formation (Mitchell and Edwards, 2016), 554 which marks the start of the shallowing upwards, regressive phase of the MTR1 sequence. The 555 Jerusalem Formation contains rudist bivalves, gastropods, oysters, algal material and larger benthic 556 foraminifera, commonly Kathina jamaicensis and Sulcoperculina angulata (Fig. 6), which are typical 557 of Maastrichtian age strata in Jamaica (e.g. Brown and Bronnimann, 1957). The regressive phase of 558 the MTR1 sequence is interpreted to follow the Ma1 MFS and terminate at the Ma2 sequence 559 boundary (Fig. 6). 560 561 5.5. Maastrichtian 2 (MTR2)

562 The second Maastrichtian T-R cycle (MTR2) initiates at the Ma2 SB (69.27 Ma), reaching 563 average water depths of 12m at the maximum transgressive inflection of the RSL curve correlated to 564 the Ma2 MFS (69.04 Ma). The regressive phase of the MTR2 sequence culminates with the Ma5 SB 565 (68.20 Ma) where terrestrially-influenced strata top the sequence (Fig. 6). The MTR2 sequence is best 566 represented by middle to late Maastrichtian age successions of the Sunderland, Marchmont and 567 Central Inliers, Cockpit-1, Content-1/A, Hertford-1, Negril Spots-1, Retrieve-1 and West Negril wells. 568 The MTR2 sequence commences with deposition of the Masemure, Slippery Rock and Thomas 569 River Formations, and the Veniella Shales. The Masemure Formation comprises well-sorted, fine- 570 grained sandstone to siltstone. Grains are sub-angular with moderate sphericity comprising quartz, 571 plagioclase feldspar and clay minerals including chlorite within calcite cement. Masemure Formation 572 sediments are therefore classified as texturally and mineralogically immature sandstones. Mitchell and 573 Edwards (2016) report that the Masemure Formation does not yield fossils, yet samples collected from 574 this formation at 18.3154°N, 78.228902°W recorded Rotalia skourensis and Heterohelix spp. (Fig. 6), 575 indicative of a Maastrichtian age. This unit shares lithological characteristics to the Veniella shales 576 which contain large bivalves including Veniella/Roudairia spp. The bivalves Veniella/Roudairia are

Page 18 of 54 577 reported from shallow marine deposits in California (Kirby and Saul, 1995). The immaturity of the 578 sandstones and microfaunal assemblage within the Masemure Formation suggests deposition within a 579 marginal, paralic, marine environment. The Masemure Formation is also correlated to the Thomas 580 River Formation (Mitchell and Edwards, 2016) and Slippery Rock Formation. Asymmetrical ripples 581 on bedding surfaces observed in the Thomas River Formation at 18.113194°N, 77.382544°W indicate 582 deposition under a unidirectional current presently directed to the north. It is interpreted that this 583 locality represents a fluvial deposit, which is further supported by the presence of freshwater 584 charophytes, particularly Platychara spp. (Fig. 6) and Azolla spp. spores (Peck and Forester, 1979; 585 Kumar and Grambast-Fessard, 1984; Kumar and Oliver, 1984). It is interpreted that the Masemure, 586 Slippery Rock and Thomas River Formations, and the Veniella Shales were deposited at the maximum 587 regressive inflection of the Jamaican RSL curve between the MTR1 and MTR2 sequences, associated 588 with the Ma2 SB. 589 The Slippery Rock and Thomas River Formations, and the Veniella Shales, are overlain by 590 rudist limestones of the Guinea Corn Formation and Titanosarcolites Beds in the Sunderland, 591 Marchmont and Central Inliers. The Titanosarcolites Beds contain a fossil assemblage characterised 592 by rudist bivalves Titanosarcolites, Thyrastylon, Trechmannites and Oligosarcolites, which are often 593 preserved in life position, dasycladacean green algae and common calcareous and miliolid 594 foraminifera, including Kathina jamaicensis and Chubbina jamaicensis (Fig. 6), indicating deposition 595 in a shallow, photic and oligotrophic marine setting. Mitchell (2005) suggests Chubbina is an index 596 fossil for the late Maastrichtian. It is interpreted that the Titanosarcolites Beds were deposited in a 597 proximal inner platform setting. 598 The Guinea Corn Formation is also dominated by a rudist bivalve assemblage including 599 Chiapasella radiolitiformis and Titanosarcolites spp. Larger benthic foraminifera including Orbitoides 600 megaloformis and Lepidorbitoides spp. are also observed. The foraminifer Orbitoides megaloformis is 601 reported from Titanosarcolites bearing strata in Jamaica ranging into the late Maastrichtian (Gunter et 602 al., 2002). Consequently, the Guinea Corn Formation and Titanosarcolites Beds are considered 603 correlative and assigned a late Maastrichtian age based on the presence of Chubbina jamaicensis. The 604 occurrence of Orbitoides megaloformis and Lepidorbitoides spp. in the Yankee River and Moravia 605 Members of the Guinea Corn Formation suggest moderate hydrodynamic energy and water depths up 606 to 20m, therefore these two members are interpreted to represent the distal inner platform close to fair- 607 weather wave base. The interbedded Two Meetings Member may represent a slight oscillation of RSL 608 into shallower water and represent paralic sediments before transgressing back into the deeper waters 609 of the Moravia Member rudist build-ups. Rare planktonic foraminifera including Globotruncanella 610 pschadae, Globotruncana ventricosa and Rugoglobigerina rugosa in samples from the Two Meetings 611 Member indicate an age no younger than the Abathomphalus mayaroensis Zone, which is consistent

Page 19 of 54 612 with the interpretation of the Cretaceous-Tertiary unconformity, where the Plummerita 613 hantkeninoides-Pseudoguembelina hariaensis Zones are absent. 614 The Titanosarcolites beds and Guinea Corn Formation are both interpreted to have been 615 deposited in an inner platform setting during a transgressive episode following the deposition of 616 marginal marine strata of the Veniella shales, Masemure, Slippery Rock and Thomas River Formations 617 which are associated with the Ma2 SB. Consequently, the Titanosarcolites beds and Guinea Corn 618 Formation represent the transgressive phase of the MTR2 sequence culminating with the Ma2 MFS. 619 The Guinea Corn Formation is overlain by the Summerfield Group in the Central Inlier. The 620 Summerfield Group comprises massive sandstones and polymict conglomerates. These conglomerates 621 occur in repetitive parallel bedded fining upwards sequences between 30 and 60 cm thick. Each 622 sequence is floored by conglomerates that fine upwards from cobble to pebble sized polymict clasts 623 that include andesites, mudstones and quartz within a sandstone matrix. The conglomerates then fine 624 into coarse and then medium grained sandstones. These conglomerates represent the Mahoe River 625 Formation within the Summerfield Group. Fining upwards successions from basal conglomerate units 626 within the Mahoe River Formation are interpreted to represent fluvial deposits of a braid delta system 627 (Mitchell, 2000). This represents a shallowing upwards sequence following deposition of the inner 628 platform Guinea Corn Formation limestones. The terrestrially influenced sediments of the Mahoe 629 River Formation represent the culmination of the regressive phase of the MTR2 sequence deposited 630 during a period of low RSL and are interpreted to be associated with the Ma5 SB. 631 The MTR2 sequence is topped by the Cretaceous-Tertiary unconformity (Ma5 to Da1, ca. 632 68.20 to 65.76 Ma), lasting a time span of at least 2.44 Ma. Biostratigraphic analyses indicate that 633 there are no Maastrichtian age sediments younger than the Abathomphalus mayaroensis Zone, where 634 the Plummerita hantkeninoides-Pseudoguembelina hariaensis zones are absent at the Cretaceous- 635 Tertiary unconformity (Fig. 6). 636 637 5.6. Paleocene – Early Eocene (PETR1)

638 The Paleocene to Early Eocene T-R cycle (PETR1) initiates at the Da1 SB (65.76 Ma), reaching 639 average water depths in excess of 100m at the maximum transgressive inflection of the RSL curve 640 correlated to the Da4 MFS (62.23 Ma). The regressive phase of the PETR1 sequence culminates with 641 the Yp9 SB (50.04 Ma) interpreted to correspond to subaerial exposure resulting in the formation of 642 the Early Eocene unconformity (Fig. 7). The PETR1 sequence is represented by Paleocene-Early 643 Eocene age successions of the Blue Mountain Block and Wagwater Trough in the east of Jamaica, and 644 Ecclesdown corehole. 645 The PETR1 sequence initiates with the deposition of the Moore Town Formation which 646 comprises well-bedded, dark grey interbedded shales and siltstones in repetitive fining upwards

Page 20 of 54 647 sequences. At the base of each graded bed, siltstones exhibit bed-parallel cylindrical trace fossils and 648 unidirectional groove structures oriented 030. The siltstones grade normally to parallel and planar 649 shale beds which exhibit internal laminations. The fining upwards siltstone and mudstone packages 650 that exhibit sole structures are interpreted to represent bathyal turbidite deposits. Outer ramp to bathyal 651 water depths are interpreted to have occurred during a period of high RSL, tentatively correlated to the 652 Da4 MFS. The Moore Town Formation is reported to be Danian to Selandian age by Fluegeman 653 (1998). Fluegeman (1998) assigned the Moore Town Formation to micropalaeontological Zones P1c- 654 P3a, and to nannofossil Zones NP4-NP5. 655 The Wagwater Formation stratigraphically overlies the Moore Town Shales. It comprises 656 various lithologies including polymict conglomerates and sandstones with limestone lenses. The 657 limestone lenses contain a fossil assemblage including bivalves, echinoids, ostracods and calcareous, 658 agglutinated and miliolid benthic foraminifera. The foraminifer Ranikothalia catenula is observed 659 (Fig. 7). This taxon is described from the Wagwater Formation and is reported to represent a Late 660 Paleocene, Thanetian, age (Ramsook and Robinson, 2009). Based on the Ranikothalia limestone 661 observed along the , the Wagwater Formation is assigned an age within the range of 662 micropalaeontological Zones P4b-P5. The Wagwater Formation represents a shallowing upwards, 663 regressive phase of the PETR1 sequence following the deposition of the deep-water turbidites of the 664 Moore Town Formation. 665 The Wagwater Formation is succeeded by the Richmond Formation in the Blue Mountain 666 Block. The Richmond Formation comprises parallel and horizontally bedded fining upwards 667 sequences of medium- to fine-grained sandstones, siltstones and shales. The Richmond Formation 668 contains age-diagnostic palynomorphs including Diphyes colligerum, Homotryblium tenuispinosum 669 and Cyclopsiella ‘reticulata’ indicating an age between Early Eocene, intra-Ypresian, and Middle 670 Eocene, Lutetian. Consequently, the age of the Richmond Formation is interpreted to be Ypresian. It is 671 interpreted that the alternating sandstone and shale beds of the Richmond Formation were deposited as 672 fan-delta to submarine-fan deposits as coarse sediments were shed from the highlands of the Blue 673 Mountain Block into prograding deltas encroaching a steep submarine slope (Wescott and Ethridge, 674 1983). 675 The PETR1 sequence is topped by the Early Eocene unconformity to the east of Jamaica (Yp9 676 to Yp10, ca. 50.04 to 49.58 Ma), lasting a duration of at least 0.46 Ma (Fig. 7). 677 678 5.7. Yellow Limestone Group 1, Ypresian-Lutetian (YTR1)

679 The Yellow Limestone Group of Jamaica is divided into three T-R cycles (YTR1-3). The basal 680 YTR1 cycle is equivalent to the lower cycle of Maharaj and Mitchell (2000) and the Stettin cycle of 681 Matchette-Downes and Mitchell (2005). The YTR1 cycle initiates at the Yp10 SB (49.58 Ma),

Page 21 of 54 682 reaching average water depths of 15m at the maximum transgressive inflection of the RSL curve 683 correlated to the Lu1 MFS (46.82 Ma). The regressive phase of the YTR1 sequence culminates with 684 the Lu3 SB (43.61 Ma). The YTR1 sequence is represented by Early to Middle Eocene age 685 successions (Fig. 7) of predominantly the Clarendon Block of central Jamaica, Content-1/A, Hertford- 686 1 and Pedro Bank-1 wells, and Blowfire Hill corehole. 687 The New Ground Formation is here included as a basal unit of the lower Yellow Limestone 688 Group T-R cycle (YTR1); this is in contrast to Mitchell et al. (2011) who make no comment to its 689 stratigraphic position within the Paleogene successions of Jamaica. This formation marks the most 690 proximal depositional setting before the transgression of later marine-influenced sediments. The New 691 Ground Formation comprises matrix-supported, polymict conglomerates that contain moderately 692 sorted, sub-angular, low sphericity grains that attain a maximum diameter of 6mm. The matrix consists 693 of a medium-grained, texturally and mineralogically immature sandstone. The conglomerate fines 694 upwards to fine-grained, well-sorted sandstones and siltstones that are texturally and mineralogically 695 mature. The fining upwards cycles within the New Ground Formation were lain down in episodes of 696 initial high energy, depositing conglomeratic units and subsequent finer grains as the initial energy 697 was waning. This is interpreted to have been deposited within a fluvial system. The age of this 698 formation is given as Early Eocene by Mitchell et al. (2011). 699 The organic-rich Litchfield Formation (Section 5.8) was previously termed the ‘Guys Hill 700 Formation’, but subsequently renamed after sediments at the Guys Hill type locality were found to be 701 Early Eocene in age (Mitchell, 2016). The ‘true’ Guys Hill Formation described here refers to 702 sediments similar in appearance to the Litchfield Formation (Section 5.8), but are older and exposed 703 only in a small area close to the town of Guys Hill, west of the Benbow Inlier and to the east of the 704 Clarendon Block. The Guys Hill Formation comprises well bedded, arkosic sandstones that are coarse 705 to medium-grained, containing well-rounded and well-sorted grains, and occasional lignitic laminae. 706 The sandstones display planar cross bedding, with low angle foresets dipping towards the north. The 707 planar bedded sandstones are overlain by a trough cross-bedded sands that contain climbing foresets 708 which grade upwards into nodular impure limestone. Within this limestone, the presence of the benthic 709 foraminifera Eoconuloides wellsi and Fabularia colei and the rare occurrence of the planktonic 710 foraminifer Globanomalina chapmani constrain the age of the Guys Hill Formation to 711 micropalaeontological Zone E4, Ypresian. This is supported by the nannofossil Discoaster kuepperi 712 which is assigned to Zones NP14a-NP11, within the Ypresian. The textural immaturity of the 713 siliciclastic components and highly fragmentary nature of the bioclasts within the Guys Hill Formation 714 suggests deposition in a proximal, high energy, marine setting. The presence of planar cross-beds may 715 indicate foreshore deposits with seaward inclined laminae directed towards the north. These are 716 overlain by trough cross-bedded, upper shoreface deposits. This suggests a transgressive sequence that 717 deepens upwards towards shallow marine sediments of the overlying Freeman’s Hall and Stettin

Page 22 of 54 718 Formations. The transition from the Guys Hill to Freeman’s Hall and Stettin Formations represents the 719 transgressive phase of YTR1 following the Yp10 sequence boundary. 720 The Freeman’s Hall and Stettin Formations consist of conglomerates, sandstones, mudrocks 721 and irregularly bedded, flaggy limestones that contain a mix of wackestones, packstones and 722 grainstones. These units yield an abundant and diverse fossil assemblage that includes corals, molluscs 723 (including large lucinid bivalves) and large benthic foraminifera. The diverse microfossil assemblage 724 within these limestones (Fig. 7) include the larger benthic foraminifera Discorinopsis gunteri, 725 Fallotella cookei, Fabiania cassis, Fabularia colei, Coskinolina douvillei, Fabiania cassis, 726 Amphistegina parvula, Eoconuloides wellsi, Verseyella jamaicensis and Helicostegina gyralis, and the 727 planktonic foraminifera Morozovella crater which restrict the age of these units to 728 micropalaeontological Zones E8-E9. The assemblage, including common miliolid and agglutinated 729 foraminifera plus significant micrite within packstones and wackestones, indicates deposition within 730 low energy, oligotrophic waters of the proximal inner ramp. The Freeman’s Hall and Stettin 731 Formations are overlain by foraminiferal grainstones which are dominated by the benthic foraminifera 732 Helicostegina gyralis. The foraminiferal grainstones are constrained to micropaleontological Zone E9 733 based on the presence Eoconuloides lopeztrigoi, Helicostegina gyralis and Pseudolepidina trimera 734 (Fig. 7). The grainstones which contain large amounts of spar, robust lepidocyclinid foraminifera, 735 highly fragmented bioclasts and encrusting coralline red algae indicate a high energy depositional 736 setting and may represent a foraminiferal shoal at fair-weather wave base at the distal inner ramp. It is 737 interpreted that the Freeman’s Hall and Stettin Formations, and the foraminiferal grainstones, were 738 deposited during the transgressive phase of the YTR1 sequence, which culminates with the Lu2 MFS. 739 In the Maldon Inlier, the Stettin Formation is unconformably overlain by the Litchfield Formation 740 along an unconformity that is interpreted to correspond to the Lu3 SB (Fig. 7). 741 742 5.8. Yellow Limestone Group 2, Lutetian-Bartonian (YTR2)

743 The second Yellow Limestone Group T-R cycle (YTR2) is equivalent to the upper cycle of 744 Maharaj and Mitchell (2000) and the Chapelton-Troy cycle of Matchette-Downes and Mitchell (2005). 745 The YTR2 cycle initiates at the Lu3 SB (43.61 Ma), reaching average water depths of 75m at the 746 maximum transgressive inflection of the RSL curve correlated to the Lu4 MFS (40.40 Ma). The 747 regressive phase of the YTR2 sequence culminates with the Bart1 SB (39.17 Ma). The YTR2 748 sequence is represented by Middle Eocene age successions (Fig. 7) of the Hanover and Clarendon 749 Blocks of western and central Jamaica, Arawak-1, Content-1/A, Negril Spots-1, Pedro Bank-1, 750 Portland Ridge-1 and Santa Cruz-1 wells, and Elderslie and Pindars River-3 coreholes. 751 The Litchfield Formation is found within the Maldon, Marchmont and Central Inliers in the 752 Clarendon Block of Jamaica. This formation contains dark, organic-rich shales and was previously

Page 23 of 54 753 termed the ‘Guys Hill Formation’, but subsequently renamed after sediments at the Guys Hill type 754 locality were found to be Early Eocene in age (Mitchell, 2016). The Litchfield Formation is observed 755 to comprise weakly cemented, porous sandstones and lignite-rich dark mudstones in fining-upwards 756 sequences. In the Maldon Inlier, the Litchfield Formation rests unconformably on top of the Stettin 757 Limestone, where occasionally it slumps into sinkholes developed in the karstic boundary between the 758 two units. The Litchfield Formation is characterised by the presence of high numbers of Lanagiopollis 759 crassa, which have been documented from the Middle Eocene Avon Park Formation in west central 760 Florida (Jarzen and Dilcher, 2006). A Middle Eocene age is supported by the occurrence of 761 Margocolporites vanwijhei and Monoporites annulatus, both indicating an age no older than Middle 762 Eocene in the Caribbean region, and Spinizonocolpites echinatus, which indicates an age no younger 763 than Eocene (Germeraad et al., 1968). Additionally, the co-occurrence of other taxa such as 764 Spirosyncolpites spiralis, Psilatricolporites caribbiensis, Retitricolporites irregularis, 765 Spinizonocolpites baculatus, Polysphaeridium subtile, Homotryblium spp. and Operculodinium spp. is 766 fully consistent with this age assignment (González-Guzmán, 1967; Jaramillo et al., 2011). The high 767 abundances of the miospore Lanagiopollis crassa reflect input from significant developments of 768 mangrove-like vegetation along the adjacent coastline. Consequently, it is interpreted that the 769 Litchfield Formation was deposited within a marginal marine, paralic, setting. This supports Robinson 770 and Mitchell’s (1999) interpretation that this formation was deposited within a tide-dominated 771 estuarine complex. The Litchfield Formation is assigned an age correlated with the E9 772 micropalaeontological Zone based on its stratigraphic position between the Stettin and Chapelton 773 Formations. Consequently, deposition of the Litchfield Formation is interpreted to correspond to a 774 period of low RSL associated with the Lu3 SB at the base of the YTR2 sequence. The transgressive 775 phase of the YTR2 sequence initiates with paralic and terrestrially-influenced deposits of the 776 Litchfield Formation, transitioning to more marine sediments of the Chapelton Formation. In tropical 777 areas of West Africa (Gulf of Guinea) and SE Asia (South China Sea), it has been shown that 778 ecosystems containing the palynological assemblage listed above spread during periods of increasing 779 sea-level (Poumot, 1989; Morley, 1991). 780 The Chapelton Formation overlies the Litchfield Formation within the Yellow Limestone 781 Group. This study recognises two sub-divisions of the Chapelton Formation, herein referred to the 782 lower and upper Chapelton Formation. The base of the Chapelton Formation consists of a mollusc-rich 783 coquina that contains large, thick-walled bivalves, gastropods and larger benthic foraminifera. The 784 coquina is succeeded by a foraminiferal packstone, which is dominated by miliolid foraminifera, 785 including Yaberinella jamaicensis. Lepidocyclinids are absent from the lower part of the Chapelton 786 Formation. The lower Chapelton Formation is constrained to micropalaeontological Zones E9-10 787 based on the presence of Verseyella jamaicensis and Yaberinella hottingeri. Consequently, the lower 788 division of the Chapelton Formation is entirely middle to late Lutetian age.

Page 24 of 54 789 The upper Chapelton Formation comprises massive, nodular, fossiliferous packstone. It 790 contains abundant molluscan material, miliolids and lepidocyclinids including Eulepidina spp. and 791 Polylepidina spp. which are diagnostic of the upper half of Chapelton Formation. Lepidocyclinids 792 increase in abundance, diversity and morphological complexity towards the top of the Chapelton 793 Formation. The upper Chapelton Formation is constrained to micropalaeontological Zones E10-12 794 based on the presence of Pseudophragmina advena, Lepidocyclina (Lepidocyclina) macdonaldi, 795 Fabularia vaughani/gunteri, Coleiconus christianaensis, Yaberinella jamaicensis and Polylepidina 796 chiapasensis. Therefore, the upper Chapelton Formation is assigned a late Lutetian to early Bartonian 797 age. The disappearance of the foraminifera Polylepidina chiapasensis (Fig. 7) marks the top of the 798 Chapelton Formation. 799 The Chapelton Formation displays a deepening upwards sequence, marking a lateral facies 800 transition from paralic sediments, that share similarities to the Litchfield Formation, to shoreline 801 mollusc-rich coquinas and quiet water, proximal inner ramp deposits dominated by miliolid 802 foraminifera that are typical of much of the lower part of the formation. The upper part of the 803 Chapelton Formation is interpreted to have been deposited in a distal inner ramp setting close to an 804 energetic foraminiferal shoal. The Chapelton Formation is described as having been deposited during a 805 north to south transgression over Jamaica (Robinson and Mitchell, 1999). This is evident from the 806 relatively deeper water facies which contain orthophragminid foraminifera, such as Pseudophragmina 807 advena (Fig. 7), that occur close to the north coast. 808 The Healthy Hill Formation overlies the Chapelton Formation (Mitchell, 2013; 2016). It 809 comprises a pink-coloured foraminiferal grainstone which contains large miliolids including 810 Yaberinella jamaicensis and is dominated by large lepidocyclinid foraminifers including Eulinderina 811 spp. and Polylepidina chiapasensis. The foraminifer Polylepidina chiapasensis (Fig. 7) is restricted to 812 micropalaeontological Zones E10-E12 therefore the Healthy Hill Formation is assigned a late Lutetian 813 to early Bartonian age. The depositional setting and microfaunal assemblage of the Healthy Hill 814 Formation is similar to the upper Chapelton Formation. However, it is interpreted that the Healthy Hill 815 Formation was deposited within an energetic foraminiferal shoal that transgressed diachronously 816 across the top of the upper Chapelton Formation. This represents the transgressive phase of the YTR2 817 sequence. 818 The Preston Hill Marl Member of the Font Hill Formation represents a deep-water lateral 819 facies transition from the relatively shallower water Chapelton and Healthy Hill Formations. In places, 820 the Preston Hill Marl is observed to transgress across the top of the Chapelton and Healthy Hill 821 Formations. This unit is classified as a planktonic foraminiferal carbonate mudstone and comprises 822 beige-coloured marl, consisting of abundant planktonic foraminifera and echinoid spines visible in 823 hand specimen. The Preston Hill planktonic foraminiferal marl contains abundant carinate planktonic 824 foraminifera including Morozovelloides spp. and thick-walled taxa including Globigerinatheka spp.

Page 25 of 54 825 These morphologies are indicative of outer ramp to upper bathyal deposits (Bé, 1977). The Preston 826 Hill Marl also contains the planktonic foraminifer Orbulinoides beckmanni which is restricted to 827 micropalaeontological Zone E12 and is often found in maximum flooding surface deposits associated 828 with the Lu4 MFS. Consequently, the deposition of the Preston Hill Marl is interpreted to mark the 829 maximum transgressive inflection of the YTR2 sequence corresponding with the Lu4 MFS (Fig. 7). 830 Regression following deposition of the Preston Hill Marl culminated with the formation of the Bart1 831 sequence boundary. 832 833 5.9. Yellow Limestone Group 3, Bartonian-Priabonian (YTR3)

834 The final Yellow Limestone Group T-R cycle (YTR3) initiates at the Bart1 SB (39.17 Ma), 835 reaching average water depths of 15m at the maximum transgressive inflection of the RSL curve 836 correlated to the Bart1 MFS (38.46 Ma). The regressive phase of the YTR3 sequence culminates with 837 the Pr1 SB (37.75 Ma) which separates the Yellow and White Limestone Groups (Fig. 7). The YTR3 838 sequence is represented by latest Middle Eocene age successions of the Hanover and Clarendon 839 Blocks of western and central Jamaica, Arawak-1, Content-1/A, Negril Spots-1, Portland Ridge-1 and 840 Santa Cruz-1 wells. 841 The YTR3 sequence is represented by the Ipswich Formation which unconformably overlies the 842 Preston Hill Marl, Healthy Hill and Chapelton Formations. This formation is the uppermost unit of the 843 Yellow Limestone Group and is unconformably overlain by the White Limestone Group. The lower 844 part of the Ipswich Formation comprises a pale orange to beige-coloured foraminiferal packstone 845 containing a large amount of micrite. This lower unit contains abundant miliolids including large 846 specimens of Yaberinella jamaicensis and Fabularia hanzawai. The lower part of the Ipswich 847 Formation is constrained to micropalaeontological Zones E13-E14, based on the highest occurrences 848 of Fabularia vaughani/gunteri and Cushmania americana and the presence of Fabularia hanzawai 849 which is restricted to these two biozones. The large number of miliolids and agglutinated foraminifera, 850 together with a large amount of micrite matrix, indicates this unit was deposited in the low energy, 851 oligotrophic waters of the proximal inner ramp. Continued transgression resulted in the deposition of 852 the overlying distal inner ramp deposits of the upper part of the Ipswich Formation within the YTR3 853 sequence. 854 The upper part of the Ipswich Formation is observed to comprise a grey coloured foraminiferal 855 mud-lean packstone to grainstone. This unit is dominated by an abundant and diverse lepidocyclinid 856 assemblage including Lepidocyclina (Lepidocyclina) macdonaldi, L. (L.) pustulosa and L. (L.) ariana, 857 and rare specimens of large Yaberinella jamaicensis. The presence of Fabularia hanzawai, 858 Polylepidina proteiformis and Lepidocyclina (Lepidocyclina) macdonaldi restrict the age of the upper 859 part of the Ipswich Formation to micropalaeontological Zones E13-E14. The presence of

Page 26 of 54 860 Lepidocyclina (Lepidocyclina) ariana in some samples is restricted to Zone E13. The dominance of 861 lepidocyclinids and large amount of calcite spar within these grainstones indicate it was deposited in 862 an energetic environment such as a foraminiferal shoal at fair weather wave base of the distal inner 863 ramp. This represents a minor deepening trend following on from the proximal inner ramp deposits of 864 the lower part of this formation. This deepening trend marks the transgressive phase of the YTR3 865 sequence, culminating with the Bart1 MFS. Subsequent regression terminates at the Pr1 SB and base 866 of the overlying White Limestone Group (Fig. 7). 867 868 5.10. White Limestone Group, Priabonian-Rupelian (WTR1)

869 The White Limestone Group T-R cycle (WTR1) initiates at the Pr1 SB (37.75 Ma), reaching 870 average water depths of approximately 75m. The WTR1 sequence is represented by Late Eocene- 871 Oligocene age successions (Fig. 7) of Jamaica, Arawak-1, Negril Spots-1, Pedro Bank-1, Portland 872 Ridge-1 and Santa Cruz-1 wells, and Ecclesdown corehole. Only one T-R cycle was recorded within 873 the White Limestone Group by this study due to sampling reasons, however with higher resolution 874 sampling further sequences may be determined. 875 The Troy Formation unconformably overlies the Yellow Limestone Group and is the basal unit 876 of the White Limestone Group. It comprises a dolomitic limestone that is black in colour on its 877 weathered surface and pale orange/beige in colour on its fresh surface. Samples from the Troy 878 Formation display pervasive dolomitisation with loss of original fabric; however rare exceptions retain 879 ‘ghosts’ of larger benthic foraminifera. Consequently, samples analysed for biostratigraphy yielded 880 poor results due to the presence of pervasive dolomitisation. However, due to the Troy Formation’s 881 stratigraphic position between two well constrained units (the Ipswich and Swanswick Formations) it 882 is assigned to micropalaeontological Zone E14. ‘Ghosts’ of conical dictyoconids and presumed 883 abundance of fine-grained micrite prior to dolomitisation suggest the Troy Formation was deposited 884 within very shallow, low energy, oligotrophic waters of the proximal inner ramp during a period of 885 low RSL associated with the Pr1 depositional SB. 886 The Swanswick and Claremont Formations overlie the Troy Formation within the White 887 Limestone Group succession. These units are observed to comprise massive, white coloured bioclastic 888 grainstone. They contain an abundant and diverse foraminiferal assemblage dominated by the small 889 rotaliid Neorotalia mexicana and large morphologically advanced lepidocyclinids. The presence of 890 Lepidocyclina (Nephrolepidina) yurnagunensis within the Swanswick Formation indicates an age no 891 older the Zone E16. Consequently, the Swanswick Formation is assigned a Late Eocene, late 892 Priabonian, age and continues the deepening upwards, transgressive trend of the WTR1 sequence 893 transitioning from the proximal inner ramp deposits of the Troy Formation. The presence of large platy 894 lepidocyclinids, encrusting red algae and grainstone facies suggest that the Swanswick and Claremont

Page 27 of 54 895 Formations were deposited in a distal inner ramp setting at fair-weather wave base where high 896 hydrodynamic energy favours the precipitation of coarse calcite spar. 897 The Swanswick and Claremont Formations are overlain by the Montpelier Formation, an 898 extremely thick and widespread unit found across most of Jamaica. The Montpelier Formation 899 comprises interbedded chalk and chert, interpreted to represent deep-water calciturbidites deposited 900 above the Calcite Compensation Depth (CCD) in water depths in excess of 200m (Underwood and 901 Mitchell, 2004). However, these turbidites must have been deposited a short distance from the shelf to 902 receive transported (allochthonous) shallow water taxa including Lepidocyclina (Lepidocyclina) 903 canellei, Lepidocyclina (Eulepidina) undosa, Heterostegina antillea and Miogypsinoides spp., sourced 904 from the contemporaneous Brown’s Town Formation that was deposited upslope of the Montpelier 905 Formation. The presence of Lepidocyclina (Lepidocyclina) cannellei, Lepidocyclina (Nephrolepidina) 906 yurnagunensis, Heterostegina antillea and Nummulites fichteli in samples from the Montpelier 907 Formation constrain this unit to Zones O3-O5. These biozones represent the ‘late’ Rupelian and ‘early’ 908 Chattian stages of the Oligocene. The presence of the foraminifera Lepidocyclina (Nephrolepidina) 909 yurnagunensis, Heterostegina antillea, Nummulites fichteli, Lepidocyclina (Lepidocyclina) canellei, 910 Lepidocyclina (Eulepidina) undosa (Fig. 7) within the Brown’s Town Formation are also restricted to 911 the Oligocene, micropalaeontological Zone O5. This corroborates the age of the transported units 912 within the calciturbidites observed within the Monteplier Formation. The deposition of the Montpelier 913 Formation is associated with the Ru3 MFS and marks the culmination of the WTR1 transgressive 914 sequence that initiated at the Pr1 sequence boundary with the deposition of the Troy Formation (Fig. 915 7). 916 917 6. Discussion

918 It has been described that both the Nicaraguan Rise and northern Bahamas have been influenced 919 by global trends in eustasy, particularly within the Neogene (Hine et al., 1992).The Jamaican RSL 920 curve presented by this study (Fig. 8) indicates that both local tectonics and global eustasy were 921 important factors in the deposition of regional sequences since the Cretaceous. It is interpreted that 922 large unconformities (duration >1 Ma), quantified through biostratigraphic analyses, are related to 923 significant regional tectonic events, with minor overprint of a global eustatic signal, while smaller 924 unconformities (duration <1 Ma) are produced by global eustatic trends. The signature of the tectonic 925 events is most apparent when the Jamaican RSL curve is compared to global sea-level trends (Fig. 8) 926 of Haq and Al-Qahtani (2005), Kominz et al. (2008), Müller et al. (2008), Snedden and Liu (2010) and 927 Miller et al. (2005; 2011). Many of the depositional sequences initiate during a period of low RSL, 928 attributed to tectonic uplift, and are followed by transgressive episodes influenced by global eustasy or 929 tectonic subsidence (Fig. 8).

Page 28 of 54 930 931 6.1. Tectonic controls

932 Four Jamaican unconformities lasted durations in excess of 1 Ma. The Turonian-Coniacian 933 unconformity between the EKTR1 and STR1 sequences lasted a duration of at least 2.39 Ma. This 934 unconformity is interpreted to have formed during regional uplift associated with the assembly of 935 CLIP fragments (Burke et al., 1978; Burke, 1988; Mitchell, 2006; Hastie et al., 2010a; 2010b). The 936 regressive phase of the EKTR1 sequence leading up to the formation of the Turonian-Coniacian 937 unconformity is interpreted to have been deposited as RSL decreased during regional uplift. This 938 conflicts with the global eustatic trend of Haq and Al-Qahtani (2005) which shows increasing global 939 sea-level during the same period (Fig. 8), implying a tectonic control on the EKTR1 sequence. 940 Following the formation of the Turonian-Coniacian unconformity, the transition to deep-water 941 sedimentation of the STR1 sequence is attributed to rapid subsidence and increase in the rate of RSL 942 rise associated with intra-arc rifting (Mitchell, 2003). The early Campanian unconformity between the 943 STR1 and CTR1 sequences lasted a duration of at least 2.55 Ma and is attributed to jarring of buoyant 944 CLIP against the subduction zone of the Great Arc of the Caribbean (Burke et al., 1978; Mitchell, 945 2003; Hastie et al., 2010a; 2010b) resulting in regional uplift and producing the regressive phase of the 946 STR1 sequence. 947 The regressive phase of the CTR1 sequence also conflicts with the trend of increasing global 948 sea-level over the same time period of Haq and Al-Qahtani (2005). This localised regression is 949 attributed to the regionally significant collision of the Chortis Block and volcanic arc with the CLIP 950 (Mitchell 2003; 2006; Rogers et al., 2007; Carvajal-Arenas et al., 2015; Sanchez et al., 2016) resulting 951 in compressional shortening and collisional uplift expressed as a series of thrusts onshore Jamaica and 952 by the late Campanian to early Maastrichtian unconformity (Mitchell 2003; 2006). The duration of this 953 unconformity is calculated as 1.86 Ma between the CTR1 and MTR1 sequences. 954 The Cretaceous-Tertiary unconformity between the MTR2 and PETR1 sequences is calculated 955 to have lasted a duration of at least 2.44 Ma and is attributed to the collision of the Caribbean Plate 956 with the Bahamas and suturing of the Gonâve Microplate to the rest of Jamaica along the Plantain 957 Garden Fault Zone (Arden Jr, 1969; 1975; Mann et al., 1995; Calais, 1999; McCann, 1999; James- 958 Williamson et al., 2014). Following this event, extension related to sinistral strike-slip movement 959 between the North American Plate and the northern margin of the Caribbean Plate initiated in the 960 Paleocene (Pindell, 1994; Mitchell, 2003). This Paleocene extension caused significant regional 961 tectonic subsidence resulting in the formation of the Montpelier-Newmarket and Wagwater Troughs 962 (Draper, 1987; Mann and Burke, 1990; James-Williamson et al., 2014; Carvajal-Arenas et al., 2015). 963 This tectonically influenced regional subsidence is recorded in the Jamaica by a significant increase in 964 RSL that contrasts to a contemporaneous fall in global sea-level (Fig. 8) recorded by Haq and Al-

Page 29 of 54 965 Qahtani (2005), Kominz et al. (2008), Müller et al. (2008), Snedden and Liu (2010) and Miller et al. 966 (2005; 2011). 967 The Yellow and White Limestone Groups are separated by an angular unconformity which is 968 represented by a palaeokarstic surface formed during subaerial exposure and interpreted to be tectonic 969 in origin (Mitchell, 2016). According to Mitchell (2016) this unconformity youngs to the western 970 margin of the Clarendon Block, lasting a duration of at least 4 Ma. Mitchell (2016) suggests that the 971 youngest unit of the Yellow Limestone Group, the Ipswich Formation, is only present in the western 972 part of the Clarendon Block. This study identifies limestone of a similar age (early E14) and facies 973 (distal inner ramp grainstones) to the Ipswich Formation farther east (18.394091°N, 77.606896°W; 974 18.398820°N, 77.205972°W). Consequently, the unconformity dividing the Yellow and White 975 Limestone Groups is calculated by this study to have lasted a duration of <1 Ma, restricted to the 976 Bartonian part of the E14 micropalaeontological biozone. A typical limestone dissolution rate of 977 100m3/km2/annum (Gunn, 1981; Gabrovšek, 2009) is equivalent to the removal of a 10cm depth of 978 limestone per square kilometre per year. This suggests that the palaeokarst surface between the Yellow 979 and White Limestone Groups need not take millions of years to form. However, given the angular 980 nature of the unconformity it is interpreted by this study to be undoubtedly tectonic in origin, possibly 981 related to the reactivation of extensional to reverse faults, and may record both a tectonic and eustatic 982 signature where tectonic uplift and subaerial exposure is contemporaneous with global sea-level fall, 983 attributed to the Pr1 sequence boundary. 984 985 6.2. Eustatic controls

986 While local tectonism is shown to exert a significant control on the deposition of Jamaican 987 sequences, particularly within the Cretaceous, global eustasy is interpreted to be an important 988 controlling factor in the deposition of Maastrichtian and Paleogene sequences. The signature of 989 tectonic episodes, particularly collisional events, is best expressed by the regressive phases of the 990 Jamaican depositional sequences culminating in the formation of major unconformities. However, it is 991 interpreted that some of the maximum transgressive and regressive inflections of RSL correlate to 992 globally recognised sequence stratigraphic surfaces of Hardenbol et al. (1998) particularly where these 993 events are well constrained to short-lived biozones. The best examples of this are the maximum 994 transgressive inflection of the YTR2 sequence which is constrained to the E12 biozone, based on the 995 presence of Orbulinoides beckmanni, corresponding to the Lu4 MFS and the maximum regressive 996 inflection of the RSL curve between the YTR1 and YTR2 sequences, responsible for deposition of the 997 Litchfield Formation, constrained to the E9 biozone that corresponds to the Lu3 SB (Fig. 8). 998 Consequently, where depositional sequences are bound by unconformities lasting a duration of less

Page 30 of 54 999 than 1 Ma, it is interpreted that these unconformities correlate with global sequence boundaries formed 1000 during periods of eustatic sea-level fall. 1001 The Maastrichtian sequences, MTR1-2, are the most strongly influenced by eustasy of the 1002 Cretaceous sequences. The MTR1 sequence correlates with a complete cycle of sea-level rise and fall 1003 recorded by the short-term eustatic trend of Haq and Al-Qahtani (2005) between the Ma1 and Ma2 1004 sequence boundaries (Fig. 8). The MTR2 sequence also correlates to the short-term rise and fall of 1005 Kominz et al. (2008), Snedden and Liu (2010), and Miller et al. (2005; 2011) eustatic trends between 1006 the Ma2 and Ma5 sequence boundaries (Fig. 8). 1007 The Early Eocene unconformity between the PETR1 and YTR1 sequences is calculated to have 1008 lasted a duration of at least 0.46 Ma. This unconformity coincides with a significant drop in the short- 1009 term mean global sea-level trend, between 60m and 110m, recorded by Haq and Al-Qahtani (2005) 1010 and Snedden and Liu (2010) correlated with the Yp10 sequence boundary. 1011 Minor oscillations in the Jamaican RSL curve within the Yellow Limestone Group correlate 1012 with oscillations in the short-term sea-level trend of Haq and Al-Qahtani (2005). The maximum 1013 transgressive inflection of the YTR1 sequence correlates with peaks in the short-term mean sea-level 1014 trend of Haq and Al-Qahtani (2005), Kominz et al. (2008), Snedden and Liu (2010), and Miller et al. 1015 (2005; 2011). Similarly, the maximum transgressive inflection of the YTR2 sequence coincides with 1016 significant maximum transgressive inflections of the short-term mean global sea-level curves of Haq 1017 and Al-Qahtani (2005), Kominz et al. (2008) and Snedden and Liu (2010), correlated to the Lu4 MFS 1018 (Fig. 8). The final Yellow Limestone Group sequence, YTR3, also correlates with a complete cycle of 1019 sea-level rise and fall recorded by the short-term eustatic trend of Haq and Al-Qahtani (2005) and 1020 Snedden and Liu (2010) between the Bart1 and Pr1 sequence boundaries (Fig. 8). 1021 1022 6.3. Magnitude of Jamaican stratigraphic cycles

1023 Transgressive-regressive cycles identified within the Jamaican sea-level curve represent second 1024 (107 years) to fourth (105 years) order sequences, although most tie to globally recognised third order 1025 (106 years) sequences of Hardenbol et al., (1998). The EKTR1 and PETR1 sequences are regarded as 1026 second order sequences, lasting a duration of approximately 23 Ma and 16 Ma, respectively. The 1027 duration of these sequences is consistent with long term regional tectonic events including the 1028 assembly of CLIP fragments and back/fore-arc spreading in the ‘Middle’ Cretaceous, and Late 1029 Cretaceous to Paleogene rifting across the Nicaraguan Rise. The STR1 (ca. 4 Ma), CTR1 (ca. 7.5 Ma), 1030 MTR1 (ca. 3 Ma), YTR1 (ca. 5.5 Ma), YTR2 (ca. 4.5 Ma) and WTR1 (ca. 9.25 Ma) depositional 1031 cycles represent third order sequences and are regarded to be controlled by global eustatic trends. The 1032 MTR2 (ca. 1 Ma) and YTR3 (ca. 1.25 Ma) sequences may represent fourth order cycles.

Page 31 of 54 1033 In carbonate dominated sequences of the Maastrichtian, and Paleogene Yellow and White 1034 Limestone groups the average rates of RSL rise are calculated as approximately 26.6 m/Ma for MTR1, 1035 10 m/Ma for MTR2, 3.6 m/Ma for YTR1, 15.5 m/Ma for YTR2, 4 m/Ma for YTR3 and 7 m/Ma for 1036 WTR1. Given the lowest accumulation rates of prograding carbonates is regarded as 30 m/Ma 1037 (Schlager, 1981; Handford and Loucks, 1993; Wissler et al., 2003), production rates of Maastrichtian 1038 and Paleogene carbonates could sufficiently keep pace with the low rates of RSL rise calculated for 1039 these sequences. This accounts for the several kilometres thick carbonate successions of Maastrichtian 1040 and Paleogene age units observed in Jamaica. However, the sharp change in gradient in the WTR1 1041 sequence in the late Rupelian is calculated to require an increase in the rate of RSL rise to 27.5 m/Ma. 1042 Although optimal carbonate accumulation rates should keep pace with this rate of RSL rise, Jamaican 1043 carbonate platforms nevertheless succumbed to drowning during the Oligo-Miocene evident by thick 1044 sequences of the deep-water Montpelier Formation overlying shallow platform carbonates across the 1045 island. It is therefore interpreted that environmental factors (e.g. Mutti et al., 2005) exerted a greater 1046 control than global eustasy later on during the WTR1 sequence. 1047 1048 7. Conclusions

1049 The ten Jamaican T-R sequences identified by this study each correspond to one complete 1050 cycle of RSL rise and fall. Comparisons of the Jamaican RSL curve with global trends show that local 1051 tectonism exerts a significant control on the deposition of sedimentary sequences in Jamaica. 1052 Consequently, the Jamaican depositional sequences are interpreted to have formed as a result of a 1053 complex interplay of tectonics (collisional and extensional) and long- and short-term sea-level 1054 fluctuations. Major unconformities, particularly within the Cretaceous, lasting durations in excess of 1 1055 Ma are related to major tectonic episodes during the assembly of the crustal fragments that form 1056 Jamaica and the Nicaraguan Rise, while minor unconformities lasting durations less than 1 Ma are 1057 influenced by global eustasy. The relatively low rates of RSL rise and fall calculated in Jamaica and 1058 thick successions of carbonate sequences suggest that carbonate production rates were able to respond 1059 adequately to gradual changes in RSL during the Cretaceous to Paleogene until succumbing to 1060 drowning during a sharp increase in the rate of RSL rise and environmental deterioration in the Oligo- 1061 Miocene. 1062 1063 Acknowledgements

1064 The authors would like to thank CGG for funding this research and the Petroleum Corporation of 1065 Jamaica for providing samples and assistance during fieldwork and continued support throughout the 1066 study. We would also like to thank Prof. Simon Mitchell and the University of West Indies for 1067 assistance in the field. Thanks are also reserved for the constructive comments of three anonymous

Page 32 of 54 1068 reviewers and Tullow Oil Ltd., particularly Madeleine Slatford and Jim Hendry. Tullow Oil Ltd. are 1069 also thanked for permission to use the photomicrograph in Figure 7. Finally, we would like to credit 1070 CGG NPA Satellite Mapping for the creation of the geological map in Figure 2. 1071 1072 Funding

1073 This research was funded by CGG. 1074 1075 Declaration of interest: None

Page 33 of 54 1076 References

1077 Abbott Jr, R.N., Bandy, B.R. and Rajkumar, A., 2013. Cenozoic burial metamorphism in eastern 1078 Jamaica. Caribbean Journal of Earth Science, 46, pp.13-30. 1079 1080 Aguilar, M., Bernaus, J.M., Caus, E. and Hottinger, L., 2002. Lepidorbitoides minima Douvillé from 1081 Mexico, a foraminiferal index fossil for the Campanian. J. Foramin. Res., v. 32, no. 2, p.126-134. 1082 1083 Arden, D.D., Jr., 1969. Geologic history of the Nicaraguan Rise. Trans. Gulf Coast Assoc. Geol. Soc, 1084 v. 19, p. 295-309. 1085 1086 Arden, D.D., Jr., 1975. Geology of Jamaica and the Nicaragua Rise. In: Nairn, A.E.M., Stehli, F.G. 1087 and Brown, G.M. (eds.), The ocean basins and margins (Volume 3): the Gulf of Mexico and the 1088 Caribbean. Geophysical Journal of the Royal Astronomical Society, v. 48, no. 1, p.617-661. 1089 1090 Arnaud-Vanneau, A. 1994. Benthic foraminifer distribution and sequence stratigraphy. Examples from 1091 Lower Cretaceous carbonate platforms in France and Azerbaijan. Geologie Mediterraneenne, 21, 13- 1092 15. 1093 1094 Azeredo, A.C., Wright, V.P. and Ramalho, M.M., 2002. The Middle–Late Jurassic forced regression 1095 and disconformity in central Portugal: eustatic, tectonic and climatic effects on a carbonate ramp 1096 system. Sedimentology, 49(6), pp.1339-1370. 1097 1098 Baker, R.D., Hallock, P., Moses, E.F., Williams, D.A. and Ramirez, A., 2009. Larger foraminifers of 1099 the Florida reef tract, USA: Distribution patterns on reef-rubble habitats. J. Foramin. Res., v. 39, no. 4, 1100 p. 267-277. 1101 1102 Bandy O. 1..., 19$3. Ecology and paleoccology of sorne Cal ifornia Forarninifera. Part Il. Forarninile 1103 ral evidence of subsidence rates in the Ventura basin. J. Paleontol., 27, 200-203. 1104 1105 Banner, F.T. and Simmons, M.D., 1994. Calcareous algae and foraminifera as water-depth indicators: 1106 an example from the Early Cretaceous carbonates of northeast Arabia. Micropalaeontology and 1107 Hydrocarbon Exploration in the Middle East, p.243-252. 1108

Page 34 of 54 1109 Barker, R.W. and Grimsdale, T.F., 1936. A contribution to the phylogeny of the orbitoidal 1110 foraminifera, with descriptions of new forms from the Eocene of Mexico. Journal of Paleontology, 1111 p.231-247. 1112 1113 Bé, A.W.H., 1977 An ecological, zoogeographic and taxonomic review of Recent planktonic 1114 foraminifera. In: Ramsey, A.T.S (ed.) Oceanic micropalaeontology, p.1-100. Academic Press, London. 1115 1116 Beavington-Penney, S.J. and Racey, A., 2004. Ecology of extant nummulitids and other larger benthic 1117 foraminifera: applications in palaeoenvironmental analysis. Earth-Science Reviews, v. 67, p. 219-265. 1118 Benton, M.J. and Harper, D.A.T., 1997. Basic palaeontology. Harlow: Addison Wesley Longman. 1119 342pp. 1120 1121 Berggren, W.A. and Pearson, P.N., 2005. A revised (sub)tropical planktonic foraminiferal zonation of 1122 the Eocene and Oligocene. J. Foramin. Res., v. 35, p. 279-298. 1123 1124 Berggren, W.A., Kent, D.V., Swisher, C.C. III and Aubry, M-P., 1995. A revised Cenozoic 1125 geochronology and chronostratigraphy. In: Berggren, W.A., Kent, D.V., Aubry M-P and Hardenbol, J. 1126 (eds.), Geochronology, time scales and global stratigraphic correlation. SEPM (Soc. Sed. Geol.), Spec. 1127 Publ., no. 54, p. 129-212. 1128 1129 BouDagher-Fadel, M.K., 2008. Evolution and geological significance of larger benthic foraminifera. 1130 Developments in palaeontology and stratigraphy, v. 21, Elsevier, 539pp. 1131 1132 BouDagher-Fadel, M.K., 2015. Biostratigraphic and geological significance of planktonic foraminifera 1133 2nd edition. UCL Press, 306pp. 1134 1135 Brown, I. and Mitchell, S.F., 2010. Lithostratigraphy of the Cretaceous succession in the Benbow 1136 Inlier, Jamaica. Caribbean Journal of Earth Science, v. 41, no. 2, p. 25-37. 1137 1138 Brown, N.K. and Bronnimann, P., 1957. Some Upper Cretaceous rotaliids from the Caribbean region. 1139 Micropaleontology, p. 29-38. 1140 1141 Bunge, R., Yusri, Y., Miller, S., Ferro, E. and Bova, J., 2016. Seismic Carbonate Reservoir Prediction 1142 and Drilling Results, Offshore Nicaragua, Caribbean Sea. In AAPG/SEG International Conference & 1143 Exhibition. 1144

Page 35 of 54 1145 Burchette, T.P. and Wright, V.P., 1992. Carbonate ramp depositional systems. Sedimentary Geology, 1146 79(1-4), p.3-57. 1147 1148 Burke, K. 1988. Tectonic evolution of the Caribbean. Annual Reviews of Earth and Planetary 1149 Sciences, 16, 201–230. 1150 1151 Burke, K., Coates, A.G. and Robinson, E., 1969. Geology of the Benbow Inlier and 1152 surrounding areas, Jamaica. Proc. 4th Caribbean Geol. Conf., Trinidad, p. 299-307. 1153 1154 Burke, K., Fox, P.J. and Şengör, A.M.C., 1978. Buoyant ocean floor and the evolution of the 1155 Caribbean. Journal of Geophysical Research: Solid Earth, 83(B8), p.3949-3954. 1156 1157 Cabioch, G., Montaggioni, L.F., Faure, G. and Ribaud-Laurenti, A., 1999. Reef coralgal assemblages 1158 as recorders of paleobathymetry and sea level changes in the Indo-Pacific province. Quaternary 1159 Science Reviews, 18(14), p.1681-1695. 1160 1161 Calais, E., 1999. Plate boundary zone deformation in the northern Caribbean: observations, 1162 measurements and models. In: Penrose Conference: Subduction to Strike-Slip Transitions on Plate 1163 Boundaries, Jan. 18the24th. ABS, p. 13-16. 1164 1165 Carvajal-Arenas, L. C., Torrado, L. and Mann, P., 2015. Early and Late Eocene/Oligo-Miocene (!) 1166 petroleum system on the Nicaraguan Rise: Insights from basin and three-dimensional petroleum 1167 system modeling, in C. Bartolini and P. Mann, eds., Petroleum geology and potential of the Colombian 1168 Caribbean Margin: AAPG Memoir 108, p. 615–646. 1169 1170 Catuneanu, O., 2006. Principles of sequence stratigraphy. Elsevier. 1171 1172 Catuneanu, O., Galloway, W.E., Kendall, C.G.S.C., Miall, A.D., Posamentier, H.W., Strasser, A. and 1173 Tucker, M.E., 2011. Sequence stratigraphy: methodology and nomenclature. Newsletters on 1174 stratigraphy, 44(3), p.173-245. 1175 1176 Cochran, W.J., Spotila, J.A., Prince, P.S. and McAleer, R.J., 2017. Rapid exhumation of Cretaceous 1177 arc-rocks along the Blue Mountains restraining bend of the Enriquillo-Plantain Garden fault, Jamaica, 1178 using thermochronometry from multiple closure systems. Tectonophysics, 721, 292-309. 1179

Page 36 of 54 1180 DeMets, C. and Wiggins-Grandison, M., 2007. Deformation of Jamaica and motion of the Gonave 1181 microplate from GPS and seismic data. Geophysical Journal International, 168(1), p.362-378. 1182 1183 Di Lucia, M., Sayago, J., Frijia, G., Cotti, A., Sitta, A. and Mutti, M., 2017. Facies and seismic 1184 analysis of the Late Carboniferous–Early Permian Finnmark carbonate platform (southern Norwegian 1185 Barents Sea): An assessment of the carbonate factories and depositional geometries. Marine and 1186 Petroleum Geology, 79, pp.372-393. 1187 1188 Draper, G., 1987. A revised tectonic model for the evolution of Jamaica. In: Ahmad, R. (Ed.), 1189 Proceedings of a Workshop on the Status of Jamaican Geology, Special Issue (No. 10) 1987, 1190 Geological Society of Jamaica. Signart Printing House, Kingston, Jamaica, p. 24-51. 1191 1192 Draper, G., 1998. Geological and tectonic evolution of Jamaica. Contributions to Geology, UWI, 1193 Mona, 3, p.3-9. 1194 1195 Droxler, A., Cunningham, A., Hine, A.C., Hallock, P., Duncan, D., Rosencrantz, E., Buffler, R. and 1196 Robinson, E., 1993. Late middle Miocene Segmentation of an Eocene-early Miocene carbonate 1197 megabank on the Northern Nicaragua Rise tied to the tectonic activity at the North America/Caribbean 1198 plate boundary zone. AAPG Bulletin, v. 77. 1199 1200 Droxler, A., Morse, J.W., Glaser, K.S., Haddad, G.A. and Baker, P.A., 1991. Surface sediment 1201 carbonate mineralogy and water column chemistry: Nicaragua Rise versus the Bahamas. Marine 1202 geology, v. 100, no. 1-4, p.277-289. 1203 1204 Eberli, G.P. and Ginsburg, R.N., 1987. Segmentation and coalescence of Cenozoic carbonate 1205 platforms, northwestern Great Bahama Bank. Geology, 15, 75-79. 1206 1207 Eberli, G.P. and Ginsburg, R.N., 1989. Cenozoic progradation of northwestern Great Bahama Bank, a 1208 record of lateral platform growth and sea-level fluctuations. In: Controls on Platform and Basin 1209 Development (Eds P.D. Crevello, J.L. Wilson, J.F. Sarg and J.F. Read), Soc. econ. Paleont. Mineral., 1210 Spec. Publ., 44, 339-351. 1211 1212 Eva, A.N., 1976. The palaeoecology and sedimentology of Middle Eocene larger foraminifera in 1213 Jamaica. 1st. Int. Symp. on benthonic foraminifera of continental margins, Part B. Paleoecology and 1214 Biostratigraphy, Maritime Sediments, Spec. Pub., v. 1, p. 467-475. 1215

Page 37 of 54 1216 Eva, A.N., 1980. Pre-cyclical chamber arrangement in the foraminiferal genus Polylepidina Vaughan, 1217 1924. Micropaleontology, p.90-94. 1218 1219 Ewing , J. I., Antoine ,J.& Ewing , M. 1960. Geophys- ical measurements in the western Caribbean 1220 and Gulf of Mexico. Journal of Geophysical Research , 65 , 4087–4126. 1221 1222 Filipescu, S., and Gîrbacea, R., 1997. Lower Badenian sea level drop on the western border of the 1223 Transylvanian Basin: foraminiferal paleobathymetry and stratigraphy. Geol Carpath, 48(5), pp.325- 1224 334. 1225 1226 Fluegeman, R.H., 1998. Planktonic and benthic foraminifera from the Moore Town Shales (Paleocene) 1227 at the Fellowship Bridge section, Portland, Jamaica. In: Caribbean Geology: Into the 3rd Millennium: 1228 Transactions of the 15th Caribbean Geological Conference, 29th June-2nd July, p. 149-154. 1229 1230 Gabrovšek, F., 2009. On concepts and methods for the estimation of dissolutional denudation rates in 1231 karst areas. Geomorphology, 106(1-2), pp.9-14. 1232 1233 Gary, A., Sjogren, G., Jones, G. and Waters, V., 1996. The Integrated Paleontological System: A 1234 Software Application for the Analysis, Interpretation, and Integration of Biostratigraphic Data. The 1235 Paleontological Society Special Publications, 8, p.135-135. 1236 1237 Germeraad, J.H., Hopping, C.A., and Muller, J., 1968. Palynology of Tertiary sediments from tropical 1238 areas. Review of Palaeobotany and Palynology, 6: 189–348. 1239 1240 Gómez-Pérez, I., Fernández-Mendiola, P.A. and García-Mondéjar, J., 1998. Constructional dynamics 1241 for a Lower Cretaceous carbonate ramp (Gorbea Massif, north Iberia). Geological Society, London, 1242 Special Publications, 149(1), pp.229-252. 1243 1244 Gonzalez-Guzman, A.E., 1967. A palynological study on the upper Los Cuervos and Mirador 1245 formations (Lower and middle Eocene; Tibu area, Colombia). Brill, Leiden 129 p. 1246 1247 Gordon, M.B. & Muehlberger, W.R.: Evidence from the Valle de Catacamas supports a rightlateral, 1248 neotectonic sense of slip for the Guayape Fault of Honduras. Am. Assoc. Pet. Geol. Bull. 72 (1988), 1249 pp.190. 1250

Page 38 of 54 1251 Gradstein, F.M., Ogg, J.G., Schmitz M.D. and Ogg, G.M., 2012. The Geologic Time Scale 2012. 1252 Elsevier, Amsterdam, v.1-2, p. 1-1144. 1253 1254 Grippi, J., 1978. Geology of the Lucea Inlier, western Jamaica. PhD thesis 1255 1256 Grippi, J., 1980. Geology of the Lucea Inlier, Western Jamaica. The Journal of the Geological Society 1257 of Jamaica, 19, 1-24. 1258 1259 Grippi, J. & Burke, K., 1980. Submarine-canyon complex among Cretaceous island-arc sediments, 1260 western Jamaica. Geological Society of America Bulletin, 91, 179-184. 1261 1262 Gunn, J., 1981. Limestone solution rates and processes in the Waitomo district, New Zealand. Earth 1263 Surface Processes and Landforms, 6(5), pp.427-445. 1264 1265 Gunter, G.C., Robinson, E. and Mitchell, S.F., 2002. A new species of Omphalocyclus 1266 (Foraminiferida) from the Upper Cretaceous of Jamaica and its stratigraphical significance. Journal of 1267 Micropalaeontology, 21(2), pp.149-153. 1268 1269 Hallock, P. and Glenn, E.C., 1986. Larger foraminifera: A tool for paleoenvironmental analysis of 1270 Cenozoic carbonate depositional facies. Palaios, v. 1, p. 55-64. 1271 1272 Handford C.R. and Loucks, R.G., 1993. Carbonate depositional sequences and systems tracts– 1273 responses of carbonate platforms to relative sea-level changes. In: Loucks RG, Sarg JF (eds) 1274 Carbonate sequence stratigraphy, vol 57. American Association of Petroleum Geologists, Memoir, 1275 Tulsa, pp 3–41 1276 1277 Hanzawa, S., 1962. Upper Cretaceous and Tertiary three-layered larger foraminifera and their allied 1278 forms. Micropaleontology, pp.129-186. 1279 1280 Haq, B.U., Al-Qahtani, A.M., 2005. Phanerozoic cycles of sea-level change on the Arabian Platform. 1281 GeoArabia 10 (2), 127-160. 1282 1283 Haq, B.U., Hardenbol, J. and Vail, P.R., 1988. Mesozoic and Cenozoic chronostratigraphy and cycles 1284 of sea-level change. 1285

Page 39 of 54 1286 Hardenbol, J., Thierry, J., Farley, M.B., Jacquin, T., De Graciansky, P-C. and Vail, P.R., 1998. 1287 Mesozoic and Cenozoic sequence chronostratigraphic framework of European basins. In: de 1288 Graciansky, P-C. et al. (eds.), Mesozoic and Cenozoic sequence stratigraphy of European basins. 1289 SEPM Spec. Publ., no. 61, p. 109-127. 1290 1291 Hastie, A.R., Kerr, A.C., Mcdonald, I., Mitchell, S.F., Pearce, J.A., Wolstencroft, M. and Millar, I.L., 1292 2010a. Do Cenozoic analogues support a plate tectonic origin for Earth’s earliest continental crust? 1293 Geology, v. 38, p. 495–498. 1294 1295 Hastie, A.R., Kerr, A.C., McDonald, I., Mitchell, S.F., Pearce, J.A., Millar, I.L., Barfod, D. and Mark, 1296 D.F., 2010b. Geochronology, geochemistry and petrogenesis of rhyodacite lavas in eastern Jamaica: A 1297 new adakite subgroup analogous to early Archaean continental crust? Chemical Geology, v. 276, p. 1298 344-359. 1299 1300 Hastie, A.R., Mitchell, S.F., Kerr, A.C., Minifie, M.J., and Millar, I.L., 2011. Geochemistry of rare 1301 high-Nb basalt lavas: Are they derived from a mantle wedge metasomatised by slab melts? 1302 Geochimica et Cosmochimica Acta, v. 75, p. 5049-5072. 1303 1304 Heckel, P.H., 1986. Sea-level curve for Pennsylvanian eustatic marine transgressive-regressive 1305 depositional cycles along midcontinent outcrop belt, North America. Geology, 14(4), pp.330-334. 1306 1307 Hine, A.C., Locker, S.D., Tedesco, L.P., Mullins, H.T., Hallock, P., Belknap, D.F., Gonzales, J.L., 1308 Neumann, A.C. and Snyder, S.W., 1992. Megabreccia shedding from modern, low-relief carbonate 1309 platforms, Nicaraguan Rise. Geological Society of America Bulletin, 104(8), pp.928-943. 1310 1311 Hohenegger, J., 2005. Estimation of environmental paleogradient values based on presence/absence 1312 data: a case study using benthic foraminifera for paleodepth estimation. Palaeogeography, 1313 Palaeoclimatology, Palaeoecology, v. 217, p. 115-130. 1314 1315 Holcombe , T. L., Ladd , J. W., Westbrook , G., Edgar , N. T. & Bowland , C. L. 1990. Caribbean 1316 marine geology: ridges and basins of the plate interior. In : Dengo , G. & Case , J. E. (eds) The 1317 Caribbean Region . The Geology of North America, H , Geological Society of America, Boulder, CO, 1318 231–260. 1319 1320 Horowitz, D.H., 1976. Mathematical modeling of sediment accumulation in prograding deltaic 1321 systems. In Quantitative techniques for the analysis of sediments (pp. 105-119).

Page 40 of 54 1322 1323 Jackson, T.A., 1987. The petrology of Jamaican Cretaceous and Tertiary volcanic rocks and their 1324 tectonic significance. Proceedings of a workshop on the status of Jamaican geology: Geological 1325 Society of Jamaica, special issue, p. 107-119. 1326 1327 James, K. H., 2007. Structural geology: from local elements to regional synthesis. In: Bundschuh, J. 1328 and Alvarado, G. E. (Eds.) Central America: Geology, resources and hazards 1329 1330 James-Williamson, S.A., Mitchell, S.F. and Ramsook, R., 2014. Tectono-stratigraphic development of 1331 the Coastal Group of south-eastern Jamaica. Journal of South American Earth Sciences, 50, p.40-47. 1332 1333 Jaramillo, C., Rueda, M., Torres, V., 2011. A palynological zonation for the Cenozoic of the Llanos 1334 and Llanos Foothills of Colombia. Palynology, 35(1): 46–84. 1335 1336 Jarzen, D.M. and Dilcher, D.L., 2006. Middle Eocene terrestrial palynomorphs from the Dolime 1337 Minerals and Gulf Hammock Quarries, Florida, USA. Palynology, v. 30, no. 1, p. 89-110. 1338 1339 Jiang, M.,J. and Robinson, E., 1987. Calcareous nannofossils and larger foraminifera in Jamaican 1340 rocks of Cretaceous to early Eocene age. Status of Jamaican Geology, p. 24-51 1341 1342 Jones, B., M. Simmons, and J. Whittaker, 2004. "Chronostratigraphic and paleoenvironmental 1343 significance of agglutinated and associated larger benthonic foraminifera from the Lower to “Middle” 1344 Cretaceous of the Middle east. In: Bubik, M., and Kaminski, M.A. (Eds.) Proceeding of the Sixth 1345 International Workshop on Agglutinated Foraminifera, Grzybowski Found. Spec. Publ, vol. 8, pp. 1346 229-235. 1347 1348 Katz, M.E. and Miller, K.G., 1993. Miocene-Pliocene bathyal benthic foraminifera and the uplift of 1349 Buff Bay, Jamaica. Geological Society of America Memoirs, 182, p.219-254. 1350 1351 Kauffman, E.G. and Sohl, N.F., 1974. Structure and evolution of Antillean Cretaceous rudist 1352 frameworks. Verh. Nat. Forsch. Ges. Basel, 84/1: 339-467. 1353 1354 Kerr, A.C., White, R.V., Thompson, P.M. E., Tarney, J., Saunders, A.D., 2004. No oceanic plateau— 1355 no Caribbean plate? The seminal role of an oceanic plateau in Caribbean plate evolution. In: Bartolini, 1356 C., Buffler, R.T., Blickwede, J. (eds.). The Circum-Gulf of Mexico and the Caribbean: Hydrocarbon

Page 41 of 54 1357 habitats, basin formation, and plate tectonics. American Association of Petroleum Geologists Memoir 1358 79, 126–168. 1359 1360 Kirby, M.X. and Saul, L.R., 1995. The Tethyan bivalve Roudairia from the Upper Cretaceous of 1361 California. Palaeontology, v. 38, no. 1, p.23-38. 1362 1363 Kominz, M.A., Browning, J.V., Miller, K.G., Sugarman, P.J., Mizintseva, S. and Scotese, C.R., 2008. 1364 Late Cretaceous to Miocene sea‐level estimates from the New Jersey and Delaware coastal plain 1365 coreholes: An error analysis. Basin Research, v. 20, no. 2, p.211-226. 1366 1367 Krijnen, J.P., 1971. Analysis of pseudorbitoids from Glenbrook, Jamaica, and discussion of a study by 1368 Cole & Applin (1970). Rijksmuseum van Geologie en Mineralogie. 1369 1370 Krijnen, J.P., 1972. Morphology and Phylogeny of Pseudorbitoid Foraminifera from Jamaica and 1371 Curaçao: A Revisional Study. Rijksmuseum van Geologie en Mineralogie. 1372 1373 Krijnen, J.P., Macgillavry, H.J. and Van Dommelen, H., 1993. Review of Upper Cretaceous orbitoidal 1374 larger foraminifera from Jamaica, West Indies, and their connection with rudist assemblages. 1375 Geological Society of America Memoirs, v. 182, p.29-64. 1376 1377 Kumar, A. and Grambast-Fessard, N., 1984. Maastrichtian charophyte gyrogonites from Jamaica. 1378 Micropaleontology, v. 30, no. 3, p. 263-267. 1379 1380 Kumar, A. and Oliver, R., 1984. The occurrence and geological significance of charophyte 1381 gyrogonites from the Slippery Rock Formation (Maastrichtian), Central Inlier, Jamaica. Caribbean 1382 Journal of Science, v. 20, p. 29-34. 1383 1384 Leckie, R.M. and Olson, H.C., 2003. Foraminifera as proxies for sea-level change on siliciclastic 1385 margins In: Micropaleontologic Proxies for Sea-Level Change and Stratigraphic Discontinuities 1386 1387 Lewis, J.F., Kysar Mattietti, G., Perfit, M. and Kamenov, G., 2011. Geochemistry and petrology of 1388 three granitoid rock cores from the Nicaraguan Rise, Caribbean Sea: implications for its composition, 1389 structure and tectonic evolution. Geologica Acta: an international earth science journal, 9(3-4). 1390

Page 42 of 54 1391 Maharaj, R. and Mitchell, S.F., 2000. Sequence stratigraphy of mixed clastic-carbonate systems–a case 1392 example from the Eocene of Jamaica. Geological Society of Trinidad and Tobago SPE Convention 1393 and Exhibition technical presentation. 1394 1395 Mann, P., and K. Burke, 1984, Cenozoic rift development in the northern Caribbean: Geology, v. 12, 1396 p. 732–736. 1397 1398 Mann, P., DeMets, C., Wiggins-Grandison, M., 2007. Toward a better understanding of the Late 1399 Neogene strike-slip restraining bend in Jamaica: geodetic, geological, and seismic constraints. Geolog. 1400 Soc. Lond. Spl. Publ. 290, 239e253. http:// dx.doi.org/10.1144/SP290.8. 1401 1402 Mann, P., Draper, G. and Burke, K. 1985. Neotectonics of a strike-slip restraining bend system, 1403 Jamaica. In: K. Biddle and N. Christie-Blick (Eds.), Strike-slip deformation, basin formation, and 1404 sedimentation. SEPM Special Issue, 37, 211-226. 1405 1406 Mann, P., Rogers, R., and Gahagan, L. 2006. Overview of plate tectonic history and its unresolved 1407 tectonic problems. In: Bundschuh, J. & Alvarado, G. E. (eds) Central America: Geology, Resources 1408 and Hazards. Taylor & Francis/, Balkema, 201–237. 1409 1410 Mann, P., Taylor, F.W., Lawrence Edwards, R. and Ku, T-L., 1995. Actively evolving microplate 1411 formation by olblique collision and sideways motion along strike-slip faults: an example from the 1412 northeastern Caribbean plate margin. Tectonophys., v.246, p.1-69. 1413 1414 Martini, E., 1971. Standard Tertiary and Quaternary calcareous nannoplankton zonation. Proc. 2nd 1415 Planktonic Conf., Roma, 1969, v. 2, p. 739-785. 1416 1417 Matchette-Downes, C. and Mitchell, S.F., 2005. Jamaica's petroleum potential prompts a first licensing 1418 round. first break, 23(4). 1419 1420 McCann, W.R., 1999. Collisional tectonics of the Bahama Bank along the northern margin of the 1421 Caribbean plate: cause and effects. In: Penrose Conference: Subduction to Strike-slip Transitions on 1422 Plate Boundaries, Jan. 18th-24th. ABS, p. 70. 1423 1424 McFarlane, N., 1977. Geology of Jamaica, 1:250,000 scale. Mines & Geology Division, 1425 Ministry of Mining & Natural Resources. 1426

Page 43 of 54 1427 Mercedes-Martín, R., Salas, R. and Arenas, C., 2013. Facies heterogeneity and depositional models of 1428 a Ladinian (Middle Triassic) microbial-dominated carbonate ramp system (Catalan Coastal Ranges, 1429 NE Spain). Marine and Petroleum Geology, 46, pp.107-128. 1430 1431 Mercedes‐Martín, R., Salas, R. and Arenas, C., 2014. Microbial‐dominated carbonate platforms during 1432 the Ladinian rifting: sequence stratigraphy and evolution of accommodation in a fault‐controlled 1433 setting (Catalan Coastal Ranges, NE Spain). Basin Research, 26(2), pp.269-296. 1434 1435 Miller, K.G., M.A. Kominz, J.V. Browning, J.D. Wright, G.S. Mountain, M.E. Katz, P.J. Sugarman, 1436 B.S. Cramer, N. Christie- Blick, and S.F. Pekar. 2005. The Phanerozoic record of global sea-level 1437 change. Science 310:1,293–1,298. 1438 1439 Miller, K.G., Mountain, G.S., Wright, J.D., Browning, J.V., 2011. A 180-million-year record of sea 1440 level and ice volume variations from continental margin and deep-sea isotopic records. Oceanography 1441 24 (2), 40-53 1442 1443 Mitchell, S.F. and Blissett, D., 2001. Lithostratigraphy of the late Cretaceous to ?Paleocene succession 1444 in the western part of the Central Inlier of Jamaica. Caribbean Journal of Earth Science, 35, 19-31. 1445 1446 Mitchell, S.F. and Edwards, T.C.P., 2016. Geology of the Maastrichtian (Upper Cretaceous) 1447 succession of the Jerusalem Mountain Inlier in western Jamaica. Caribbean Journal of Earth Science, 1448 48, 29-36. 1449 1450 Mitchell, S.F. and Ramsook, R., 2009. Rudist bivalve assemblages from the Back Rio Grande 1451 Formation (Cretaceous: Campanian) of Jamaica and their stratigraphic significance. Cretaceous 1452 Research, v. 30, no. 2, p.307-321. 1453 1454 Mitchell, S.F., 2000. SS03 Facies analysis of a Cretaceous-Paleocene volcaniclastic braid-delta. GSTT 1455 2000 SPE conference proceedings, 1-9. 1456 1457 Mitchell, S.F., 2003. Sedimentary and tectonic evolution of central Jamaica. In: Bartolini, C., Buffler, 1458 R.T. and Blickwede, J.F. (eds.) The Circum-Gulf of Mexico and the Caribbean: hydrocarbon habitats, 1459 basin formation, and plate tectonics. AAPG Memoir, v. 79, Tulsa, Arizona, USA, p. 605-623. 1460

Page 44 of 54 1461 Mitchell, S.F., 2005. Biostratigraphy of Late Maastrichtian larger foraminifers in Jamaica and the 1462 importance of Chubbina as a Late Maastrichtian index fossil. Journal of Micropalaeontology, v. 24, 1463 no. 2, p.123-130. 1464 1465 Mitchell, S.F., 2006. Timing and implications of Late Cretaceous tectonic and sedimentary events in 1466 Jamaica. Geologica Acta: an international earth science journal, 4(1-2). 1467 1468 Mitchell, S.F., 2013. Stratigraphy of the white limestone of Jamaica. Bulletin de la Société Géologique 1469 de France, 184(1-2), pp.111-118. 1470 1471 Mitchell, S.F., 2016. Geology of the western margin of the Benbow Inlier-implications for the 1472 relationship between the Yellow Limestone and White Limestone groups (with the description of the 1473 Litchfield Formation, new name). Caribbean Journal of Earth Science, v. 48, p. 19-25. 1474 1475 Mitchell, S.F., Ramsook, R., Coutou, R. and Fisher, J., 2011. Lithostratigraphy and age of the St. 1476 Ann's Great River Inlier, northern Jamaica. Caribbean Journal of Earth Science, v. 42, p. 1-16. 1477 1478 Mitchum Jr, R.M., Vail, P.R. and Thompson III, S., 1977. Seismic stratigraphy and global changes of 1479 sea level: Part 2. The depositional sequence as a basic unit for stratigraphic analysis: Section 2. 1480 Application of seismic reflection configuration to stratigraphic interpretation. 1481 1482 Morley, R.J., 1991. Tertiary stratigraphic palynology in Southeast Asia: current status and new 1483 directions. Geological Society of Malaysia Bulletin, 28: 1–36. 1484 1485 Morley, R.J., 1996. Biostratigraphic characterization of systems tracts in Tertiary sedimentary basins. 1486 IPA Proceedings of the International Symposium on Sequence Stratigraphy in SE Asia, p. 49-71 1487 1488 Müller, R.D., Sdrolias, M., Gaina, C., Steinberger, B., Heine, C., 2008. Long-term sealevel 1489 fluctuations driven by ocean basin dynamics. Science 319, 1357-1362. 1490 1491 Murray, J.W., 1991. Ecology and palaeoecology of benthic foraminifera. Longman Group UK Ltd. 1492 397pp. 1493 1494 Murray, J.W., 2006. Ecology and applications of benthonic foraminifera. Cambridge University Press, 1495 426pp. 1496

Page 45 of 54 1497 Murris, R.J., 1981. Middle East--Stratigraphic Evolution and Oil Habitat. AAPG Bulletin, 65(7), 1498 pp.1358-1358. 1499 1500 Mutti, M., Droxler, A.W. and Cunningham, A.D., 2005. Evolution of the Northern Nicaragua Rise 1501 during the Oligocene–Miocene: drowning by environmental factors. Sedimentary Geology, v. 175, no. 1502 1, p. 237-258. 1503 1504 Naish, T. and Kamp, P.J., 1997. Foraminiferal depth palaeoecology of Late Pliocene shelf sequences 1505 and systems tracts, Wanganui Basin, New Zealand. Sedimentary Geology, 110(3-4), pp.237-255. 1506 1507 Olsson, R.K. and Nyong, E.E., 1984. A paleoslope model for Campanian-lower Maestrichtian 1508 foraminifera of New Jersey and Delaware. The Journal of Foraminiferal Research, 14(1), pp.50-68. 1509 1510 Olsson, R.K., Hemleben, C., Berggren, W.A. and Huber, B.T., 1999. Atlas of Paleocene Planktonic 1511 foraminifera. Smithsonian Institution Press. 1512 1513 Ott, B., Mann, P. and Saunders, M., 2013, December. Role of the offshore Pedro Banks left-lateral 1514 strike-slip fault zone in the plate tectonic evolution of the northern Caribbean. In AGU Fall Meeting 1515 Abstracts. 1516 1517 Pearson, P.N., Olsson, R.K., Huber, B.T., Hemleben, C. and Berggren, W.A., 2006. Atlas of Eocene 1518 planktonic foraminifera. Cushman Foundation, Spec. Publ., no. 41. 1519 1520 Peck, R.E. and Forester, R.M., 1979. The genus Platychara from the Western hemisphere. Review of 1521 Palaeobotany and Palynology, v. 28, no. 2, p.223-236. 1522 1523 Perch-Nielsen, K., 1979. Calcareous nannofossil zonation at the Cretaceous/Tertiary boundary in 1524 Denmark. In: Birkelund, T. and Bromley, R.G. (eds.), Symposium on the Cretaceous/Tertiary 1525 boundary events; I. The Maastrichtian and Danian of Denmark. Univ. Copenhagen, Denmark, p. 115- 1526 135. 1527 1528 Perch-Nielsen, K., 1985a. Cenozoic calcareous nannofossils. In: Bolli, H.M., Saunders, J.B. and 1529 Perch-Nielsen, K. (eds.), Plankton stratigraphy. Cambridge Univ. Press, p. 427-454. 1530 1531 Perch-Nielsen, K., 1985b. Mesozoic calcareous nannofossils. In: Bolli, H.M., Saunders, J.B. and 1532 Perch-Nielsen, K. (eds.), Plankton stratigraphy. Cambridge Earth Sci. Ser., p. 326-426.

Page 46 of 54 1533 1534 Petrizzo, M.R., Falzoni, F. and Silva, I.P., 2011. Identification of the base of the lower-to-middle 1535 Campanian Globotruncana ventricosa Zone: comments on reliability and global correlations. 1536 Cretaceous Research, 32(3), pp.387-405. 1537 1538 Pflum C. E., Frerichs W. E. , 1976. Gulf of Mexico deep-water foraminifers. edited by W. V. Sliter. 1539 Cushman Found. Foraminiferal Res. Spec. Publ.. 14,11-125. 1540 1541 Pindell, J. L. 1993. Regional synopsis of Gulf of Mexico and Caribbean evolution. In: Pindell, J. L. & 1542 Perkins, R. F. (eds) Mesozoic and Early Cenozoic Development of the Gulf of Mexico and Caribbean 1543 Region. Transactions of the 13th Annual GCSSEPM Research Conference, Houston, TX, 251–274. 1544 1545 Pindell, J. L., Cande, S., Pitman, W. C., Rowley, D. B., III, Dewey, J. F., LaBrecque, J. & Haxby, W. 1546 1988. A plate-kinematic framework for models of Caribbean evolution. Tectonophysics, 155, 121– 1547 138. 1548 1549 Pindell, J., Kennan, L., 2001. Kinematic evolution of the Gulf of Mexico and Caribbean. In: Fillon, 1550 R.H.,Weimer, P., Lowrie, A., Pettingill, H., Phair, R. L., Roberts, H.H., van Hoorn, B. (eds.). 1551 Petroleum Systems of Deep-Water Basins: Global and Gulf of Mexico Experience. Gulf Coast Section 1552 Society of Economic Paleontologists and Mineralogists. 21st Annual Bob Perkins Research 1553 Conference, Petroleum Systems of Deep-Water Basins, 193-220. 1554 1555 Pindell, J.L. and Kennan, L., 2009. Tectonic evolution of the Gulf of Mexico, Caribbean and northern 1556 South America in the mantle reference frame: an update. In: James, K.H., Lorente, M.A. and Pindell, 1557 J.L. (eds.), The origin and evolution of the Caribbean Plate. Geol. Soc. Lon. Spec. Publ., v. 328, p. 1- 1558 55. 1559 1560 Pindell, J.L., 1994. Evolution of the Gulf of Mexico and the Caribbean. In: Donovan, S.K. and 1561 Jackson, T.A. (eds.), Caribbean geology: an introduction. University of West Indies Publishers’ 1562 Association. Kingston, Jamaica, p. 13-39. 1563 1564 Pindell, J.L., Barrett, S.F., 1990. Geological evolution of the Caribbean region: a plate tectonic 1565 perspective. In: Dengo, G., Case, J.E. (eds.). The Caribbean. Decade of North American Geology, 1566 Boulder, Colorado, Geological Society of America, Volume H, 404-432 1567

Page 47 of 54 1568 Pindell, J.L., Kennan, L., Maresch, W.V., Stanek, K-P., Draper, G. And Higgs, R., 2005. Plate- 1569 kinematics and crustal dynamics of circum-Caribbean arccontinent interactions: Tectonic controls on 1570 basin development in Proto-Caribbean margins. Geol. Soc. Am. Spec. Pa. 394, p. 7-52. 1571 1572 Posamentier, H.W., James, D.P., Summerhayes, C.P., Haq, B.U. and Allen, G.P., 1993. An overview 1573 of sequence-stratigraphic concepts: uses and abuses. Sequence stratigraphy and facies associations, 1574 pp.3-18. 1575 1576 Poumot, C., 1989. Palynological evidence for eustatic events in the tropical Neogene. Bulletin des 1577 Centres de Recherches exploration-production Elf-Aquitaine, 13(2): 437–453. 1578 1579 Premoli-Silva, I. and Verga, D., 2004. Practical manual of Cretaceous planktonic foraminifera. 1580 Tipografia Pontefelcino, Perugia, p. 1-283. 1581 1582 Ramsook, R. and Robinson, E., 2009. A new record of the Paleocene larger foraminifer Ranikothalia 1583 catenula in the Wagwater Formation of eastern Jamaica. Caribbean Journal of Earth Science, v. 40, p. 1584 21-23. 1585 1586 Read, J.F., 1985. Carbonate platform facies models. AAPG bulletin, 69(1), pp.1-21. 1587 1588 Reiss, Z. & Hottinger, L. 1984. The Gulf of Aqaba - Ecological Micropaleontology. 354 pp. Springer- 1589 Verlag, New York. 1590 1591 Robaszynski, F. 1998. Planktonic foraminifera-Upper Cretaceous, Chart of Cretaceous 1592 Biochronostratigraphy. In P.-C. de Graciansky, J. Hardenbol and P.R. Vail (Eds.), Mesozoic and 1593 Cenozoic Sequence Stratigraphy of European Basins. Society for Sedimentary Geology (SEPM), 1594 Special Publication, v. 60, p. 782. 1595 1596 Robinson, E. and Mitchell, S.F., 1999. Upper Cretaceous to Oligocene stratigraphy in Jamaica. In: 1597 Mitchell, S. F. (ed.) Middle Eocene to Oligocene stratigraphy and palaeogeography in Jamaica: A 1598 window on the Nicaragua Rise, Contributions to Geology, UWI, Mona, no. 3, p. 1-47. 1599 1600 Robinson, E. and Wright, R.M., 1993. Jamaican Paleogene larger foraminifera. In: Wright, R. M. and 1601 Robinson, E. (eds.) Biostratigraphy of Jamaica. Geol. Soc. Am., v. 182, p. 283-345. 1602

Page 48 of 54 1603 Robinson, E., 2004. Zoning the White Limestone Group of Jamaica using larger foraminiferal genera: 1604 a review and proposal. Cainozoic Research, v. 3, no. 1-2, p. 39-75. 1605 1606 Rogers, R.D., Mann, P., Emmet, P.A. and Venable, M.E., 2007. Colon fold belt of Honduras: 1607 Evidence for Late Cretaceous collision between the continental Chortis block and intra-oceanic 1608 Caribbean arc. Special Papers, Geological Society of America, 428, p.129. 1609 1610 Sahagian, D., O. Pinous, A. Olferiev and V. Zakharov 1996. Eustatic curve for the Middle Jurassic – 1611 Cretaceous based on Russian Platform and Siberian stratigraphy zonal resolution. American 1612 Association of Petroleum Geologists Bulletin, v. 80, p. 1433-1458. 1613 1614 Sanchez, J., Mann, P. and Emmet, P.A., 2016. Late Cretaceous–Cenozoic tectonic transition from 1615 collision to transtension, Honduran Borderlands and Nicaraguan Rise, NW Caribbean Plate boundary. 1616 In: Nemčok, M., Rybár, S., Sinha, S. T., Hermeston, S. A. & Ledvényiová , L. (eds) Transform 1617 Margins: Development, Controls and Petroleum Systems. Geological Society, London, Special 1618 Publications, 431(1), pp.273-297. 1619 1620 Schlager W (1981) The paradox of drowned reefs and carbonate platforms. Geol Soc Am Bull 1621 92(4):197–211 1622 1623 Schmidt, W., 1988. Stratigraphy and depositional environment of the Lucea Inlier, western Jamaica. 1624 The Journal of the Geological Society of Jamaica, 24, 15-35. 1625 1626 Sharland, P.R., D.M. Casey, R.B. Davies, M.D. Simmons and O.E. Sutcliffe 2004. Arabian Plate 1627 Sequence Stratigraphy – revisions to SP2. GeoArabia, v. 9, no. 1, p. 199-214. 1628 1629 Sharland, P.R., R. Archer, D.M. Casey, R.B. Davies, S.H. Hall, A.P. Heward, A.D. Horbury and M.D. 1630 Simmons 2001. Arabian Plate Sequence Stratigraphy. GeoArabia Special Publication 2, Gulf 1631 PetroLink, Bahrain, 371 p. 1632 1633 Sigurdsson, H., Leckie, R. M. and Acton, G. D., 1997. Proceedings of the Ocean Drilling Program, 1634 Initial Reports, v. 165. 1635 1636 Simmons, M.D., Bidgood, M.D., Brenac, P., Crevello, P.D., Lambiase, J.J. and Morley, C.K. 1999. 1637 Microfossil assemblages as proxies for precise palaeoenvironmental determination – an example from 1638 Miocene sediments of northwest Borneo. in Jones, R.W., & Simmonds, M.D., (eds) Biostratigraphy in

Page 49 of 54 1639 Production and Development Geology. Geological Society of London Special Publication, no.152, 1640 219-241. 1641 1642 Simmons, M.D., Sharland, P.R., Casey, D.M., Davies, R.B. and Sutcliffe, O.E., 2007. Arabian Plate 1643 sequence stratigraphy: Potential implications for global chronostratigraphy. GEOARABIA, 12(4), 1644 p.101-130. 1645 1646 Simmons, M.D., Whittaker, J.E. and Jones, R.W., 2000. Orbitolinids from Cretaceous sediments of the 1647 Middle East–a revision of the FRS Henson and Associates Collection. In Proceedings of the 5th 1648 International Workshop on Agglutinated Foraminifera (Vol. 7, pp. 411-437). 1649 1650 Sissingh, W., 1977. Biostratigraphy of Cretaceous calcareous nannoplankton. Geol. Mijnbouw, v. 56, 1651 p. 37-65. 1652 1653 Sliter, W.V., 1989. Biostratigraphic zonation for Cretaceous planktonic foraminifers examined in thin 1654 section. J. Foramin. Res., v. 19, no. 1, p. 1-19. 1655 1656 Snedden, J.W., Liu, C., 2010. A compilation of Phanerozoic sea-level change, coastal onlaps and 1657 recommended sequence designations. AAPG Search Discov. Article 40594. 1658 1659 Steineck, P.L., 1974. Foraminiferal palaeoecology of the Montpelier and Lower Coastal Groups 1660 (Eocene-Miocene), Jamaica, West Indies. Palaeogeography, Palaeoclimatology, Palaeoecology, v. 16, 1661 p. 217-242. 1662 1663 Steineck, P. L., 1981, Upper Eocene to middle Miocene ostracode faunas and paleo-oceanography of 1664 the North Coastal Belt, Jamaica: Marine Micropaleontology, v. 6, p. 339-366. 1665 1666 Tipsword H. L., Setzer F. M., Smith F. L. , Jr., 1966. Interpretation of depositional environmcm in 1667 Gulf Coast petrorcum exploration from paleoecology and relaled stratigr.lphy. TraitS. Gulf CUO$I. A 1668 ssoc. Geol. Suc.. 16, 119-130. 1669 1670 Trechmann, C. T., 1923. The Yellow Limestone of Jamaica and its Mollusca. Geological Magazine, v. 1671 60, p. 337-367. 1672 1673 Tucker, M.E. and Wright, V.P., 1990. Carbonate sedimentology. John Wiley & Sons. 1674

Page 50 of 54 1675 Underwood, C.J. and Mitchell, S.F., 2004. Sharks, bony fishes and endodental borings from the 1676 Miocene Montpelier Formation (White Limestone Group) of Jamaica. Cainozoic Research, 3, pp.157- 1677 165. 1678 1679 Vail, P.R., 1987. Seismic stratigraphy interpretation using sequence stratigraphy: Part 1: Seismic 1680 stratigraphy interpretation procedure. 1681 1682 Van Wagoner, J.C., Mitchum Jr, R.M., Posamentier, H.W. and Vail, P.R., 1987. Seismic stratigraphy 1683 interpretation using sequence stratigraphy: Part 2: Key definitions of sequence stratigraphy. 1684 1685 Van Wagoner, J.C., Posamentier, H.W., Mitchum, R.M.J., Vail, P.R., Sarg, J.F., Loutit, T.S. and 1686 Hardenbol, J., 1988. An overview of the fundamentals of sequence stratigraphy and key definitions. 1687 1688 Wade, B.S., Pearson, P.N., Berggren, W.A. and Pälike, H., 2011. Review and revision of Cenozoic 1689 tropical planktonic foraminiferal biostratigraphy and calibration to the geomagnetic polarity and 1690 astronomical time scale. Earth Science Reviews, v. 55, p. 111-142. 1691 1692 Wadge, G., Jackson, T.A., Isaacs, M.C., Smith, T.E., 1982. The ophiolitic Bath-Dunrobin Formation, 1693 Jamaica: significance for the Cretaceous plate margin evolution in the north-western Caribbean. 1694 Journal of the Geological Society of London, 139, 321-333. 1695 1696 Wescott, W.A. and Ethridge, F.G., 1983. Eocene fan delta‐submarine fan deposition in the Wagwater 1697 Trough, east‐central Jamaica. Sedimentology, v. 30, no. 2, p.235-247. 1698 1699 Wilson, M.E., Bosence, D.W. and Limbong, A., 2000. Tertiary syntectonic carbonate platform 1700 development in Indonesia. Sedimentology, 47(2), pp.395-419. 1701 1702 Wissler L, Funk H, Weissert H (2003) Response of Early Cretaceous carbonate platforms to changes 1703 in atmospheric carbon dioxide levels. Palaeogeogr Palaeoclimatol Palaeoecol 200(1):187–205 1704 1705 Wright, R. M., 1971. Tertiary biostratigraphy of central Jamaica: tectonic and environmental 1706 implications. In: P. H. Mattson (editor), Trans. 5th Caribbean Geol. Conference (St. Thomas). Geol. 1707 Bull. 5, Queens College Press, Flushing, N.Y., p. 129 (abstract). 1708 1709 Wright, V.P. and Burchette, T.P., 1998. Carbonate ramps: an introduction. Geological Society, 1710 London, Special Publications, 149(1), pp.1-5.

Page 51 of 54 1711 1712 Zans, V.A., 1958. Geology of Jamaica map, 1958. 1:250,000 scale. Geol. Surv. Jamaica. 1713 1714 Zans, V.A., 1962. Synopsis of the geology of Jamaica: An explanation of the 1958 1715 provisional geological map of Jamaica (no. 4). Geol. Surv. Dept. 1716

Page 52 of 54 1717 Figure Captions

1718 Fig. 1 – Regional map of the western Caribbean showing bathymetric expression of the Nicaraguan 1719 Rise. NW-SE profiles across the Nicaraguan Rise show the gently inclined slope of the distally 1720 steepened carbonate ramp of the south-eastern margin of the Nicaraguan Rise. Image and profiles 1721 created using GeoMapApp software using Global Multi-Resolution Topography (NR – Nicaraguan 1722 Rise, J – Jamaica). 1723 1724 Fig. 2 – Geological map of Jamaica displaying locations of wells, coreholes and fieldwork localities. 1725 Geological map based on 1:250,000 scale Geological Survey of Jamaica map (Zans, 1958; 1726 1962),1:250,000 scale Mines & Geology Division, Ministry of Mining & Natural Resources, Jamaica 1727 map (McFarlane, 1977) and new observations from remote sensing of satellite data and structural 1728 observations recorded during fieldwork. 1729 1730 Fig. 3 – Ranges of key larger benthic foraminifera observed in this study. Ranges taken from 1731 references mentioned in the text and observations by this study. Stratigraphic column created using 1732 Timescale Creator (version 6.1.2, www.tscreator.org) following the timescale of Gradstein et al. 1733 (2012). 1734 1735 Fig. 4 – Jamaican RSL curve of the Early to ‘Middle’ Cretaceous (EKTR1 sequence). The highest sea- 1736 level is associated with the Al8 MFS with the regressive phase culminating with the formation of the 1737 Tu4 SB. A) The planktonic foraminifera Thalmanninella appenninica are observed within the top of 1738 the Rio Nuevo Formation in the Retrieve-1 well, B) Caprinid rudists and dasycladacean green algae 1739 occur within the Seafield Limestone of the Windsor-1 well. 1740 1741 Fig. 5 – Jamaican RSL curve of the Coniacian to Campanian (STR1 and CTR1 sequences). High sea- 1742 level correlates with the Co1 MFS in the STR1 sequence and Cam5 MFS in the CTR1 sequence. The 1743 planktonic foraminifera Globotruncana bulloides (A) and Globotruncana neotricarinata (B) occur 1744 within the Drax Hall Formation. The larger benthic foraminifera Pseudorbitoides trechmanni (C) 1745 displaying small number of primary chambers in juvenarium which comprises an initial spiral of 4-5 1746 peri-embryonic chambers (e.g. Krijnen, 1971; 1972) and Lepidorbitoides minima (D) within the 1747 Stapleton Limestone and Back Rio Grande Formation are characteristic of the middle to late 1748 Campanian. 1749 1750 Fig. 6 – Jamaican RSL curve of the Maastrichtian (MTR1-2 sequences). High sea-level of the MTR1 1751 sequence correlates with the Ma1 MFS. Two transgressive-regressive cycles occur within the

Page 53 of 54 1752 Maastrichtian. The larger benthic foraminifera Orbitoides megaloformis (A), Kathina jamaicensis (B) 1753 and Sulcoperculina angulata (C) are common in Maastrichtian strata from Jamaica. Immature calcite 1754 cemented sandstones of the Masemure Formation contain rare planktonic foraminifera including 1755 Heterohelix spp. (D). The miliolid foraminifer Chubbina jamaicensis (E) is an index fossil for the late 1756 Maastrichtian observed within the Titanosarcolites beds. Freshwater charophytes including Platychara 1757 spp. (F) occur within the Slippery Rock Formation indicating terrestrially influenced deposition. 1758 1759 Fig. 7 – Jamaican RSL curve of the Paleogene (PETR1, YTR1-3 and WTR1 sequences). Paleocene 1760 limestones of the Wagwater Formation contain rare Ranikothalia catenula (A). Grainstones within the 1761 YTR1 sequence contain common Eoconuloides lopeztrigoi (B), Fabularia colei (B), Pseudolepidina 1762 trimera (C) and Helicostegina gyralis (C) which displays long spire comprising 3 whorls (e.g. Barker 1763 and Grimsdale, 1936). The disappearance of the foraminifer Polylepidina chiapasensis (D), displaying 1764 pre-cyclical bispiral nepionic arrangement and 8 primary chambers which almost completely surround 1765 the proloculus (e.g. Hanzawa, 1962; Eva, 1970), is characteristic for the top of the Chapelton 1766 Formation and YTR2 sequence(Photograph of Polylepidina chiapasensis courtesy of Tullow Oil 1767 Ltd.). The YTR3 sequence is characterised by larger forms of Yaberinella jamaicensis (E). 1768 Lepidocyclinids increase in abundance and diversity through the upper part of the Yellow Limestone 1769 and into the White Limestone Group. Large Lepidocyclina (Eulepidina) undosa (F) characterise the 1770 Brown’s Town Formation which is often resedimented in calciturbidites of the Montpelier Formation. 1771 1772 Fig. 8 – The Jamaican RSL curve compared to global mean sea-level curves of Haq and Al-Qahtani 1773 (2005), Kominz et al. (2008) Muller et al. (2008), Snedden and Liu (2010) and Miller (2005; 2011). 1774 Major Jamaican unconformities are interpreted to be tectonic in origin, whereas shorter duration 1775 unconformities may correspond to global periods of sea-level fall. 1776

Page 54 of 54 Cuba

Hispaniola Yucatán rough J peninsula Cayman T

NR

Honduras A Nicaragua B

Hess escarpment

NW SE

2

(km) 4 Elevation 0 100 200 300 400 500 Distance (km) A: Vertical exaggeration = 10

NW SE

0

(km) 2 Elevation 0 100 200 300 400 500 Distance (km) B: Vertical exaggeration = 10

1 79°W 78°W 77°W 76°W

0 10 20 40 60 80 100

Kilometres

Sunderland Potosi-2 ¯ Blowfire Hill Sunderland Blackshop-1 Elderslie Windsor-1 Marchmont Cockpit-1 Windsor !! Jerusalem Mountain !! !! ! >!> ! ! ! >!>!! !! !! ! !!!! !! ! >! !! !!!! ! ! !! !! >!!! ! !!!>! ! !! ! !!>! ! West Negril-1 >! >! >! ! >! !! ! ! !>! !!! !! !! !!! !!! ! !! ! ! ! ! !!! ! >! ! ! !! !!!!!! ! ! Negril Spots-1 Hertford-1 !! ! !! Ecclesdown !!! !! !>! !!! !! >!

18°N 18°N Content-1 ! >! Retrieve-1 Santa Cruz-1 Pindars River-3 >! Pindars River-10 Portland Ridge-1

>! Arawak-1

17°N 17°N >! Pedro Bank Legend Pedro Bank-1 ! Fieldwork localities >! Wells >! Coreholes

2 79°W 78°W 77°W 76°W .) advena

Planktonic Standard Chronostratigraphy Foraminifers erseyella jamaicensis aberinella hottingeri aberinella jamaicensis Lepidocyclina (E.) undosa Ranikothalia catenula Rotalia skourensis Lepidorbitoides minima V Lepidocyclina (L.) ariana Orbitoides megaloformis Fabularia hanzawai Kathina jamaicensis Coleiconus christianaensis Miogypsinoides spp. (Americas) Polylepidina proteiformis Montcharmontia appeninica Fabularia colei Cushmania americana Eoconuloides lopeztrigoi Fissoelphidium operculiferum Y Discorinopsis gunteri Fabiania cassis Eulinderina subplana Heterostegina antillea Pseudorbitoides chubbi Helicostegina gyralis Pseudorbitoides trechmanni Eoconuloides wellsi Pseudophragmina (P Polylepidina chiapasensis Sulcoperculina angulata Nummulites fichteli Coskinolina douvillei Lepidocyclina (L.) yurnagunensis Fabularia vaughani/gunteri Lepidocyclina (L.) pustulosa Lepidocyclina (L.) canellei Y Fallotella cookei Chubbina jamaicensis Amphistegina parvula Lepidocyclina (L.) macdonaldi Neorotalia mexicana/metacapensis Ma Period Epoch Age/Stage Sub-Tropical Zone Pseudolepidina trimera

24 O7

25 Chattian 26 O6

27 O5 28 Oligocene O4 29

30 O3 Rupelian 31 O2 32

33 O1 34 E16 35 Priabonian E15 36 37 E14 38

39 Bartonian E13 40 E12 41 E11 42 E10 43 E9 44 Paleogene Lutetian 45 Eocene E8

46

47

48 E7

49

50 E6 51 E5 52 Ypresian

53 E4 54

55 E3 E2 56 E1 57 P5 Thanetian 58

59 P4

60 Selandian 61 Paleocene P3 62 P2 63 Danian P1 64

65 Pa P0 66 P. hantkeninoides 67 P. hariaensis 68 A. mayaroensis 69 Maastrichtian

70 R. fructicosa

71 P. palpebra 72 G. gansseri 73 G. aegyptiaca 74 Cretaceous Late

75 G. havanensis

76 R. calcarata

77

78 Campanian C. plummerae

79

80

81 G. elevata 82 3 83 Proximal Proximal Shallow-water Shelf Deep-water Distal shales sands carbonate sands carbonate shales

Standard Chronostratigraphy Planktonic 3rd Order Foraminifers Sequences Stratigraphy Relative sea-level (m) Sequences

Ma Period Epoch Age/Stage Sub-Tropical Zone Nanno 1 90 CC13 -10 10 20 30 40 50 60 70 80 90 Tu4 0 D. concavata 0 0+ 91 Turonian M. schneegansi CC12 92 H. helvetica 93 CC11 Tiber 94 W. archaeocretacea

CC10 Ce5 95 Late R. cushmani 96 T. reicheli 97 Cenomanian EKTR1 98 B T. globotruncanoides 99 CC9 100 Seafield 101 Cretaceous P. appenninica A 102

103 P. ticinensis

104

105 P. subticinensis 106 Early Albian Rio Nuevo 107 T. praeticinensis Al8 108 CC8

109 T. primula

110

111 T. madecassiana

112 M. rischi

Caprinid rudist

T. appenninica

100um A 1000um 4 B Proximal Proximal Shallow-water Shelf Deep-water Distal shales sands carbonate sands carbonate shales

Standard Chronostratigraphy Planktonic 3rd Order Foraminifers Sequences Stratigraphy Relative sea-level (m) Sequences

Ma Period Epoch Age/Stage Sub-Tropical Zone Nanno 1 -10 10 20 30 40 50 60 70 80 90

G. gansseri 0 0

73 0+ G. aegyptiaca 74 CC23 Cam9 G. havanensis 75 St Ann’s Back Rio Grande D 76 R. calcarata CC22 Lime Hall Blue 77 CC21 Mountain Campanian Stapleton Cascade 78 C. plummerae C CC20 Block CTR1 Newmans Hall A 79 Drax Hall Sunderland B 80 CC19 Liberty Hall Fm Cam5 Johns Hall 81 G. elevata CC18 Cam2 82 Sunderland & CC17 83 Marchmont Cretaceous Late Inliers 84 Sa3 Santonian 85 D. asymetrica CC16 Middlesex / 86 Windsor/ Dias Clamstead STR1 87 CC15 88 Coniacian D. concavata Co1 St Ann’s Inlier CC14 Lucea Inlier 89

G. neotricarinata

G. bulloides

200um 200um A B

P. trechmanni L. minima

200um C 500um D

5 Proximal Proximal Shallow-water Shelf Deep-water Distal shales sands carbonate sands carbonate shales

Standard Chronostratigraphy Stratigraphy Relative sea-level (m) Sequences Planktonic Foraminifers 3rd Order

Sequences 100+ -10

Sub-Tropical Nanno 90 60 70 10 30 20 40 80 50

Zone 0 Ma Period Epoch Age/Stage 66 P. hantkeninoides

Moravia 67 CC26 P. hariaensis Two Meetings Waterworks Yankee River Mahoe River Peckham 68 Green River Ma5 Ma5 Summerfield Grp. A. mayaroensis Titanosarcolites Guinea MTR2 69 Maastrichtian CC25 Beds E Corn Ma2 Veniella Thomas River/ Masemure Masemure Cretaceous Late R. fructicosa D Shales Slippery Rock F 70 B Jerusalem C Bath Belleisle Lst. 71 P. palpebra MTR1 CC24 Thicket Kensington Rio River Lst. A Grande Shepherds Mint Hall 72 Ma1Ma1 G. gansseri Jerusalem Marchmont & Central Blue Mountain Sunderland Inlier Mountain Campanian 73 Inlier Inliers Block G. aegyptiaca

74

O. megaloformis

O. megaloformis S. angulata K. jamaicensis

500um A 500um B 200um C

Heterohelix spp.

Platychara spp. C. jamaicensis 100um D 500um E 200um F

6 Proximal Proximal Shallow-water Shelf Deep-water Distal shales sands carbonate sands carbonate shales Plank. Foraminifers Standard Chronostratigraphy 3rd Order Sub-Tropical N,P Sequences Stratigraphy Relative sea-level (m) Sequences

Ma Period Epoch Age/Stage Zone Zones Nanno

1 -10 0 10 20 30 40 50 60 70 80 90 24 0 O7 0+ 25 NP25 Chattian P22 26 O6

27 O5 Montpelier F 28 Oligocene P21 NP24 O4 29 Ru3 30 O3 P20 Rupelian NP23 31 O2 P19 32 NP22 Somerset 33 O1 P18 NP21 WTR1 34 E16 White Limestone Grp. P16 / Claremont 35 P17 NP19 Swanswick Priabonian E15 36 -20 Troy 37 E14 P15 NP18 38 Pr1 Ipswich NP17 E YTR3 39 Bartonian E13 P14 Bart1 Preston Hill Marl 40 E12 P13 41 Healthy E11 Upper D Blue Mountain NP16 Hill YTR2 42 P12 Lu4 Chapelton Block E10 43 Lower Chapelton E9 P11 Lu3 Litchfield 44 Paleogene Lutetian P10 Foram. grst. / Palmetto Grove / C Eocene E8 NP15 Stettin 45 Freemans Hall B 46 ellow Limestone Grp. 47 YTR1 P9 Lu1 Y NP14 48 E7 Guys Hill New Ground 49

50 NP13 Yp9 E6 P8 51 Hanover & E5 P7 52 Ypresian NP12 Clarendon

53 Blocks Richmond E4 P6b 54 NP11 E3 55 P6a NP10 E2 56 E1 P5 NP9 57 P5 Thanetian NP8 58 Wagwater PETR1 NP7 A 59 P4 P4 NP6 60 Selandian NP5 61 Paleocene P3 P3 62 P2 P2 NP4 63 Da4 Moore Town Danian P1 64 P1 NP3 Pa 65 NP2 66 P0 Pa NP1 Da1

R. catenula

F. colei

P. trimera

H. gyralis

E. lopeztrigoi

500 um A 500 um B 1000 um C

Lepidocyclina spp.

L. (E.) undosa Y. jamaicensis

P. chiapasensis L. (E.) undosa

500 um D 1000 um E 1000 um F

7 Planktonic Standard Chronostratigraphy Foraminifers 3rd Order Jamaican Sequences Sequences Jamaican relative sea-level (m) Global mean sea-level curves (m)

Ma Period Epoch Age/Stage Sub-Tropical Zone Nanno

24 O7 0 0 20 90 10 20 30 80 100+ 40 60 70 40 50 60 25 80 -40 -20 100 120 -80 -60 160 -100 140 180 200 220 240 -10 280 300 320

NP25 260 Chattian 26 O6

27

O5 28 Oligocene NP24 O4 29 Ru3 O3 30 Rupelian NP23 31 O2 WTR1 32 NP22 33 O1 NP21 34 E16 35 E15 NP19- Priabonian 36 20

37 E14 NP18 38 Pr1 YTR3 39 NP17 Bartonian E13 Bart1 40 E12 41 E11 NP16 YTR2 42 Lu4 E10 43 E9 Lu3 44 Paleogene Lutetian NP15 45 Eocene E8

46 YTR1 47 Lu1 NP14

48 E7

49

50 NP13 Yp9 E6 51 Ypresian E5 52 NP12

53 E4 54 NP11 E3 55 NP10 E2 56 E1 NP9 57 P5 Thanetian 58 NP8 PETR1 NP7 59 P4 NP6 60 Selandian NP5 61 Paleocene P3 62 P2 NP4 63 Da4 Danian P1 64 NP3 Pa 65 NP2 P0 NP1 Da1 66 P. hantkeninoides 67 CC26 P. hariaensis 68 Ma5 A. mayaroensis 69 Maastrichtian CC25 MTR2 Ma2 70 R. fructicosa MTR1 71 P. palpebra CC24 72 G. gansseri Ma1 73 G. aegyptiaca 74 CC23 Cam9 G. havanensis 75

76 R. calcarata CC22 77 CC21 Campanian C. plummerae CTR1 78 CC20 79 CC19 80 Cam5 81 G. elevata CC18 Cam2 82 CC17 83 Late 84 Sa3 Santonian 85 D. asymetrica CC16 86 STR1

87 CC15 88 Coniacian Co1

89 D. concavata CC14

90 CC13 Tu4

91 Turonian M. schneegansi CC12 92 H. helvetica 93 CC11 94 W. archaeocretacea CC10 Ce5 95 R. cushmani 96 T. reicheli 97 Cenomanian 98 T. globotruncanoides 99 CC9 100 EKTR1 101 P. appenninica 102

103 P. ticinensis

104

105 Cretaceous P. subticinensis 106 Early Albian Al8 107 T. praeticinensis 108 CC8 Haq & Al-Qahtani (2005)

109 T. primula Kominz et al. (2008)

110 Muller et al. (2008)

111 T. madecassiana Snedden & Liu (2010)

112 Miller et al. (2005; 2011) M. rischi 8