You are on page 1of 22

Mesozoic Gulf of Mexico basin evolution from a planetary perspective

and petroleum system implications


Richard H. Fillon
Earth Studies Group, 3730 Rue Nichole, New Orleans, Louisiana 70131, USA (e-mail: fillorh@bellsouth.net)

ABSTRACT: The lack of a sharp boundary between North and South Atlantic stress
fields recorded in the arcs of transform faults and ridge segments suggests a gradual
merging of mantle stresses within a broad Central Atlantic plate boundary zone. The
nature of descending slabs deep in the mantle beneath North and South America
suggests that this intra-American plate boundary zone has existed since the Early
Cretaceous, at which time it was located beneath the Gulf of Mexico Basin. Simple
Euler sums of North and South America to Africa rotation poles validate the
concept of merging stress fields, providing a geologically reasonable trajectory and
rotation data for the Yucatan microplate with respect to Africa. The new Yucatan
rotation geometry is consistent with initiation of back-arc spreading in the western
Gulf of Mexico Basin during the Late Berriasian or Early Valanginian, c. 140 Ma,
triggered by a strengthening South Atlantic stress field. Continued spreading and
rotation of Yucatan likely persisted through the Late Albian, c. 100 Ma. These
findings are supported by Early Cretaceous deposystem architecture, basin margin
reef trends and source-rock distribution. Kinematic analysis predicts that most Gulf
of Mexico seafloor (c. 60%) was created during the Cretaceous period of stable
normal geomagnetic polarity, c. 125–83.5 Ma (the ‘Cretaceous Quiet Zone’). Salt-
lubricated detachment faulting in the young Gulf of Mexico likely covered newly
formed oceanic crust with large allochthons of Oxfordian–Valanginian strata.
KEYWORDS: Gulf of Mexico, plate tectonics, back-arc spreading, subduction, intra-American plate
boundary, Greater Antilles Arc, slab detachment, Early Cretaceous, deposystem architecture

INTRODUCTION consistent with planetary-scale geodynamic paradigms that are


The Gulf of Mexico Basin and associated areas have been the evolving largely due to evidence arising from investigations of
focus of many significant plate kinematic, plate tectonic and global seismic tomography (Grand et al. 1997; Dziewonski
related stratigraphic studies (e.g. Freeland & Dietz 1971; 2005). Discussion of points introduced herein should lead to a
Dickinson & Coney 1980; Walper 1980; White 1980; Pindell & better understanding of the Gulf Basin, including development
Dewey 1982; Anderson & Schmidt 1983; Duncan & Hargraves of one of the world’s most significant petroleum systems.
1984; Buffler & Sawyer 1985; Pindell 1985; Klitgord & Why seek a planetary geodynamic view for a small basin? In
Schouten 1986; Dunbar & Sawyer 1987; Michaud 1987; basic plate kinematic analyses all plates from super-continents
Salvador 1987; Blair 1988; Christenson 1990; Sedlock et al. to microplates are treated by the same rules. However, from a
1993; Marton & Buffler 1994; Pindell & Kennan 2001a, b; planetary geodynamic perspective, large plates and microplates
Jacques & Clegg 2002; Bird et al. 2005). In assessing the interact differently with the stresses created by planetary-scale
available geological and geophysical data, these studies have mantle dynamics (see summary in DeMets 1995). Microplate or
proposed or adopted the common postulate that the western plate-fragment systems, like the Japan–Japan Sea (Celaya &
Gulf of Mexico Basin opened by seafloor spreading and McCabe 1987) and Baja California–Gulf of California (Atwater
counterclockwise rotation of the Yucatan (Maya) block in the & Stock 1998) and, as argued in this paper, the Yucatan–Gulf
Late Jurassic–earliest Cretaceous (Bird et al. 2005) following a of Mexico system, are affected by the behaviour of neighbour-
>60 Ma long Late Triassic to Middle Jurassic interval of rifting. ing larger plates. Larger plates in such systems can move small
Such wide agreement on the principal chronologies and events continental fragments along complicated trajectories before
leading to formation of the Gulf of Mexico is surprising, given finally leaving them sutured onto continental portions of the
the healthy disagreements that have marked discussions of larger plates.
Caribbean Basin evolution (e.g. Meschede & Frisch 1998; Microplate rotation poles that honour basin geometry,
Muller et al. 1999; Pindell & Kennan 2001b; Kerr et al. 2003). basement geology and proposed chronologies can work
In this paper an analysis of Euler geometries of the large kinematically but they must also be consistent with fundamen-
plate systems surrounding the Gulf of Mexico Basin is pre- tal geodynamic processes driving the larger plates. The better-
sented which provides new insights into Gulf Basin evolution. constrained kinematic solutions of North and South Atlantic
A new kinematic model is proposed for the Gulf Basin that is large plate systems should be able to successfully predict the
Petroleum Geoscience, Vol. 13 2007, pp. 105–126 1354-0793/07/$15.00  2007 EAGE/Geological Society of London
106 R. H. Fillon

kinematics of the smaller Gulf of Mexico Basin system. This arc which was involved in the assembly of Mexican terranes
paper confirms the basis of that postulate and proposes that the along the western margin of Pangaea (Bartolini et al. 2001;
forces operating within the Gulf Basin are related to and Keppie 2004). At the younger end of their range the distri-
predictable from large-scale plate forces. bution of these rocks indicates considerable horizontal sub-
Several lines of evidence locate the Gulf of Mexico Basin in duction of Pacific crust (van Hunen et al. 2000). Andesites
an intraplate tectonic setting through much of the Mesozoic, interbedded with Albian–Turonian (c. 112–89 Ma) limestones,
including: (1) the history of circum-Gulf Basin volcanism (e.g. reported from a few industry wells located near the northwest-
Byerly 1991); (2) stress directions recorded in Atlantic fracture ern coast of the Yucatan Peninsula, formerly believed to be
zones (Muller et al. 1997; Smith & Sandwell 1997); (3) the associated with this group (Paine & Meyerhoff 1970; Lopez
nature of the critical North American–South American plate 1981; Byerly 1991), are now considered to be target melts
boundary (Muller et al. 1997); and (4) tomographic evidence of associated with the Chicxulub bolide event at the end of the
an ancient intra-American plate boundary beneath the Gulf Cretaceous (Alvarez et al. 1995).
Basin (Grand et al. 1997). Andesites of Late Aptian–Early Campanian age (c. 115–
Critically, this investigation reveals that new Yucatan Euler 81 Ma) are also widely reported from the Caribbean: in eastern
rotation poles derived entirely from published North and South Cuba (Kerr et al. 1999), Jamaica, Hispaniola, Puerto Rico, and
Atlantic pole data support adoption of an intra-American plate on the Aves Ridge (Almy 1969; Arden 1969; Fox et al. 1971;
boundary view of Gulf Basin tectonics. Derivation of the new Horsefield & Roobol 1974; Jackson & Smith 1979; Robinson
Yucatan poles is explained and a new kinematic model is 1994; Kerr et al. 1999). They are presumed to reflect subduction
presented for opening the Gulf of Mexico Basin, which of proto-Caribbean crust prior to creation and/or interjection
produces a young, i.e. largely Early Cretaceous, chronology of the Caribbean plate between the North and South American
for Yucatan rotation and seafloor spreading. The new Gulf plates in the Late Cretaceous (Pindell & Barrett 1990; Meschede
Basin chronology and geometry are tested by comparison & Frisch 1998; Muller et al. 1999; Acton et al. 2000). These
with chronologically calibrated deposystem extents drawn Aptian–Campanian volcanics may include remnant portions of
from industry and published sources in a series of six Early proto-Caribbean seafloor (Kerr et al. 2003). Although develop-
Jurassic–Early Cretaceous plate tectonic reconstructions. ment of an Early Cretaceous Greater Antilles arc-trench com-
plex is evident from the lithology and age data, the dip direction
of the subducting proto-Caribbean slab is open to differing
EVIDENCE FOR MESOZOIC INTRA-AMERICAN interpretations (Kerr et al. 1999; Pindell & Kennan 2001).
TECTONICS
Late Albian–Early Paleocene category Reports of Late Albian to
Circum-Gulf Basin volcanism Early Paleocene (c. 115–62 Ma) alkalic basalts, nephelenites and
Volcanism is a principal surface manifestation of planetary- phonolites in the northern Gulf Basin, including the Magnet
scale forces. It is therefore necessary to review the timing and Cove, Monroe Uplift and Jackson Dome localities onshore, and
significance of igneous activity in the greater Gulf Basin the Alderdice Bank and Door Point sites offshore (Braunstein
(Fig. 1). Regionally, Mesozoic igneous rocks cluster by age, type & McMichael 1976; Hunter & Davies 1979; Rezak & Tieh
and locale into three categories: (1) Rhaetian to Hettangian 1980; Byerly 1991; Baksi 1997), imply a deep-mantle source
(c. 204–197 Ma) dominantly tholeiitic basalts with some and probable mid-plate hotspot volcanism (McHone 2000;
andesite and rhyolite in eastern Mexico and the southeastern Anderson & Natland 2005). Volcanic rocks of similar types and
USA; (2) Tithonian to Early Campanian (c. 151–81 Ma) pillow ages are reported from as far west as central Texas (Byerly
lavas, andesites and bentonites in eastern Mexico and the 1991) and the Rio Grande Rift area of New Mexico, where they
Yucatan Peninsula; and (3) Late Albian to Early Paleocene predate Palaeogene rifting (Lasky 1938; McLemore et al. 2000a,
(c. 115–62 Ma), dominantly alkalic basalts with some quartz- b). Although the distribution of these rocks along an approxi-
poor nephelenite and phonolite lavas in the northern Gulf mate NW–SE axis is consistent with typical Atlantic hotspot
Basin (cf. Byerly 1991). tracks, their ages are not sufficiently constrained to confirm a
Rhaetian–Hettangian category Rhaetian–Hettangian (c. 204– propagation direction. The interpretation of a NW–SE positive
197 Ma) tholeiites are interpreted as rift-related basalts of the gravity anomaly in the Keathley Canyon area of the Gulf of
Central Atlantic Magmatic Province (Marzoli et al. 1999; Dalziel Mexico as a still undated hotspot track (Bird et al. 2005) is
et al. 2000; McHone 2000; McHone et al. 2005). They comprise consistent with these reports.
giant dykes and dyke-orientation swarms associated with the
initial fragmentation of Pangaea (Dalziel et al. 2000; McHone Spreading directions and the intra-American plate
et al. 2005). Byerly (1991) discussed uncertainties involved in boundary
dating that make some post-Hettangian tholeiitic basalt ages Nowhere is the intra-American (North American–South
equivocal. Reliable tholeiitic basalt ages in the circum-Gulf American) plate boundary defined clearly. The much smaller
Basin are now believed to group tightly around a Central Caribbean plate today occupies about two-thirds of the region
Atlantic regional average of 199.9 Ma (McHone 2000; Knight between the two larger plates, concealing much of the actual
et al. 2004), within the late Rhaetian using the time-scale of location and nature of the boundary between the continental
Gradstein et al. (2004). These rocks are closely similar in age and portions of the plates. To the east of the Windward Islands in
composition to African (Bertrand 1991; Knight et al. 2004) and the Central Atlantic, where the North and South American
South American (Marzoli et al. 1999) Central Atlantic Magmatic plates come into direct contact, the boundary between them is
Province (CAMP) basalts. indistinguishable. It is not marked by earthquake foci or by
Tithonian–Early Campanian category Tithonian–Early Campanian prominent tectonic boundary features, appearing rather to be a
(c. 151–81 Ma) pillow lavas, andesites and bentonites, sugges- zone of merging stresses, sensu Muller et al. (1999). This zone
tive of island-arc magmatism, are found in eastern Mexico also conceals an assumed North American–South American–
(Lopez 1981; Byerly 1991; Eguiluz de Antuñano 2001; African triple-junction which must lie somewhere along the
Goldhammer & Johnson 2001; Lawton et al. 2001), where they mid-ocean ridge (Escartin et al. 2003). If asthenospheric mantle
are attributed to island-arc volcanism, e.g. within the Chortis flow vectors immediately to the north and south of the
Gulf of Mexico back-arc tectonics 107

Fig. 1. Late Triassic to Late Cretaceous stratigraphy of the intra-American plate boundary region, including the Gulf of Mexico Basin, is
compared with hotspot and large igneous province (LIP) events, rift and ridge systems, geomagnetic polarity events, arc-trench systems and
related volcanism, and selected geotectonic reconstructions. MCU, Mid-Cretaceous Unconformity; MJU, Mid-Jurassic Unconformity. Numbered
superscripts in figure are keyed to additional comments and referenced sources. 1, Mexico (Byerly 1991); 2, Jamaica (Dickinson 1970; Arden
1969); 3, Jamaica, Cuba, Hispaniola (Pindell & Kennan 2001b) GCSSEPM; 4, Cuba (Iturralde-Vinent 1994, 1997; Garcia-Casco et al. 2003);
5, Hispaniola (Bowin 1966); 6, Puerto Rico (Kerr et al. 1999); 7, Triassic arcs in eastern Mexico – La Boca formation volcanic flows, andesite,
rhyolite, Eagle Mills equivalent redbeds in the Tampico area (Goldhammer & Johnson 2001); Zacatecas Formation turbidites (Monod & Calvet
1992; Bartolini et al. 2001); 8, Cretaceous arcs in Mexico – Monod & Calvet (1992) assign an early Cretaceous age to pillow lava and limestone
in the vicinity of Zacatecas city (c. 400 km SW of Monterrey), may be correlative with the Chilitos Formation (de Cserna 1976);
9, Triassic–Cretaceous arcs in Mexico – Chortis and Motagua terranes include andesite and other volcanic arc rocks of middle Jurassic–Upper
Cretaceous age, a metamorphic event c. 131–137 Ma ago in Juarez terrane, and a Maya terrane rotation 140 Ma ago (Keppie 2004); 10, paired
Sierra Leone Rise and Ceara Rise oceanic plateaus are dated to the Late Cretaceous (c. 73 Ma ago) by Kharin et al. (1977), Vogt & Jung (2005);
11, Florida volcanic province alkalic basalts (c. 140–180 Ma; Barnett 1975; Byerly 1991) are likely related to the location of the Sierra Leone
Rise–Ceara Rise hotspot in the Early Jurassic; 12, CAMP basalts date from c. 200 Ma (McHone et al. 2005); 13, evidence for Aves–Beata Ridge
hotspot or melting anomaly – Beata Ridge (Sinton et al. 2000) and northern Aves Ridge basalts (Fox et al. 1971) date c. 81 Ma; 14, Bermuda
Hotspot in the Cretaceous (Pollitz et al. 2001; Cox & Van Arsdale 2002); 15, South American and intra-American rifting (Van Siclen 1983; Lugo
& Mann 1995; Filho et al. 2000; McHone 2000); 16, Circum North Atlantic rifting in Late Triassic–Early Jurassic (Phair & Buffler 1983); rifting
and transition to North Atlantic spreading (Withjack et al. 1998); 17, South Atlantic spreading (Klitgord & Schouten 1986); 18, Caribbean plate
evolution (Pindell & Kennan 2001b); 19, development of Chortis arc (Pindell & Kennan 2001b) obduction and metamorphism in the Jurassic
(Horne et al. 1976) in response to clockwise rotation of the Chortis terrane and counterclockwise rotation of Mexican terranes (Keppie 2004),
the latter possibly induced by spreading at a Kulla–Farallon (or Phoenix) ridge; 20, Jurassic magnetic quiet interval (Gradstein et al. 2004);
21, Jackson Dome basalt ages 792.9 to 692.9 Ma (Saunders & Harrelson 1992); 22, Alderdice Bank alkalic basalt age 76.83 Ma (Rezak &
Tieh 1980); 23, Door Point basalt age 828 Ma (Braunstein & McMichael 1976); 24, COSUNA chart (Braunstein et al. 1985; Salvador 1985);
25, geotectonic reconstructions – Figures 6–11.

boundary differed significantly a more defined plate boundary region between c. 2 S and c. 17 N has persisted for much of
should be present. Accordingly, the mantle forces responsible the interval since the Late Cretaceous (Fig. 2a, b). This relatively
for significant differential motion between the North and South large area of the Central Atlantic containing the intra-American
American plates since the Late Cretaceous (Muller et al. 1999) plate boundary zone has, over time, accommodated varying
must then transition across a relatively broad plate boundary differences in North and South Atlantic spreading reflected in a
zone. complex geometry of oceanic fracture zones (Fig. 2a).
Is there direct evidence for a broad intra-American plate
boundary zone? Ridge-transform fault systems constrain the
extension of oceanic crust and, in the fracture zones they Tomographic evidence of an ancient plate boundary
produce, provide a cumulative record of crustal extension Adding the vertical dimension to kinematic analysis is made
(Garfunkel 1986; Muller et al. 1997; Smith & Sandwell 1997; possible largely by recent advances in global seismic tomogra-
Silver et al. 1998; Anderson 2001; Fairhead & Wilson 2005). By phy (Grand et al. 1997; Dziewonski 2005; Walker et al. 2005)
backtracking along fracture zones the drift trajectories of and derived mantle circulation models (Becker et al. 2003).
continental blocks and therefore the history of stresses imposed Global seismic tomography, in recording seismic velocity vari-
on the lower lithosphere by the flow of material in the ations within the Earth’s mantle (higher velocities=cooler,
asthenosphere can be deduced. Central Atlantic fracture zones denser material; lower velocities=warmer, less dense material),
record that a gradual merging of mantle stresses over a broad provides new insights into the configuration of descending
108 R. H. Fillon

Fig. 2. Fracture zones detected by satellite altimetry data (a, c) and seafloor magnetic anomalies (b) provide a record of stresses imposed on
oceanic crust by asthenospheric mantle flow (Becker et al. 2003). The altimetry-derived bathymetry maps (a. c) are from Smith & Sandwell (1997)
and the magnetic anomaly-derived age of seafloor map (b) is from Muller et al. (1997). Well-organized North and South Atlantic (NATL &
SATL) fracture zones (areas enclosed in polygons) clearly depict small circle arcs of rotation of North and South America away from Africa (a–c).
Instead of a clear boundary between North and South Atlantic fracture zone orientations and implied stress fields, however, the Central Atlantic
intra-American plate boundary zone contains complicated rotation arc geometry pointing to a gradual merging of mantle stresses. (a, c) reprinted
from Smith & Sandwell (1997) with permission from Science; (b) reprinted from Muller et al. (1997) with permission from the American
Geophysical Union.

crustal plates at scales ranging from hundreds to thousands of Jurassic–Late Cretaceous arc formation and subduction should
kilometres (e.g. Grand et al. 1997). Recent tomographic studies be consistent with tomographically resolved slab geometries.
and re-evaluations of kinematic analyses emphasize the impor- The agreement of Kulla–Farallon mantle circulation models
tance of examining plate kinematics as part of a system of with tomographic data, demonstrated by Bunge & Grand
planetary forces driven by temperature contrasts in the mantle, (2000), is consistent with a meridional north–south gradient in
especially in the intra-American plate boundary region (see flow directions that is expected to underlie an intra-American
discussion by Pindell & Kennan 2001b). plate boundary zone (Muller et al. 1997). Mantle circulation
Recently analysed and modelled tomographic data suggest models that are conservative with respect to slabs imaged in the
that during the Mesozoic Era the intra-American plate bound- lower mantle therefore support a dynamic linkage between the
ary zone may have extended beneath the Gulf Basin. Cool position and nature of intra-American and Kulla–Farallon plate
dense slabs reside in the mantle beneath present-day North and boundaries in the Cretaceous.
South America at depths between 900 km and 1500 km (Grand In Figure 3a and b most of the cool 1300–1500 km deep
et al. 1997; Bunge & Grand 2000; Pindell & Kennan 2001b). mass beneath North America is attributed to the portion of the
Comparing the location of these slabs to the reconstructed Kulla plate that was subducted beneath North America in the
positions of continents in the Cretaceous (Fig. 3) provides a Middle Jurassic to Early Cretaceous, while most of the 900 km
basis for identifying the sources of the descending material. deep mass beneath South America is attributed to the portion
Correlation of fast anomalies near the base of the mantle of the Farallon plate that was subducted beneath South
(c. 2600 km) to ancient subduction zones (Castle et al. 2000) America in the Early to Late Cretaceous. The locations of the
suggests that subducted slabs may remain as seismically distin- southern limit of the descending Kulla slab and the northern
guishable masses in the deep mantle for as long as 180 Ma. This limit of the descending Farallon slab and the inferred Kulla–
means that plate kinematic models proposing to explain Middle Farallon plate boundary, which otherwise are ill defined, are
Gulf of Mexico back-arc tectonics 109

Fig. 3. Seismic tomographic evidence for a deep mantle zone of merging North and South Atlantic stresses. Tomographic images of higher
velocity anomalies representing cool descending slabs at 1500 km and 900 km (blue) are adapted from Grand et al. (1997) and Bunge & Grand
(2000). The images, adapted from Grand et al. (1997), are compared with positions of the continents in the Cretaceous (c. 135–80 Ma).
Identification of the Kulla, Phoenix and Farallon slabs, the slab window, the intra-American boundary zone and the plate kinematic model are
explained in the text. Note that at 135 Ma (a, c) Yucatan and the Greater Antilles arc lie within the boundary zone; by 94 Ma (b, e) they are
moving out; and, by 80 Ma (f), they lie wholly to the north of the boundary zone above the Kulla slab. Reprinted from Grand et al. (1997) with
permission from the Geological Society of America.

indicated in the tomographic data (Engebretson et al. 1986; significant change in mantle flow direction south of the slab
Bunge & Grand 2000). window coincident with formation of the Farallon and South
A closer examination of the data presented by Grand et al. American plates and the intra-American plate boundary zone in
(1997) and Bunge & Grand (2000) in the area of the predicted the Early Cretaceous (c. 135 Ma).
Early Cretaceous intra-American plate boundary reveals a The likely plate boundary observed in the tomographic data
distinct break in the slab 1300 km beneath northwestern suggests that northern Farallon plate crust originally belonged
Colombia (Fig. 3c). The gap may record development of a slab to a once larger Kulla plate or to the Phoenix plate. This
window sensu Breitsprecher & Thorkelson (2001) or slab tear process of transferring crust from one plate to another and
sensu Pindell & Kennan (2001b) at the boundary between the implicitly from one asthenospheric flow regime to another is
Kulla and Farallon plates, or it could be evidence of a plate consistent with changes in fracture zone orientation observed
boundary between the Kulla and Phoenix plates sensu Sager et al. on Central Atlantic crust within the intra-American boundary
(1998) and Pockalny et al. (2002). In any event, this break zone. The Kulla–Farallon plate boundary, and the Kulla–
effectively isolates the southernmost portion of the slab at Phoenix plate boundary that preceded it in the same part of
1500 km from the much larger northern portion. At the 900 km the Pacific (Sager et al. 1998; Pockalny et al. 2002), like the
level, however, the southernmost portion of the slab is located intra-American plate boundary zone, lay within a region of
beneath northwestern Brazil, well to the east of the northern shifting stresses presumably related to the same large-scale
portion. It merges there with the descending Farallon slab meridional gradient in asthenospheric flow. In the Early
(Fig. 3b). Cretaceous that meridional gradient very likely encompasses the
Comparison of the slabs and the slab window, or tear, with Gulf of Mexico Basin (Fig. 3a).
the late Early Cretaceous locations of North and South Because quadruple plate junctions are inherently unstable, a
America (Fig. 3c–f) suggests that the southern isolated slab is link between the intra-American and Kulla–Farallon plate
evidence of subduction associated with what was most likely an boundaries is expected to be manifest by two triple junctions
oceanic arc-trench system located in the Pacific Ocean well in reasonably close proximity, comprising a conjugate pair.
seaward of the American plates. The abrupt eastward displace- They are constrained from moving too far apart by prevailing
ment of the southern slab at shallower depths may record a mantle flow directions, and from moving too close together
110 R. H. Fillon

Table 1. Euler pole and rotation angle data describing the opening of the North and South Atlantic Oceans and rotation of North and South America clockwise relative to Africa

Age (Ma) Euler Poles: NAM to NWAF from 0 Ma Euler Poles: SAM to Africa from 0 Ma
References Lat Lon Angle Published data Lat Lon Angle Ang Vel
245.0 66.95* 12.02* 75.55* Royer et al. (1992) 49.10 33.70 55.17
200.0 66.95* 12.02* 75.55* 49.10* 33.70* 55.17* 0.00
175.0 Klitgord & Schouten (1986) 66.95 12.02 75.55 49.10* 33.70* 55.17* 0.00
170.0 Klitgord & Schouten (1986) 67.02 13.17 72.10 49.10* 33.70* 55.17* 0.00
156.5 Klitgord & Schouten (1986) 67.15 16.00 64.70 49.10* 33.70* 55.17* 0.00
155.0 67.01* 16.45* 64.10* 49.10* 33.70* 55.17* 0.00
149.5 Klitgord & Schouten (1986) 66.50 18.10 61.92 49.10* 33.70* 55.17* 0.00
145.0 66.28* 18.27* 60.72* 49.10* 33.70* 55.17* 0.00
141.5 Klitgord & Schouten (1986) 66.10 18.40 59.79 49.10* 33.70* 55.17* 0.00
135.0 66.00* 18.47* 58.24* 49.10* 33.70* 55.17* 0.00
131.5 Klitgord & Schouten (1986) 65.95 18.50 57.40 Royer et al. (1992) 49.10 33.70 55.17 0.00
126.5 66.11* 18.95* 56.48* Royer et al. (1992) 49.30 33.80 54.29 0.18
126.0 Klitgord & Schouten (1986) 66.13 19.00 56.39 49.36* 33.84* 54.23* 0.12
125.0 66.15* 19.11* 56.12* 49.49* 33.91* 54.11* 0.12
121.0 66.24* 19.56* 55.05* Royer et al. (1992) 50.00 34.20 53.64 0.12
118.7 66.29* 19.82* 54.44* Royer et al. (1992) 50.10 34.60 52.78 0.37
118.0 Klitgord & Schouten (1986) 66.30 19.90 54.25 50.34* 34.59* 52.39* 0.56
115.0 67.20* 19.97* 52.08* 51.34* 34.54* 50.72* 0.56
109.5 68.86* 20.11* 48.09* 53.19* 34.44* 47.67* 0.56
106.0 69.92* 20.19* 45.55* 54.36* 34.38* 45.72* 0.56
105.0 70.22* 20.22* 44.83* 54.70* 34.37* 45.15* 0.57
102.0 71.12* 20.29* 42.65* 55.71* 34.31* 43.50* 0.55
94.0 73.54* 20.49* 36.85* 58.39* 34.17* 39.06* 0.56
92.0 74.14* 20.53* 35.40* 58.82* 34.11* 37.95* 0.55
84.0 Klitgord & Schouten (1986) 76.55 20.73 29.60 Cande et al. (1988) 61.75 34.00 33.50 0.56

Interpolated Euler parameters are indicated by an asterisk

by quadruple junction instability. Constraints operating on the two neighbouring large-plate systems (Table 1) should
paired triple junctions explain why once South America provide resultant pole solutions that are consistent with fracture
separates from Africa, the resulting intra-American plate zone and tomographic evidence. Table 1 lists Euler pole and
boundary develops not as a discrete feature but as a broad and rotation angle data describing the opening of the North and
persistent plate boundary zone within which the paired triple South Atlantic Oceans and rotation of North and South
junctions are free to move about. The initial development of America clockwise relative to Africa (from Klitgord &
this boundary zone likely coincides with and possibly predates Schouten 1986; Cande et al. 1988; Royer et al. 1992). The
the initial creation of the Kulla–Farallon spreading centre. resultant poles presented in Table 2 offer an attractive new
Paired triple junctions in the Central Atlantic might explain kinematic solution for the tectonic evolution of the Gulf basin.
differential motion of northern and southern Africa accommo- Kinematic reconstructions based on these poles should be
dated within the Benue Trough (Fairhead & Wilson 2005). consistent with merging mantle flows within the intra-American
Is there evidence of a small descending slab beneath the Gulf boundary zone.
of Mexico Basin? A mass of higher velocity material observed North and South Atlantic rotation arcs (small circles)
in tomographic images at 1300–1500 km is positioned within describe the relative motions of the North and South American
the Gulf Basin in the Early Cretaceous (Fig. 3d). Due to the plates in the Early Cretaceous (e.g. at 135 Ma, Fig. 4a, b).
northwestward drift of North America the movement of this Contrasting North and South Atlantic rotation arcs at 135 Ma
mass relative to the Gulf Basin in the Early Cretaceous is (Fig. 4c) indicate the typically large differential stress that must
consistent with a slab remnant of northwest-dipping subduc- be accommodated across the intra-American plate boundary
tion associated with a Greater Antilles arc and trench located to zone. Because incremental plate motions reflect evolving
the southeast of Yucatan (Fig. 3d). With northwestward sub- mantle flow trajectories and crustal responses, the small circles
duction implied, back-arc spreading becomes a possible mech- calculated from Table 1 (Figs. 5–11) provide a basis for
anism for opening the western Gulf of Mexico Basin and interpreting evolving fracture zone and cumulative spreading
rotating Yucatan away from North America along a reciprocal azimuths in plate kinematic reconstructions.
trajectory to prevailing mantle drift in the intra-American
boundary zone. If so, the trajectory of the Yucatan microplate
should be predictable from the combined directions of North PLANETARY VIEW OF GULF OF MEXICO
and South Atlantic spreading. RIFTING
Plates in motion
Poles supporting an intra-American plate boundary The Euler sums of North America to Africa (North Atlantic)
Individually, North and South American rotation poles and South America to Africa (South Atlantic) poles and angular
(describing motions with respect to Africa) do not explain the displacements (Table 2) describe the incremental counterclock-
Mesozoic rotation of Yucatan. If the average direction of wise rotation of Yucatan to its present alignment relative to
asthenospheric mantle flow in an intra-American plate bound- North America. In Table 2, parameters are calculated as Euler
ary zone is predictable from North and South Atlantic spread- sums of Yucatan vs. Africa motion using North and South
ing directions, however, simple Euler sums of plate motions in Atlantic pole data. The best-fit starting point for Yucatan is
Gulf of Mexico back-arc tectonics 111

Table 2. Euler pole and rotation angle data describing the opening of the western Gulf of Mexico Basin and rotation of the Yucatan microplate relative to Africa

Age (Ma) Euler Pole: YUC to AF (NATL) Euler Pole: YUC to AF (SATL) Net Pole: YUC to AF
Lat Lon Angle + Lat Lon Angle = Lat Lon Angle
245.0
200.0 66.95 12.02 70.00 + 49.10 33.70 8.00* = 69.68 10.99 62.59*
175.0 66.95 12.02 66.72 49.10 33.70 8.00*
170.0 67.02 13.17 65.78 49.10 33.70 8.00*
156.5 67.15 16.00 63.05 49.10 33.70 8.00*
155.0 67.01 16.45 64.10 + 49.10 33.70 8.00* = 69.88 16.06 56.64
149.5 66.50 18.10 61.92 49.10 33.70 8.00*
145.0 66.28 18.27 60.72 49.10 33.70 8.00*
141.5 66.10 18.40 59.79 49.10 33.70 8.00*
135.0 66.00 18.47 58.24 + 49.10 33.70 7.54* = 68.79 18.01 51.15
131.5 65.95 18.50 57.40 49.10 33.70 6.74*
126.5 66.11 18.95 56.48 49.30 33.80 5.60*
126.0 66.13 19.00 56.39 49.36 33.84 5.49*
125.0 66.15 19.11 56.12 + 49.49 33.91 5.26* = 68.06 18.71 51.15
121.0 66.24 19.56 55.05 50.00 34.20 4.34*
118.7 66.29 19.82 54.44 50.10 34.60 3.82*
118.0 66.30 19.90 54.25 50.34 34.59 3.66*
115.0 67.20 19.97 52.08 + 51.34 34.54 2.97* = 68.26 19.69 49.25
109.5 68.86 20.11 48.09 53.19 34.44 1.71*
106.0 69.92 20.19 45.55 54.36 34.38 0.91*
105.0 70.22 20.22 44.83 54.70 34.37 0.69*
102.0 71.12 20.29 42.65 55.71 34.31 0.00)*
94.0 73.54 20.49 36.85 + 58.39 34.17 0.00 = 73.54a 20.49a 36.85a
92.0 74.14 20.53 35.40 + 58.82 34.11 0.00 = 74.14a 20.53a 35.40a

Interpolated data are indicated by an asterisk.


a
Same as NAM to AF.

Fig. 4. (a) North Atlantic and (b) South Atlantic rotation poles, and small circles for 135 Ma (from Table 1) are shown superimposed on
magnetic anomaly-derived age of oceanic crust maps (Muller et al. 1997). The poles and arcs in Table 1 describe large plate kinematics of the
North and South American plates during the Early Cretaceous (e.g. at 135 Ma). The rotation arcs are generated using GMAP-2002 software
created by the Norwegian Geological Survey (Torsvik & Smethurst 1999). Also shown are (c) contrasts in the North and South Atlantic arcs at
135 Ma. The rectangle indicates the location of the 135 Ma intra-American plate boundary zone (c). (a, b) reprinted from Muller et al. (1997) with
permission from Elsevier.

taken as c. 5 (clockwise) from the initial position of North Opening of the western Gulf Basin rotates Yucatan counter-
America with respect to the North Atlantic pole at 200 Ma and clockwise with respect to both of the Atlantic Euler poles.
8 (clockwise) from the initial position of South America with A series of kinematic and geotectonic reconstructions drawn
respect to the oldest South Atlantic pole. Convergence of from Table 2 (Figs. 5–11) illustrate the Yucatan microplate’s
Yucatan’s North Atlantic rotation angle to match North Ameri- trajectory across the western Gulf Basin without significant
ca’s (Pangaean reassembly) occurs during the interval 155– overlap of neighbouring continental blocks at any stage in the
200 Ma. Convergence of Yucatan’s South Atlantic rotation angle rotation. A reasonable starting point positions Yucatan abutting
to zero, to match North America’s motion (suturing of Yucatan the Palaeozoic and Permo-Triassic (Ouachita front) margins of
onto North America), occurs during the interval 94–135 Ma. North America (Fig. 6a).
112 R. H. Fillon

Fig. 5. Symbols used in Late Triassic through Early Cretaceous plate


boundary reconstructions (Figs. 6–11) identify plates, rifts, ridges,
arc-trench systems, transform faults, triple junctions, hotspots and
hotspot tracks. In the geotectonic reconstructions transform faults
are drawn parallel to plate rotation arcs (small circles).

Synoptic plate reconstructions of the Gulf of Mexico Basin


and surrounding intra-American plate boundary zone represent
six critical instances in Late Triassic to Early Cretaceous crustal
evolution (Fig. 1). Kinematic snapshots include the end of the
Rhaetian (latest Triassic) c. 200 Ma (Fig. 6); the beginning of the
Kimmeridgian c. 155 Ma (Fig. 7); beginning of the Hauterivian
c. 135 Ma (Fig. 8); the beginning of the Aptian c. 125 Ma
(Fig. 9); the latest Aptian c. 115 Ma (Fig. 10); and the latest
Cenomanian c. 94 Ma (Fig. 11). Dates used in the Euler tables
and reconstructions are sensu Gradstein et al. 2004. The recon-
structions accommodate volcanic ages, as well as fracture zone
and tomographic evidence.

The Yucatan problem


There is broad agreement that the Yucatan block was originally
part of North America and that it subsequently rotated
counterclockwise to its present location, thus opening the
western Gulf of Mexico Basin. However, nine different Euler
rotation poles located within or very close to the Gulf Basin, in
an area bounded by the Florida Straits, northern Florida and the
western Bahamas Platform, have been proposed to explain the
kinematics of the Yucatan microplate (Bird et al. 2005). All of
these poles empirically describe the special case of Yucatan
Fig. 6. End of the Rhaetian (latest Triassic) geotectonic reconstruc-
motion with respect to North America (Pindell 1985, 1994; tion of the Gulf of Mexico Basin and intra-American boundary zone
Dunbar & Sawyer 1987; Christenson 1990; Hall & Najmuddin (c. 200 Ma). (a) Plate position, (b) plate boundary and (c) deposystem
1994; Marton & Buffler 1994). None of the poles cited architecture maps provide views of critical elements of Gulf Basin
addresses the likelihood that once separated from North evolution. Major plates, smaller crustal blocks and microplates in
America by a spreading centre, the Yucatan block is no longer Figures 6–11, 13 are identified by abbreviations: N, North America;
fixed to a North American reference frame. Failing to consider S, South America; AF and A, Africa; K, Kulla; F, Farallon; Px,
Phoenix; pC, proto-Caribbean; iaC, intra-american Caribbean; y,
adequately the strong possibility that Yucatan does leave the Yucatan; ch, Chortis; gm, Mexican Guerrero composite terrane; m,
North American reference frame introduces considerable mobile Mexican terranes south of the Mojave–Sonora Megashear;
uncertainty into comparing successive positions of the Yucatan b, Bahamas Platform; sf, South Florida. Plate boundary and other
microplate with crustal features such as the Mexican Oaxaquia, tectonic map symbols are explained in Figure 5. Deposystem
Coahuila and Sierra Madre terranes that remain fixed with architecture and extents are shown as isopach or accumulation rate
respect to post-Pangaean North America, sensu Keppie (2004). maps (northern Gulf Basin) drawn from industry and published well
data (Fillon et al. 2005) and as lithofacies maps (southern Gulf Basin)
This study diverges from earlier ones in that once seafloor adapted from Moran-Zenteno (1994). Limits and thickness (as
spreading begins in the western Gulf Basin, the Yucatan block, colour-filled isopachs) of Eagle Mills-equivalent strata range from
Maya terrane sensu Keppie (2004), is treated as a distinct <50 ft (blue) to >2500 ft (yellow). The known extents of age-
Yucatan microplate and its rotation is examined in both North equivalent La Boca marine sediments in Mexico are shown in yellow.
Atlantic and South Atlantic to Africa reference frames. Euler Atlantic hotspots in Figures 6–11, held stationary with respect to
poles meeting this criterion cluster tightly in the North Atlantic Africa, are identified as: CV, Cape Verde hotspot; SC, Sierra Leone
Rise–Cearra Rise hotspot; PP, Saints Peter and Paul hotspot; and
(c. 69 N; 20 W), falling close to but remaining distinct from AB, Aves Ridge–Beata Ridge hotspot (melting anomaly).
Gulf of Mexico back-arc tectonics 113

North America to Africa poles. These resultant Yucatan to creation of the South American and Farallon plates; (3) prior
Africa rotation poles (Table 2) incorporate North Atlantic and development of an intra-American plate boundary zone with
South Atlantic stress azimuths and a negative (counter- paired triple junctions along its Pacific margin; and (4) an
clockwise) rotation angle for Yucatan with respect to Africa, interval during which variable North and South Atlantic spread-
consistent with a back-arc spreading trajectory. ing directions combine to move Yucatan along a trajectory
Table 2 and Figures 6–11 present new kinematics for the that follows the East Mexican transform precisely yet avoids
western Gulf of Mexico Basin and young opening dates of continental overlap.
c. 135–c. 109–102 Ma (cf. Bird et al. 2005) that are consistent Asthenospheric mantle drag within an intra-American plate
with the physics of mantle circulation. A young western Gulf of boundary zone of interfering North and South Atlantic mantle
Mexico Basin chronology is constrained by the requirements of: flow regimes provides the driving mechanism for northwest-
(1) prior establishment of South Atlantic mantle flow; (2) prior ward subduction of proto-Caribbean crust beneath Yucatan
and, ultimately, for the back-arc spreading that separates the
Yucatan microplate from North America (Figs. 6b–11b). Back-
arc spreading, as incorporated in kinematic reconstructions
(Figs. 6b–11b), is consistent with the observed distribution of
Greater Antilles arc volcanism in time and space (Fig. 1) and
because it is the most efficient means of associating a required
negative rotation angle for Yucatan (Table 2) with generally
westward mantle flow beneath an intra-American boundary
zone (Zhong & Davies 1999; Huang et al. 2003).

Chortis, Mexico and the proto-Caribbean plate


As is the case for the Yucatan microplate, the proto-Caribbean
plate, Chortis and Mexican terranes are located within the
intra-American plate boundary zone during the Jurassic and
Cretaceous and, therefore, are subject to mixtures of mantle
stresses related to both North and South Atlantic spreading.
Although the proto-Caribbean plate, the Chortis block and the
Guerrero composite terrane of Mexico (sensu Keppie 2004) are
mobile during this period and may have interacted with the
southern and eastern margins of a drifting Yucatan microplate
propelled by back-arc spreading, they are not considered in
detail in this study. It is important, however, to provide an
indication of their likely motions during the Early Cretaceous.
The proto-Caribbean plate is likely created early in the Early
Cretaceous when the largely oceanic crust west of the intra-
American portion of the North Atlantic spreading centre is
isolated from North America by development of the Chortis
arc–Greater Antilles arc systems and a proto-Caribbean
transform (Fig. 8b).
Resultant Euler poles and rotation angles are calculated for
both Chortis and the proto-Caribbean by summing North and
South Atlantic motions (Table 3) in the same general manner
that Yucatan resultant poles are calculated (Table 2). In Table 3,
Euler pole and rotation angle data describe the movement of
largely oceanic crust in the intra-American Motagua and proto-
Caribbean seaways, and rotation of the Chortis block relative to

Fig. 7. (a–b) Early Late Jurassic, beginning of the Kimmeridgian,


geotectonic reconstruction of the Gulf of Mexico Basin and intra-
American boundary zone (c. 155 Ma). Symbols and abbreviations are
explained in Figures 5 and 6. The extent of inferred and constrained
Callovian Louann salt (lighter and darker pink) is indicated (b). Salt
overlapping western Yucatan is the allochthonous Neogene Sigsbee
Salt Canopy. For convenience, in the reconstructions the approxi-
mate southeastern limit of the Guerrero composite terrane is
referenced to the location of the city of Coatzacoalcos (Coatz) in
southeastern Veracruz state. (c) Accumulation rates of Haynesville–
Smackover syn-rift and rift-fill deposits (largely Kimmeridgian–
Oxfordian) are shown as colour-filled accumulation rate polygons in
the northern Gulf Basin. Stratal accumulation rates range from
<50 ft Ma1 (blue) to >2500 ft Ma1 (yellow). (c) Oxfordian litho-
facies in parts of Mexico are adapted from Moran-Zenteno (1994)
and Kimmeridgian–Oxfordian palaeoceanography is suggested by
green arrows. For reference, the modern outlines of Cuba and
Hispaniola are shown attached respectively to the North American
and proto-Caribbean plates.
114 R. H. Fillon

Fig. 8. (a–b) Beginning of the Hauterivian (Early Cretaceous) kinematic reconstruction of the Gulf of Mexico Basin–intra-American boundary
zone (c. 135 Ma). Symbols and abbreviations are explained in Figures 5 and 6. Ridge, transform fault and arc-trench plate boundaries develop
during the Late Berriasian–Early Valanginian (c. 140 Ma), isolating to varying degrees the proto-Caribbean plate, Chortis, the Yucatan microplate
and the Mexican Guerrero composite terrane from North America and Africa/South America. The Chortis terrane jumps the proto-Caribbean
ridge, being incorporated into the larger Late Berriasian–Early Valanginian proto-Caribbean plate. Back-arc spreading probably displaces the
Guerrero composite terrane southeastward relative to other Mexican terranes (Pindell & Kennan 2001b; Keppie 2004). Back-arc spreading in the
western Gulf Basin is initiated probably during the this period and likely continues throughout the Early Cretaceous. Salt-lubricated detachments
within Jurassic syn-rift section may cover newly formed oceanic crust with allochthons of Late Jurassic–Valanginian strata. In the eastern Gulf
Basin, Late Jurassic–Valanginian forearc compression buttresses the Florida margin. (c–d) Accumulation rates of expanded Upper Cotton Valley
(Berriasian Series) and Middle–Lower Cotton Valley (Tithonian Series) deposystems are shown as colour-filled accumulation rate polygons in the
northern Gulf Basin. Accumulation rates range from <50 ft Ma1 (blue) to >2500 ft Ma1 (yellow). (b–d) Knowles and equivalent units
represent the first extensive basin-margin reef system (cyan lines – (b–d)). Cotton Valley deposition marks the final stages of rift-fill deposition.
(c, d) Berriasian lithofacies in parts of Mexico are adapted from Moran-Zenteno (1994) and Tithonian–Valanginian palaeoceanography is
suggested by green arrows. slp, San Louis Potosi platform; tp, Tuxpan platform (d).

Africa, prior to the appearance of the Caribbean plate in the Convergence of Chortis’ North Atlantic rotation angle to
Late Cretaceous. The best-fit starting point for the Chortis match North America’s begins c. 200 Ma (Keppie 2004) and
block is taken as c. 25 (clockwise) from the initial position of continues throughout the Cenozoic. Convergence of Chortis’
North America with respect to the North Atlantic pole at South Atlantic rotation to match North America’s motion, does
200 Ma, and 10 (counterclockwise) from the initial position not begin until the appearance of the Caribbean plate in the
of South America with respect to the South Atlantic pole. Late Cretaceous, and continues throughout the Cenozoic.
Gulf of Mexico back-arc tectonics 115

Fig. 10. (a–b) Early Cretaceous, latest Aptian, geotectonic recon-


Fig. 9. (a–b) Early Cretaceous, earliest Aptian, geotectonic recon- struction of the Gulf of Mexico Basin–intra-American boundary
struction of the Gulf of Mexico Basin and intra-American boundary zone (c. 115 Ma). Symbols and abbreviations are explained in
zone (c. 125 Ma). Symbols and abbreviations are explained in Figures Figures 5 and 6. (c) Lower accumulation rates continue in the
5 and 6. Chortis collides strongly with the southern Mexican terranes James–Rodessa (Aptian) deposystem. A greatly expanded basin-
as the much smaller proto-Caribbean plate is consumed by the margin reef system (Stuart City and equivalents) rings the larger basin
Greater Antilles arc-trench system. (b) Ephemeral microplates (e.g. (b, c). (c) Aptian lithofacies in parts of Mexico are adapted from
X) form to the west of the proto-Caribbean ridge and quickly Moran-Zenteno (1994) and Late Aptian palaeoceanography is
subduct beneath the Yucatan microplate. (c) Generally lower accu- suggested by green arrows.
mulation rates in the Hosston–Sligo (Barremian) deposystem record
the beginning of an open-marine phase of deposition in a basin half
the size of the modern Gulf. A more continuous basin-margin reef Initial positions suggested for both Chortis and proto-
system (Sligo–Pettet and equivalents) rings the basin. Barremian Caribbean crust, and positive rotation angles relative to Africa
lithofacies in parts of Mexico are adapted from Moran-Zenteno (indicated in Table 3), are consistent with: (1) the reported
(1994). (c) Early Barremian palaeoceanography is suggested by green initial (Triassic) separation between Chortis and Mexican
arrows.
116 R. H. Fillon

terranes; (2) the development of Early Jurassic–Cretaceous


proto-Caribbean and Motagua oceanic crust between them;
(3) their convergence, obduction and juxtaposition in the Late
Jurassic–Early Cretaceous; (4) the northwestward subduction of
proto-Caribbean crust beneath the Yucatan microplate in the
Early Cretaceous; and (5) the development of a Yucatan–
Chortis transform in the Early Cretaceous (see Chortis and
Maya discussions in Keppie (2004) and proto-Caribbean arc
discussions in Kerr et al. (1999) and Pindell & Kennan (2001b)).
During the Latest Triassic–Early Jurassic fragmentation of
Pangaea, the Cortez and Guerrero composite terranes of
northern Mexico (m+gm in Fig. 6b) move generally southwest-
ward with respect to North America north of the Mojave–
Sonora megashear. Early Cretaceous rotation angles implied for
the Guerrero composite terrane, following Triassic–Jurassic
extension, are so small, however, that Euler poles are not
calculated.
To illustrate the motion of proto-Caribbean plate crust
relative to other intra-American crustal blocks, the modern
outlines of Hispaniola and Cuba are shown attached to the
proto-Caribbean and North American plates respectively
(Figs. 6–11). Extinction of the Greater Antilles arc and creation
of the proto-Lesser Antilles or Aves arc in the late Early
Cretaceous (Fig. 11) are consistent with published ages for arc
rocks in the Greater and Lesser Antilles (Fig. 1).

The Caribbean plate question


Formation of an intra-American Caribbean plate concurrently
with development of the Lesser Antilles arc in the late Early
Cretaceous or early Late Cretaceous, c. 115–94 Ma (Fig. 11) is a
possible kinematic solution. It does not exclude the possibility
of encroachment of an allochthonous Caribbean plate from the
Pacific at this time or later in the Late Cretaceous (Burke et al.
1984; Pindell & Barrett 1990; Pindell & Kennan 2001). Neither
does it support arguments for an intra-American origin for the
Caribbean plate that separates North and South America in the
Cenozoic (James 2003). The results of this study do indicate
that if the intra-American spreading centre that created the
proto-Caribbean plate remains active for a time follow-
ing creation of the Lesser Antilles arc-trench system and
the northern South American trench–transform boundary
(Fig. 11b), a new plate is created between the proto-Caribbean
ridge and South America. When this new plate is created the
old proto-Caribbean plate located to the west of the ridge may
already be extinct, sutured onto North America. The new plate,
however ephemeral, constitutes an intra-American Caribbean
plate that is clearly distinct from the proto-Caribbean plate
because it lies to the south of the proto-Caribbean ridge.

Yucatan–Greater Antilles subduction model


considerations
In proposing that the western Gulf of Mexico Basin opens
because of back-arc spreading related to a Chortis–Greater
Fig. 11. (a–b) Late Cretaceous, latest Cenomanian, geotectonic Antilles arc and trench system, the question of why so very little
reconstruction of the Gulf of Mexico Basin and intra-American plate evidence of arc-related rocks is found on the Yucatan Peninsula
boundary zone (c. 94 Ma). Symbols and abbreviations are explained
in Figures 5 and 6. (b) cr–pa arc, Costa Rica–Panama arc. (c) Low must be addressed. Observations consistent with arc develop-
accumulation rates continue in the Lower Eagle Ford–Buda–Lower ment include: (1) andesites of Jurassic–Late Cretaceous age
Tuscaloosa (Cenomanian) deposystem. (b–c) The basin-margin located within the Motagua and Chortis terranes sutured to the
reef system (Buda and equivalents) persists, but by the end of the southern and southeastern margins of the Yucatan block; (2) a
Early Cretaceous consists of small biohermal structures on a largely c. 131–137 Ma metamorphic event reported in Juarez terrane
relict bank system. (c) Cenomanian lithofacies in parts of Mexico abutting the southwestern margin of the Yucatan block;
are adapted from Moran-Zenteno (1994) and Late Cenomanian
palaeoceanography is suggested by green arrows. (3) reported Yucatan block (Maya terrane) rotation c. 140 Ma;
and (4) pillow lavas, bentonites and a possible volcanic arc of
similar age reported from southeastern Mexico in the Chiapas
Gulf of Mexico back-arc tectonics 117

Table 3. Euler pole and rotation angle data describing the movement of crust in the intra-American Motagua and proto-Caribbean seaways, including the Chortis block, relative to Africa

Age (Ma) Euler Pole: ch–hs to AF (NATL) Euler Pole: ch–hs to AF (SATL) New Pole: ch–hs to AF
Lat Long Angle + Lat Lon Angle = Lat Lon Angle
245.0 49.10 33.70
200.0 66.95 12.02 49.55* + 49.10 33.70 10.00* = 63.42 14.61 57.94
175.0 66.95 12.02 48.75* 49.10 33.70 10.00*
170.0 67.02 13.17 48.04* 49.10 33.70 10.00*
156.5 67.15 16.00 45.05* 49.10 33.70 10.00*
155.0 67.01 16.45 48.10* + 49.10 33.70 10.00* = 63.53 17.92 57.60
149.5 66.50 18.10 45.92* 49.10 33.70 10.00*
145.0 66.28 18.27 44.72* 49.10 33.70 10.00*
141.5 66.10 18.40 43.79* 49.10 33.70 10.00*
135.0 66.00 18.47 42.24* + 49.10 33.70 10.00* = 62.40 20.03 51.81
131.5 65.95 18.50 41.40* 49.10 33.70 10.00*
126.5 66.11 18.95 40.48* 49.30 33.80 10.00*
126.0 66.13 19.00 40.39* 49.36 33.84 10.00*
125.0 66.15 19.11 40.12* + 49.49 33.91 10.00* = 62.48 20.76 49.71
121.0 66.24 19.56 39.05* 50.00 34.20 10.00*
118.7 66.29 19.82 38.44* 50.10 34.60 10.00*
118.0 66.30 19.90 38.25* 50.34 34.59 10.00*
115.0 67.20 19.97 36.08* + 51.34 34.54 10.00* = 63.44 22.00 45.72
109.5 68.86 20.11 32.09* 53.19 34.44 10.00*
106.0 69.92 20.19 29.55* 54.36 34.38 10.00*
105.0 70.22 20.22 28.83* 54.70 34.37 10.00*
102.0 71.12 20.29 26.65* 55.71 34.31 10.00*
94.0 73.54 20.49 20.85* + 58.39 34.17 10.00* = 68.46 24.47 30.58
92.0 74.14 20.53 19.40* 58.82 34.11 10.00*

Interpolated data are indicated by an asterisk.

area (Horne et al. 1976; Carfantan 1977; Byerly 1991; Moran- Briefly summarized, Celaya & McCabe (1987) suggested that
Zenteno 1994; Keppie 2004). These observations fit the Japan rifts from Asia c. 70 Ma and rotates clockwise approxi-
Chortis–Greater Antilles arc geometry and chronology sug- mately 25 as a single rigid continental crustal block, opening
gested in this study (Figs. 8b–10b; 13a). However, ‘andesites’ the Japan Basin by c. 40 Ma. Subsequently, after an arc collides
reported within Cretaceous strata penetrated by several wells with Hokkaido c. 40 Ma, northeastern Japan begins an indepen-
drilled in the Chicxulub area of the northwestern Yucatan dent counterclockwise rotation of more than 60, while south-
Peninsula (Paine & Meyerhoff 1970; Byerly 1991) are now western Japan continues rotating clockwise. Back-arc spreading
considered to be target melts related to the terminal Cretaceous ceases in the late Miocene c. 11 Ma (Celaya & McCabe 1987;
bolide event (Blum et al. 1993; Alvarez et al. 1995). Hoshi & Matsubara 1998; Miller et al. 2006).
Might other evidence of arc magmatism have been removed While examining the evidence for Yucatan–western Gulf of
from the Yucatan block? Burke (1988) proposed that during Mexico back-arc spreading, it is also useful to consider the
spreading associated with Eocene development of the Yucatan possible effects of cessation of subduction on the tectono-
back-arc basin a portion of the Yucatan block was sheared stratigraphy of the region. Extinction of the Greater Antilles
away, transported northeastward, and amalgamated into west- volcanic arc late in the Early Cretaceous would very likely be
ern Cuba as the Guaniguanico terrane (see also Pszczolkowski associated with detachment of the descending slab, thus
1999; Pindell & Kennan 2001b; Bird et al. 2005). Cretaceous creating a slab window. The slab window would put hot
Guaniguanico strata containing serpentinite debris and coarse- asthenosphere in direct contact with the crust leading to crustal
grained detritus from the Cretaceous arc are reported by heating over a wide area, and to widespread thermal uplift (sensu
Iturralde-Vinent (1996) in olistostromes of the most southerly Parsons et al. 1994; Gao et al. 2004). Observations consistent
Quninones and Felicidade belts. with regional thermal uplift following arc extinction in the late
In calling upon intra-American back-arc spreading to explain Early Cretaceous Gulf of Mexico Basin include a significant
the western Gulf of Mexico Basin, the question of scale arises. broad regional decline in sediment accumulation rates (e.g.
Is it possible for back-arc spreading to be activated 600– pre-94 Ma vs. post-94 Ma section in Fig. 12) and development
1000 km distant from the proposed arc and subduction zone? of a broad regional erosional unconformity recorded as the
The answer is probably yes. The horizontal subduction model Middle Cretaceous Unconformity (MCU) or Middle Cretaceous
of van Hunen et al. (2000) provides a thermo-mechanical basis Sequence Boundary (MCSB) in Gulf of Mexico stratigra-
for developing such long offsets between subduction and phic studies (e.g. Buffler & Sawyer 1985; Buffler 1991; Feng
oceanic basalt generation (Fig. 13). Analogues with comparable & Buffler 1991; Feng 1995). In Figure 12 the regional uncon-
offsets include apparent horizontal subduction recorded formity is located between the Fredericksburg–Paluxy and
beneath Mexico (Ferrari et al. 1999; van Hunen et al. 2000) and Selma–Eutaw series.
proposed horizontal subduction explaining both the Rocky
Mountain and Andean uplifts (Schneider et al. 1988; Parsons MAPPED DEPOSYSTEMS AND THE KINEMATIC
et al. 1994). MODEL
The Japan Sea–Japanese main islands system may offer the
best and most familiar analog. Both the scale and complex Northern Gulf of Mexico deposystems
kinematic evolution of the Japan Sea are similar to those The northern Gulf of Mexico Basin is a biostratigraphically
proposed here for the Yucatan–western Gulf of Mexico system. data-rich area where over 200 000 industry stratigraphic well
118 R. H. Fillon

Fig. 12. Stratigraphically interpreted north–south seismic tracing (see inset 135 Ma geotectonic map for location). The line transects the Early
Cretaceous basin margin in the eastern Main Pass–Viosca Knoll lease areas. Stuart City, Sligo and Knowles-equivalent biohermal facies, indicated
by stylized shell fragments, are identified from stratal thickening and seismic facies characteristics. Stratigraphy and facies are tied to wells in the
Mobile and Destin Dome areas. Correlated horizon ages are posted at right. An erosional regional unconformity, the mid Cretaceous
unconformity (MCU) of Buffler & Sawyer (1985) separates Lower Cretaceous Fredericksburg–Paluxy strata from Upper Cretaceous
Selma–Eutaw strata. Adapted from Fillon (2001), reproduced with the permission of the GCSSEPM Foundation; further reproduction requires
their express written consent.

summaries containing over 5 000 000 events have been chrono- Southern Gulf of Mexico deposystems
stratigraphically evaluated and integrated with events from Onshore and offshore biostratigraphic data are not widely
published stratigraphic data (Fillon et al. 2005). Graphic corre- available from public or industry sources in the southern Gulf
lation analysis (sensu Shaw 1964) was used to refine a regional of Mexico Basin. Ocean drilling data are also much too limited
northern Gulf of Mexico Basin deposystem chronology tied to to provide a regional picture of deposystem architecture. Useful
the Gradstein et al. (2004) geological time-scale. Synoptic information on Middle Jurassic–Cretaceous sediment thickness
deposystem accumulation rate mapping and interpretation and depositional environments in eastern Mexico and Yucatan
based on this large chronostratigraphic database (Fillon et al. is contained, however, in summary lithofacies maps (Moran-
2005) places constraints on Gulf of Mexico Basin deposystem Zenteno 1994). These maps illustrate the distribution and
architecture throughout the Middle Jurassic–Cretaceous. extents of subaerial, nearshore, deep-water, carbonate, siliciclas-
Figures 6c–11c compare these synoptic isopach and accumula- tic and volcaniclastic lithofacies that can be compared with the
tion rate deposystem maps with the basin morphology basin morphology predicted in the kinematic reconstructions
suggested from kinematic reconstructions (Figs. 6b–11b). (Figs. 6c–11c).
Gulf of Mexico back-arc tectonics 119

Fig. 13. (a) A series of scaled northwest–southeast schematic cross-sections for the Gulf of Mexico Basin and intra-American plate boundary
zone extracted from plate kinematic reconstructions in Figures 6b–11b; (b) location. Cross-sections trace intra-American tectonics from initial
rifting, c. 200 Ma, through development of active microplate boundaries and back-arc spreading, c. 140 Ma, to the Gulf of Mexico Basin and
Yucatan sutured onto North America, c. 94 Ma. (c) A model of horizontal subduction from van Hunen et al. (2000) is compared with one of
the Yucatan–Gulf of Mexico Basin cross-sections at matching scales, suggesting that the long offsets between subduction and oceanic basalt
generation implied by a Greater Antilles arc–Yucatan microplate–Gulf Basin back-arc spreading system are consistent with thermo-mechanical
properties of the crust and mantle. NA, North America; SA, South America; Gom, Gulf of Mexico; Y, Yucatan; gaa, Greater Antilles Arc; pC,
proto-Caribbean. Reprinted from van Hunen et al. (2000) with permission from Elsevier.

Latest Triassic (200 Ma) how the calculated resultant poles (Table 2) restored the
The c. 200 Ma geotectonic reconstruction of the Gulf Basin Yucatan block to a position within the supercontinent that is
(Fig. 6a–c; 13) combines plate, plate boundary and deposystem consistent with both the Palaeozoic margin of North America
views of the Gulf Basin at the end of the Rhaetian (latest and the distribution of wells containing expanded Eagle
Triassic), just prior to Pangaean fragmentation. It illustrates Mills–La Boca-equivalent (Late Triassic) continental deposits
120 R. H. Fillon

(Fig. 6c). As schematically depicted, Northern Mexico, north the Florida Platform created as South Florida and the Bahamas
of the Mojave–Sonora megashear, includes mainly terranes Platform lagged the northwestward rotation of North America
sutured to Pangaea in the Palaeozoic (Keppie 2004). Southern in the Early Jurassic. In the southern Gulf Basin, the distri-
Mexico, as shown, consists mostly of the Cortez and Guerrero bution of lithofacies suggests an elevated Yucatan block
composite terranes (Keppie 2004). It is restored to its Pangaean rimmed by eroding highlands on the east and south. Asym-
position by rotation to the northwest along the Mojave–Sonora metric rifting may have caused disparate uplift of Yucatan or,
megashear (cf. Keppie 2004), which closely follows a North noting emergent areas reported along the southeastern and
Atlantic small circle arc of Klitgord & Schouten (1986). The southwestern Yucatan margins, perhaps an early response to
South Florida and Bahamas Platform blocks are also restored northwest-dipping subduction.
along these arcs (Pindell 1985). Seafloor spreading created During this period, the Motagua Ocean, located between
oceanic crust between southeastern Mexico and Chortis, southeastern Mexico and Chortis was nearly closed, its crust
opening the Motagua Ocean (Fig. 6b; Keppie 2004). largely subducted beneath the Chortis arc. At the same time
The modern outlines of Cuba, parts of which may have been Atlantic and intra-American spreading centres were actively
located near this restored position, and Hispaniola, adjusted generating new seafloor, allowing sea water to enter the Gulf of
for Cenozoic displacement, are provided for reference. The Mexico rift basin from the intra-American (proto-Caribbean)
reported extent of Triassic–Early Jurassic rifting is indicated basin that was developing to the east (Fig. 7c). Emergent areas
(rift symbols, Figs. 5, 6b), as are the probable positions of major and marine units in southeastern and eastern Mexico and
Atlantic hotspots assuming little or no motion with respect to accumulation of Smackover–Pimienta-equivalent marine source
Africa (hotspot symbols, Figs. 5, 6c). The Cape Verde, Sierra rocks also suggest that the nascent western Gulf of Mexico
Leone–Ceara Rise and Saints Peter and Paul hotspots (McHone Basin was tortuously connected to the Pacific Ocean by
2000; Vogt & Jung 2005) would all have been positioned within seaways across Mexico and through the Motagua Ocean.
the CAMP in the earliest Jurassic. Projected tracks of major Atlantic hotspots (Fig. 7c) cross
Figure 6c compares the kinematic reconstruction with the the US Atlantic Margin in the vicinity of Blake Plateau (Cape
mapped extent of Eagle Mills–La Boca-equivalent section and Verde hotspot) and the Gulf Basin margin in the vicinity of
well control (Fig. 6c inset). The small degree of overlap of southernmost Florida–western Cuba (Sierra Leone–Ceara Rise
Eagle Mills–La Boca-equivalent strata onto the restored north- hotspot). The Saints Peter and Paul hotspot remained station-
western Yucatan margin might be eliminated entirely by the full ary relative to South America in the area of the Demerara Rise
restoration of rifted North American and Yucatan continental and Guyana Plateau (McHone 2000).
crust to a pre-rift configuration. Gulf of Mexico Minerals
Management Service (MMS) deep-water lease areas are shown
for reference in the figure (Fig. 6c inset). Earliest Hauterivian (135 Ma)
The earliest Hauterivian geotectonic reconstruction, c. 135 Ma
(Fig. 1), provides views of the Gulf Basin (Figs. 8a–d, 13)
Earliest Kimmeridgian (155 Ma) corresponding to the end of Upper Cotton Valley–Lower
The geotectonic reconstruction of the Gulf Basin at the Casita siliciclastic deposystem development (Berriasian Series,
beginning of the Kimmeridgian, c. 155 Ma (Fig. 1), captures an Fig. 8c). Architecture of the Lower Cotton Valley siliciclastic
early stage in the movement of North America away from deposystem is also shown (Tithonian Series, Fig. 8d).
Africa (Figs. 7a–c; 13). By this time, 45 million years of rifting Several million years prior to the beginning of the Hauteriv-
had shifted the South Florida and Bahamas Platform crustal ian, possibly during the Late Berriasian or Early Valanginian,
blocks into their present alignment with North America c. 140 Ma, a new ridge system was initiated within the Jurassic
(Fig. 7a–c) and had also repositioned the Yucatan block a few rift basin that had developed between the Texas–Louisiana
hundred kilometres further away from the Palaeozoic margin of margin of North America and Yucatan (cf. Figs. 6b–8b).
North America and the Ouachita Front. Restoring the South Younger, latest Jurassic–earliest Cretaceous, rather than the
Florida and Bahamas Platform plate fragments in the manner Middle Jurassic initiation of seafloor spreading in the western
and time frame suggested by Pindell (1985), provides a result Gulf Basin (see summary in Bird et al. 2005), is dictated by the
that is consistent with both the fit of the continental plates chronology of an evolving mantle stress field in the South
within Pangaea in the Triassic and the reported extent of Atlantic (Tables 1, 2).
Callovian salt deposits throughout the Gulf Basin, Cuba, the The activation of a mantle-driven South Atlantic stress field
Bahamas, Blake Plateau and the northwest African margin (Fig. 3c) responsible for early rifting between South America
(Fig. 7b); The mapped interval closely postdates the Callovian and Africa and, ultimately, for the separation of South America
deposition of Louann–Minas Viejas–Ismithian–Punta Alegre- from Africa, triggered: (1) creation of a Greater Antilles
equivalent salt throughout the intra-American boundary region arc-trench system; (2) development of a distinct proto-
(Paine & Meyerhoff 1970; Van Houten 1977; Watkins et al. Caribbean plate isolated from the Yucatan microplate by the
1978; Dillon et al. 1979; Klitgord & Schouten 1986; Salvador Greater Antilles trench and from North America by the
1987; Taylor et al. 2000; Magoon et al. 2001; Meneses-Rocha proto-Caribbean transform; (3) northwestward subduction of
2001; Pindell & Kennan 2001a, b; Moretti et al. 2003; Bird et al. proto-Caribbean crust beneath a newly isolated Yucatan micro-
2005). plate; and (4) back-arc spreading in the western Gulf Basin
The restored position of Yucatan at the end of the beginning (Fig. 8b).
of the Kimmeridgian is consistent with accumulation rates of It is suggested that the subduction polarity of the Chortis arc
syn-rift, rift-fill and platform sediments of Kimmeridgian– flipped at this time and that associated back-arc spreading in
Oxfordian-aged Haynesville–Smackover and Olvido–Zuloaga- Mexico drove the Guerrero composite terranes into Chortis,
equivalent strata (Fig. 7c; Moran-Zenteno 1994). In the obducting the remnants of Motagua seafloor, as suggested by
northern Gulf Basin mapped deposystems reflect deposition regional geotectonic reconstructions, space considerations and
centred in a Red Sea- or Gulf of California-like rifted basin that the mid-ocean ridge basalt (MORB) geochemistry of Motagua
developed between South Texas and northwestern Yucatan and ophiolites (Kerr et al. 1999; Pindell & Kennan 2001b; Dickinson
in a complex extensional terrane between southeast Texas and & Lawton 2001; Keppie 2004). The Chortis microplate jumped
Gulf of Mexico back-arc tectonics 121

the newly established proto-Caribbean spreading centre (Faja de Oro) microblock was separated fully from the Yucatan
between Mexico and northwestern South America, becoming microplate (Fig. 9b, c).
incorporated into the proto-Caribbean plate. Whether the With westward migration of the proto-Caribbean ridge,
changes in the Chortis arc triggered development of the subduction of proto-Caribbean plate seafloor beneath the
Greater Antilles arc and back-arc spreading in the Gulf Basin, Yucatan microplate potentially isolated proto-Caribbean
or the latter caused the Chortis arc to reverse polarity cannot be oceanic plate fragments, creating new microplates and transient
determined. Either explanation is consistent with the data and triple junctions west of the proto-Caribbean ridge, e.g. an
the model. expanded Chortis microplate and ephemeral microplates
During the 20 million years following the end of the (Fig. 9b). To the east of the proto-Caribbean ridge, early
Oxfordian, c. 156 Ma, active North Atlantic and proto- separation of South America from Africa in response to
Caribbean spreading centres created an open, c. 400 km wide, accelerated rifting and the growing influence of a South Atlantic
Valanginian Gulf Basin connecting the North Atlantic basin to stress field likely placed northwestern South America in
the eastern Pacific (Fig. 8c). At this time the western Gulf was northwest-southeast compression.
wide enough to allow development of the first extensive By 125 Ma the western Gulf of Mexico Basin had widened
continental margin reefs (Knowles–San Juan equivalents considerably to over 600 km, encouraging expansion of Sligo–
Figs. 8c, 12). Reefal architecture on the Tuxpan and San Louis Pettet–Cupido–La Virgen-equivalent continental margin reef
Potosi platforms suggests a trail of microblocks left behind and platform carbonate systems (Figs. 9b, c, 12). Accumulation
as the Yucatan microplate continued moving to the southeast. rates and lithofacies extents of Hauterivian–Barremian
The absence of an extensive marginal reef system before Hosston–Upper Casita-equivalent marine siliciclastics and car-
c. 135 Ma is significant because it suggests that prior to that bonates are consistent with the reconstructed shape of the
time the Gulf Basin was too narrow and restricted in its western Gulf Basin (Fig. 9c). Principal Barremian deposystems
connection to the world ocean to support shelf margin were distributed along the shelf-margin, although possible large
bioherms. It is likely that the development of extensive bypass slope-fan systems, or more likely massive salt-lubricated
basin margin bioherms and the initiation of dynamic back-arc allochthons associated with detachment faulting were located
spreading are linked. offshore South Louisiana and Texas (Fig. 9c; Fillon et al. 2006).
The restored western Gulf Basin is consistent with mapped The distribution of nearshore volcanic units and emergent
accumulation rates and lithofacies of Tithonian–Valanginian areas along the southern Yucatan margin is consistent with an
age Cotton Valley–Bossier–Knowles–Lower Casita–San active Greater Antilles arc, continued subduction beneath
Juan-equivalent strata (Fig. 8c, d). Principal Tithonian deposys- Yucatan, and an active transform or remnant arc in northwest
tems are centred in East Texas (Fig. 8c), while Berriasian– Chortis. The Mexican, Panamanian–Costa Rican and Central
Valanginian deposition (Fig. 8d) is concentrated further south. Atlantic seaways remained open connecting the intra-American
Seaward of the Gulf Basin margins, Lower Cretaceous source basins to the Pacific and North Atlantic Oceans.
rocks (sensu Cole et al. 2001) accumulated during this period.
Large salt-lubricated detachments were probably active along
the basin rim during this phase (Fillon et al. 2006). As in the Latest Aptian (115 Ma)
Gulf of California, such large-scale detachments have the The c. 115 Ma, latest Aptian geotectonic reconstruction (Figs. 1,
capacity to cover newly forming oceanic crust very quickly with 10a–c, 13) positions the Yucatan microplate, still tightly abut-
thick sedimentary allochthons (Gastil & Fenby 1991). ting the East Mexican Transform, within 200 km of its modern
In the Early Hauterivian, the Mexican seaway remained open alignment with North America (Fig. 10a–c).
to the Pacific and a Panamanian–Costa Rican seaway, located to Back-arc spreading persisted in the western Gulf of Mexico
the east of Chortis, opened as the Motagua Ocean closed, thus until the Middle to Late Albian (c. 109–102 Ma – Table 2) as
maintaining the connection between the intra-American basins northwest-dipping subduction consumed additional proto-
and the Pacific. A deep Central Atlantic seaway located to the Caribbean crust. The Costa Rica–Panama arc probably devel-
east of the Bahamas also probably opened at this time, oped off northwestern South America about this time. A new
improving the hydrographic connection of the intra-American FFR triple junction between Africa and the Bahamas Platform
basins with the North Atlantic. heralded the initiation of seafloor spreading between Africa and
Major Atlantic hotspots were located outside the Gulf of South America.
Mexico Basin by this time. The Cape Verde and Sierra The width of the western Gulf of Mexico Basin approached
Leone–Ceara Rise hotspots jumped to the east of the North 700 km in the Late Aptian. Kinematic analysis predicts that
Atlantic spreading centre (Fig. 8b, c). most Gulf of Mexico seafloor (c. 60%) was created in the
Aptian, c. 125–115 Ma, i.e. during the Cretaceous period
of stable normal geomagnetic polarity, which lasted from
Earliest Aptian (125 Ma) c. 125 Ma to c. 83.5 Ma (Gradstein et al. 2004).
The c. 125 Ma geotectonic reconstruction of the Gulf of James–Rodessa–Glen Rose–La Pena–Las Uvas–equivalent
Mexico Basin (Figs. 9a–c, 13) provides views of a widening accumulation rate and lithofacies maps (Fig. 10c) reveal depo-
western Gulf at the beginning of the Aptian (Fig. 1). Sligo– sitional architecture dominated by extensive shallow carbonate
Pettet–Cupido–La Virgen (carbonate) and Hosston–Upper platforms and deep open-marine basins separated by robust
Casita (siliciclastic) basin-margin deposystems dominated the James-equivalent continental margin reef systems (Figs. 10b, c,
depositional architecture of the basin (Fig. 9c). 12). A thick basin-margin Aptian section is also probably
Back-arc spreading in the western Gulf Basin, northwest- contained in the large allochthons located beneath the South
ward subduction of proto-Caribbean crust, and counterclock- Louisiana and Texas offshore (Fig. 10c; Fillon et al. 2006). The
wise rotation of Yucatan southeastward along the East Mexican persistence of nearshore volcanics and emergent areas in
Transform continued through this interval. The proto- southern Yucatan and adjacent terranes suggests the Greater
Caribbean transform may have extended eastward at this time, Antilles arc remained active along the southern Yucatan margin,
with a fault–fault–ridge (FFR) triple junction forming near the with continued subduction beneath Yucatan, and an active
eastern tip of the Bahamas Platform. The Tuxpan platform transform or remnant arc in northwest Chortis. Concurrent
122 R. H. Fillon

development of the Panama–Costa Rica arc (Fig. 10b) probably of the basin a few million years later (Cox & Van Arsdale
caused shoaling of the Panamanian–Costa Rican seaway. The 2002). Diamonds from the alkalic Murfreesboro lamproite in
Mexican and Central Atlantic seaways remained open, connect- southwestern Arkansas (dated c. 106–97 Ma) and alkalic basalts
ing the Gulf and other intra-American basins to the Pacific and from the Gulf rim (dated between 82 Ma and 69 Ma, Byerly
North Atlantic Oceans. 1991; Fig. 4) provide evidence of increasing mantle–upper
crustal interaction toward the end of the Early Cretaceous and
Latest Cenomanian (94 Ma) into the Late Cretaceous.
The Yucatan microplate lies within the intra-American plate
Western Gulf of Mexico back-arc spreading and west-dipping boundary zone of interfering North and South Atlantic stresses
subduction of proto-Caribbean seafloor beneath the Yucatan throughout the Early Cretaceous. Beginning in the Late Albian,
microplate, which are active elements throughout most of the
however, Yucatan moves out of the boundary zone and
Lower Cretaceous, came to an end during the Late Albian. In
is sutured (albeit loosely) to North America. In the Late
the c. 94 Ma latest Cenomanian geotectonic reconstruction of
the Gulf Basin (Figs. 1, 11a–c, 13) Yucatan had reached its Cretaceous and Cenozoic, the North American–Caribbean
modern alignment with respect to North America, creating a plate boundary passes along Yucatan’s southern margin (Pindell
western Gulf of Mexico Basin c. 800 km wide. & Kennan 2001b). The geologically measurable Late Cenozoic
During the Cenomanian, expansion of the developing South counterclockwise rotation of Yucatan with respect to North
Atlantic spreading centre led to abandonment of the Greater America, reported by Burkart & Scotese (2001), provides direct
Antilles arc and creation of a new, Lesser Antilles (Aves) evidence for the impact of interfering mantle stresses near
arc-trench system. This new arc spanned the gap between the the intra-American plate boundary zone. During the Early
eastern tip of the Bahamas Platform and South America, Cretaceous when Yucatan appears to actually lie within the
shifting the locus of west-dipping subduction further to the boundary zone, the impact and resulting rotation are likely to
east. Central Atlantic oceanic crust, generated by the South have been of considerably greater magnitude.
Atlantic spreading centre, was subducted beneath the new arc.
Cessation of subduction beneath the Greater Antilles arc
removed a critical proto-Caribbean plate boundary and there-
fore signalled extinction of the proto-Caribbean plate. Its IMPLICATIONS FOR GULF OF MEXICO
remaining crust was probably sutured onto North America at PETROLEUM SYSTEMS
this time. The new arc created an intra-American Caribbean The Early Jurassic Gulf of Mexico rift system probably looked
plate (iaC in Fig. 11b) conjugate to the extinct proto-Caribbean very much like the Rio Grande Rift Valley does today. Rifted
plate, which was replaced by the unique Caribbean plate (sensu terrane punctuated by steep fault-block mountains like the
Pindell & Kennan 2001b) that replaced it in the Cenozoic. Permian Guadalupe Mountain fault block and alluvial basins
The postulated Cenomanian intra-American Caribbean plate receiving Eagle Mills-equivalent deposition indicate an arid
consisted largely of oceanic crust. It was bordered to the subaerial environment. The Gulf of Mexico rift system evolved
northwest by the still active proto-Caribbean ridge and the into the western Gulf of Mexico Basin with accelerated
North American plate, to the southeast by a transform fault- subsidence triggered by the removal of lower crust (Karner &
trench system situated along the northern margin of South Driscoll 1999) in the Late Triassic–Early Jurassic (c. 228–
America, and to the southwest by the developing Costa 176 Ma). This stage was followed in the Middle–Late Jurassic
Rica–Panama arc and the Farallon plate.
(c. 176–146 Ma) by marine inundation and source-rock deposi-
Basin margin reef development peaked in the Albian, with
tion. By the Late Berriasian or Early Valanginian (c. 140 Ma),
the vigorous expansion of Stuart City–Aurora-equivalent bio-
herms around the Gulf rim (Figs. 10c, 12). During the back-arc spreading had begun in the western Gulf Basin and
Cenomanian, however, basin-margin reefs were much less steepened basin margins would have been inclined to fail,
significant, possibly due to regional thermal uplift related to leading to the development of large detachments on the
subducting slab detachment which may be responsible for the underlying Callovian salt. Early Cretaceous source rocks were
heating and volcanics attributed to a suggested Aves–Beata deposited in both carbonate and siliciclastic depocentres within
Ridge ‘hotspot’ or melting anomaly (Figs. 1, 12). the widening and deepening Gulf Basin.
Accumulation rates and lithofacies of Cenomanian Buda– In contrast to the western Gulf Basin, in the eastern Gulf
Lower Tuscaloosa–Lower Eagle Ford–Cuesta del Cura–Agua Basin, Berriasian–Barremian forearc compression buttressed
Nueva equivalents record extensive shallow carbonate plat- basin margins, potentially even folding Early Cretaceous strata
forms with minor shelf-edge build-ups and deep open-marine to create features such as the Destin Dome anticline (Dobson
facies (Fig. 11c). In the Cenomanian, nearshore volcanics and & Buffler 1997). Buttressing and lesser volumes of salt would
emergent areas in southern Yucatan and northern Chortis have reduced the scale of and delayed the onset of detachment
record the last stages of Greater Caribbean arc activity. faulting until after southeastward arc migration relieved forearc
Held stationary with respect to Africa, the postulated Aves– compression of the northwest Florida margin (Fillon 2001).
Beata Ridge melting anomaly was centred in the Gulf of Mexico In a tectonically active Early Cretaceous Gulf of Mexico
Basin at the end of the Cenomanian (Fig. 11c). Evidence for Basin, uplift, subsidence, fluid expulsion and hydrocarbon
this thermal anomaly comes from reports of Caribbean Large migration pathways were affected by vigorous post- and syn-
Igneous Province (CLIP) alkalic basalts dated to about 81 Ma depositional stress fields. Detachment nappes covering oceanic
on the Beata Ridge (Sinton et al. 2000) and northern Aves Ridge crustal margins of the widening Gulf, and volcanism related to
(Fox et al. 1971). Bird et al. (2005) detailed evidence for subducted slab detachment, provided post-depositional and
hotspot-generated crust in the western Gulf Basin, although syn-depositional environments in which hydrothermal fluids
they suggested a greater age. could diagenetically alter formations of Early Cretaceous age,
Heating events further north in the Mississippi Embayment potentially contributing to hydrocarbon maturation, expulsion,
portion of the Gulf Basin are thought to be associated with dolomitization, secondary porosity enhancement and illite–
the Bermuda hotspot, which transited the northern limits smectite conversion.
Gulf of Mexico back-arc tectonics 123

SUMMARY Baksi, A.K. 1997. The timing of Late Cretaceous alkalic igneous activity in the
northern Gulf of Mexico basin, southeastern USA. Journal of Geology, 105,
This study reports an acceptable kinematic solution for the 629–643.
Mesozoic rotation of Yucatan based on cumulative North Barnett, R.S. 1975. Basement structure of Florida and its tectonic
American and South American plate rotations and related implications. Gulf Coast Association of Geological Societies Transactions, 25,
122–142.
back-arc spreading. It supports a planetary geodynamic view of Bartolini, C., Lang, H., Cantú-Chapa, A. & Barboza-Gudiño, R. 2001. The
Gulf Basin evolution in which the creation of oceanic crust in Triassic Zacatecas Formation in central Mexico: Paleotectonic, paleogeo-
the Gulf of Mexico is linked directly to neighbouring larger- graphic, and paleobiogeographic implications. In: Bartolini, C., Buffler, R.T.
plate motions and merging North and South Atlantic stress & Cantú-Chapa, A. (eds) The western Gulf of Mexico Basin: Tectonics, sedimentary
basins, and petroleum systems. American Association of Petroleum Geologists
fields. Although several hundred kilometres of crustal extension Memoir, 75, 295–315.
separated Yucatan from North America in the Jurassic, North Becker, T.W., Kellogg, J.B., Ekstrom, G. & O’Connell, R.J. 2003. Compari-
and South American rotation poles and the timing of initial son of azimuthal seismic anisotropy from surface waves and finite strain
South Atlantic rifting suggest an earliest Cretaceous age for from global mantle-circulation models. Geophysical Journal International, 155,
696–714.
initiation of the seafloor spreading and oceanic crust formation Bertrand, H. 1991. The Mesozoic tholeiitic province of northwest Africa: a
which produced most of the separation of Yucatan from Texas volcano-tectonic record of the early opening of the central Atlantic. In:
and opened the western Gulf of Mexico. Angular velocities of Kampunzo, A.B. & Lubala, R.T. (eds) The Phanerozoic African Plate.
Atlantic spreading predict that Yucatan rotation slowed dra- Springer, New York, 147–191.
Bird, D.E., Burke, K., Hall, S.A. & Casey, J.F. 2005. Gulf of Mexico tectonic
matically and largely ceased in the Late Albian (c. 109–102 Ma). history: hotspot tracks, crustal boundaries, and early salt distribution.
The Late Triassic fit of a Yucatan (Maya) block against the American Association of Petroleum Geologists Bulletin, 89, 311–328.
Texas–Southwest Louisiana Palaeozoic margin and the subse- Blair, T.C. 1988. Mixed siliciclastic-carbonate marine and continental syn-rift
quent movement of Yucatan predicted by the interaction of sedimentation, Todos Santos and San Ricardo Formations, western
Chiapas, Mexico. Journal of Sedimentary Petrology, 58, 623–636.
North and South Atlantic stress fields is consistent with the
Blum, J.D., Chamberlain, C.P., Hingston, M.P., Koeberl, C., Marin, L.E.,
known distribution of Mesozoic deposystems along the north- Schuraytz, B.C. & Sharpton, V.L. 1993. Isotopic comparison of K/T
ern and southern margins of the Gulf Basin. Thick salt- boundary impact glasses with melt rock from the Chicxulub and Manson
lubricated detachment allochthons of pre-drift strata may mask impact structures. Nature, 364, 325–327.
the oldest oceanic crust in the Gulf Basin, formed prior to Bowin, C. 1966. Geology of the Central Dominican Republic (case history of
part of an island arc). In: Hess, H. (ed.) Caribbean geological studies. Geological
c. 120 Ma. Based on the implied time-scale, up to 60% of Gulf Society of America Memoir, 98, 11–84.
of Mexico seafloor could have been generated in the Aptian Braunstein, J. & McMichael, C.E. 1976. Door Point: a buried volcano in
(c. 125–112 Ma), during the Early Cretaceous stable normal southeast Louisiana. Gulf Coast Association of Geological Societies Transactions,
geomagnetic polarity episode, which spanned the interval 26, 79–80.
125–83.5 Ma. Braunstein, J., Huddlestun, P. & Beil, R. 1985. American Association of Petroleum
Geologists correlation of stratigraphic units in North America. American Associ-
Oceanic plateau basalts and other deep mantle-derived ation of Petroleum Geologists (COSUNA), Gulf Coast, Chart of the Gulf
volcanic rocks in the Gulf of Mexico and Caribbean Basins are Coast Region. AAPG, Tulsa, OK.
consistent with hotspots or melting anomalies being overridden Breitsprecher, K. & Thorkelson, D.J. 2001. Spatial coincidence of the
by the Gulf Basin as North America moved away from Africa, Kula–Farallon slab window with trench-distal volcanism of the Eocene
Magmatic Belt in the southern Cordillera. In: Cook, F. & Erdmer, P.
beginning in the Late Jurassic and continuing through the (compilers) (eds) Slave-Northern Cordillera Lithospheric Evolution (SNORCLE)
Cenomanian. Transect and Cordilleran Tectonics Workshop Meeting (February 22–25). Litho-
probe Report, 79. Pacific Geoscience Centre, 178–183.
The author wishes to acknowledge the organizers of the Return to Buffler, R.T. 1991. Seismic stratigraphy of the deep Gulf of Mexico basin and
Rifts conference and the Geological Society for hosting a marvellous adjacent margins. In: Salvador, A. (ed.) The Gulf of Mexico basin. The
meeting. Texaco (now Chevron), Louisiana State University, Paleo- Geology of North America J. Geological Society of America, Boulder,
Data, Inc. and Earth Studies Group provided support for the study Colorado, 353–387.
in its various stages. The Norwegian Geologic Survey is thanked for Buffler, R.T. & Sawyer, D.S. 1985. Distribution of crust and early history,
generously making GMAP software available on the world-wide Gulf of Mexico Basin. Gulf Coast Association of Geological Societies Transactions,
35, 333–334.
web. Bunge, H-P. & Grand, S.P. 2000. Mesozoic plate-motion history below the
northeast Pacific Ocean from seismic images of the subducted Farallon
slab. Nature, 405, 337–340.
Burkart, B. & Scotese, C. R. 2001. Cenozoic rotation of the Yucatan (Maya)
REFERENCES block along the Orizaba fault zone of southern Mexico and the faults of
Central America. Abstract presented at the Geological Society of America
Acton, G.D., Galbrun, B. & King, J.W. 2000. Paleolatitude of the Caribbean meeting. Available at: http://gsa.confex.com/gsa/2001AM/finalprogram/
plate since the late Cetaceous. In: Leckie, R.M., Sigurdsson, H., Acton, abstract_28261.htm.
G.D. & Draper, G. (eds) Proceedings of the Ocean Drilling Program, Scientific Burke, K. 1988. Tectonic evolution of the Caribbean. Annual Review of Earth
Results, 165. College Station, Texas, 149–173. and Planetary Sciences, 16, 201–230.
Almy, C.C. Jr 1969. Sedimentation and tectonism in the Upper Cretaceous Burke, K., Cooper, C., Dewey, J.F., Mann, P. & Pindell, J.L. 1984. Caribbean
Puerto Rican portion of the Caribbean Island Arc. Gulf Coast Association of tectonics and relative plate motions. In: Bonini, W.E., Hargraves, R.B. &
Geological Societies Transactions, 19, 269–279. Shagam, R. (eds) The Caribbean–South America plate boundary and regional
Alvarez, W., Claeys, P. & Kieffer, S.W. 1995. Emplacement of KT boundary tectonics. Geological Society of America Memoir, 162, 31–64.
shocked quartz from Chicxulub crater. Science, 269, 930–935. Byerly, G.R. 1991. Igneous activity. In: Salvador, A. (ed.) The Gulf of Mexico
Anderson, D.L. 2001. Topside tectonics. Science, 293, 2016–2018. basin. The Geology of North America J. Geological Society of America,
Anderson, D.L. & Natland, J.H. 2005. A brief history of the plume Boulder, Colorado, 91–108.
hypothesis and its competitors: concept and controversy. In: Foulger, G.R., Cande, S., Labrecque, J.L. & Haxby, W.F. 1988. Plate kinematics of the South
Natland, J.H., Presnall, D.C. & Anderson, D.L. (eds) Plates, Plumes & Atlantic: Chron C34 to present. Journal of Geophysical Research, 93,
Paradigms. Geological Society of America Special Paper, 388, 119–146. 13479–13492.
Anderson, T.H. & Schmidt, V.A. 1983. The evolution of middle America and Carfantan, C.J. 1977. La cobijadura de Motozintla–un paleoarco volcanico en
the Gulf of Mexico–Carribean Sea region during Mesozoic time. Geological Chiapas. Revista, Instituto de Geologia, UNAM, 1, 133–137.
Society of America Bulletin, 94, 941–966. Castle, J.C., Creager, K.C., Winchester, J.P. & van der Hilst, R.D. 2000. Shear
Arden, D.D. Jr 1969. Geologic History of the Nicaraguan Rise. Gulf Coast wave speeds at the base of the mantle. Journal of Geophysical Research, 105,
Association of Geological Societies Transactions, 19, 295–309. 21543–21558.
Atwater, T. & Stock, J. 1998. Pacific-North America plate tectonics of the Celaya, M. & McCabe, R. 1987. Kinematic model for the opening of
Neogene southwestern United States: an update. International Geology Review, the Sea of Japan and the bending of the Japanese islands. Geology, 15,
40, 375–402. 53–57.
124 R. H. Fillon

Christenson, G. 1990. The Florida lineament. Gulf Coast Association of Geological Global and Gulf of Mexico Experience. Gulf Coast Section of SEPM (Society
Societies Transactions, 40, 99–115. for Sedimentary Geology) Publication on CD-ROM.
Cole, G.A., Yu, A., Peel, F. et al. 2001. The Deep Water Gulf of Mexico Fillon, R.H., Waterman, A.S. & Lawless, P.N. 2005. Paleocene–Eocene
Petroleum System: Insights from Piston Coring, Defining Seepage, Deposystems in the Gulf of Mexico: Petroleum System Implications. Gulf
Anomalies, and Background. In: Fillon, R.H., Rosen, N., Weimer, P. et al. Coast Association of Geological Societies Transactions, 55, 195–222.
(eds) Petroleum Systems of Deep-Water Basins: Global and Gulf of Mexico Fillon, R.H., Stieglitz, T., Matava, T. & Dinkelman, M.G. 2006. Gulf of
Experience. Gulf Coast Section of SEPM (Society for Sedimentary Geology) Mexico deep seismic imaging and Mesozoic interpretation initiative. SEG
Publication on CD-ROM. Expanded Abstracts, 25, 1008. doi:10.1190/1.2369683.
Cox, R. & Van Arsdale, R. 2002. The Mississippi Embayment, North Fox, P.J., Schreiber, E. & Heezen, B.C. 1971. The geology of the Caribbean
America: a first order continental structure generated by the Cretaceous crust: Tertiary sediments, granitic and basic rocks from the Aves Ridge.
superplume mantle event. Journal of Geodynamics, 34, 163–176. Tectonophysics, 12, 89–109.
Dalziel, I.W.D., Lawver, L.A. & Murphy, J.B. 2000. Plumes, orogenesis, and Freeland, G.L. & Dietz, R.S. 1971. Plate tectonic evolution of Caribbean–
supercontinental fragmentation. Earth and Planetary Science Letters, 178, 1–11. Gulf of Mexico region. Nature, 232, 20–23.
de Cserna, Z. 1976. Geology of the Fresnillo area, Zacatecas, Mexico. Gao, W., Grand, S.P., Baldridge, W.S., Wilson, D., West, M., Ni, J.F. & Aster,
Geological Society of America Bulletin, 87, 1191–1199. R. 2004. Upper mantle convection beneath the central Rio Grande rift
DeMets, C. 1995. Plate motions and crustal deformation. U. S. National Report to imaged by P and S wave tomography. Journal of Geophysical Research, 109,
International Union of Geodesy and Geophysics, 1991–1994. Reviews of B03305, doi:10.1029/2003JB002743.
Geophysics, 33. Supplement, American Geophysical Union, Washington, García-Casco, A., de Arce, P., Millán, G. et al. 2003. Metabasites from the
D. C. northern serpentinite belt (Cuba) and a metamorphic perspective of the
Dickinson, W.R. 1970. Relations of andesites, granites, and derivative plate tectonic models for the Caribbean region. Tectónica de placas en el
sandstones to arc-trench territories. Reviews of Geophysics and Space Physics, 8, Caribe. Taller del proyecto n 433 del PICG / UNESCO. TPICG. 07,
813–860. Memorias Geomin 2003, la Habana, 24–28 de marzo (ISBN
Dickinson, W.R. & Coney, P.J. 1980. Plate tectonic constraints on the origin 959-7117-11-8), 29–37.
of the Gulf of Mexico. In: Pilger, R.H. Jr (ed.) The origin of the Gulf of Mexico Garfunkel, Z. 1986. Review of oceanic transform activity and development.
and the early opening of the central North Atlantic Ocean. Department of Geology, Journal of the Geological Society, London, 143, 775–784.
Louisiana State University and Louisiana Geological Survey, Proceedings Gastil, R.G. & Fenby, S.S. 1991. Detachment faulting as a mechanism for
of Symposium, 3–5 March, 27–36. tectonically filling the Gulf of California during dilation. In: Dauphin, J.P.
Dickinson, W.R. & Lawton, T.F. 2001. Carboniferous to Cretaceous assem- & Simoneit, B.R.T. (eds) The Gulf and Peninsular Province of the Californias.
bly and fragmentation of Mexico. Geological Society of America Bulletin, 113, American Association of Petroleum Geologists Memoir, 47, 371–375.
1142–1160. Goldhammer, R.K. & Johnson, C.A. 2001. Middle Jurassic–Upper Creta-
Dillon, W.P., Paull, C.K., Buffler, R.T. & Fail, J-P. 1979. Structure and ceous paleogeographic evolution and sequence-stratigraphic framework of
development of the southeast Georgia Embayment and northern Blake the northwest Gulf of Mexico rim. In: Bartolini, C., Buffler, R.T. &
Plateau: preliminary analysis, rifted margins. In: Watkins, J.S., Montadert, L Cantú-Chapa, A. (eds) The western Gulf of Mexico Basin: Tectonics, sedimentary
& Dickerson, P. (eds) Geological and Geophysical Investigations of Continental basins, and petroleum systems. American Association of Petroleum Geologists
Margins. American Association of Petroleum Geologists Memoir, 29, Memoir, 75, 45–81.
27–41. Gradstein, F.M., Ogg, J.G. & Smith, A. 2004. Geologic Time Scale 2004.
Dobson, L.M. & Buffler, R.T. 1997. Seismic stratigraphy and geologic history Cambridge University Press, Cambridge.
of Jurassic rocks, northeastern Gulf of Mexico. American Association of Grand, S.P., van der Hilst, R.D. & Widiyantoro, S. 1997. Global seismic
Petroleum Geologists Bulletin, 81, 100–120. tomography: a snapshot of convection in the Earth. Geological Society of
Dunbar, J.A. & Sawyer, D.S. 1987. Implications of continental crust exten- America Today, 7, 1–7.
sion for plate reconstruction: An example from the Gulf of Mexico. Hall, S.A. & Najmuddin, I.J. 1994. Constraints on the tectonic development
Tectonics, 6, 739–755. of the eastern Gulf of Mexico provided by magnetic anomaly data. Journal
Duncan, R.A. & Hargraves, R.B. 1984. Plate tectonic evolution of the of Geophysical Research, 99, 7161–7175.
Caribbean region in the mantle reference frame. In: Bonini, W.E., Horne, G.S., Clark, G.S. & Pushkar, P. 1976. Pre-Cretaceous rocks of
Hargraves, R.B. & Shagam, R. (eds) The Caribbean–South America plate northwestern Honduras: basement terrane in Sierra de Omoa. American
boundary and regional tectonics. Geological Society of America Memoir, 162, Association of Petroleum Geologists Bulletin, 60, 566–583.
81–93. Horsefield, W.T. & Roobol, M.J. 1974. A tectonic model for the evolution of
Dziewonski, A.M. 2005. The robust aspects of global seismic tomography. In: Jamaica. Journal of the Geological Society of Jamaica, 14, 31–38.
Foulger, G.R., Anderson, D.L., Natland, J.H. & Presnall, D.C. (eds) Plates, Hoshi, H. & Matsubara, T. 1998. Early Miocene paleomagnetic results from
Plumes & Paradigms. Geological Society of America Special Paper, 388, the Ninohe area, NE Japan: Implications for arc rotation and intra-arc
147–154. differential rotations. Earth Planets Space, 50, 23–33.
Eguiluz de Antuñano, S. 2001. Geologic evolution and gas resources of the Huang, J., Zhong, S. & van Hunen, J. 2003. Controls on sublithospheric
Sabinas Basin in northeastern Mexico. In: Bartolini, C., Buffler, R.T. & small-scale convection. Journal of Geophysical Research, 108(B8), 2405, doi:10.
Cantú-Chapa, A. (eds) The western Gulf of Mexico Basin: Tectonics, sedimentary 1029/2003JB002456.
basins, and petroleum systems. American Association of Petroleum Geologists Hunter, B.E. & Davies, D.K. 1979. Distribution of Volcanic sediments in the
Memoir, 75, 241–270. Gulf Coastal Province – Significance to Petroleum Geology. Gulf Coast
Engebretson, D.C., Cox, A. & Gordon, R.G. 1986. Relative motions between Association of Geological Societies Transactions, 29, 147–155.
oceanic and continental plates in the Pacific Basin. Geological Society of Iturralde-Vinent, M.A. 1994. Cuban geology: A new plate tectonic synthesis.
America Special Paper, 206, 1–59. Journal of Petroleum Geology, 17, 39–70.
Escartin, J., Smith, D.K. & Cannat, M. 2003. Parallel bands of seismicity at Iturralde-Vinent, M.A. 1996. Estratigrafia del arco volcanico Cretacico en
the Mid-Atlantic Ridge, 12–14N. Geophysical Research Letters, 30, Cuba: Evidencias de un arco primitivo (Cretacio Inferior) en Cuba. In:
1620–1623. Iturralde-Vinent, M.A. (ed.) Ofiolitas y Arcos Volcanicos de Cuba. IUGS
Fairhead, J.D. & Wilson, M. 2005. Plate tectonic processes in the South Project 364: Caribbean Ophiolites and Volcanic Arcs, Special
Atlantic Ocean: do we need deep mantle plumes?. In: Foulger, G.R., Contribution, 1.
Anderson, D.L., Natland, J.H. & Presnall, D.C. (eds) Plates, Plumes & Iturralde-Vinent, M.A. 1997. Introduccion a la geologia de Cuba. In:
Paradigms. Geological Society of America Special Paper, 388, 537–554. Furrazola, B.Z.F. & Nunez, C.B.E. (eds) Estudios sobre geologia de Cuba.
Feng, J. 1995. Post Mid-Cretaceous seismic stratigraphy and depositional history, deep Centro Nacional de Informacion Geologica, Havana, Cuba, 35–68.
Gulf of Mexico. PhD dissertation. University of Texas at Austin. Jackson, T.A. & Smith, T.E. 1979. The tectonic significance of basalts and
Feng, J. & Buffler, R.T. 1991. Preliminary age determinations for new deep dacites in the Wagwater Belt, Jamaica. Geological Magazine, 116, 365.
Gulf of Mexico Basin seismic sequences. Gulf Coast Association of Geological Jacques, J.M. & Clegg, H. 2002. Late Jurassic source rock distribution and
Societies Transactions, 41, 283–289. quality in the Gulf of Mexico: Inferences from plate tectonic modeling.
Ferrari, L., López-Martínez, M., Aguirre-Díaz, G. & Carrasco-Núñez, G. Gulf Coast Association of Geological Societies Transactions, 52, 429–440.
1999. Space–time patterns of Cenozoic arc volcanism in central Mexico: James, K.H. 2003. Caribbean plate origin: discussion of arguments claiming
From the Sierra Madre Occidental to the Mexican Volcanic Belt. Geology, to support a Pacific origin; arguments for an in-situ origin. Extended
27, 303–306. Abstract presented at the American Association of Petroleum Geologists
Filho, A.T., Pimentel Mizusaki, A.M., Milani, E.J. & de Cesero, P. 2000. International Conference Barcelona, Spain, September 21–24.
Rifting and magmatism associated with the South America and Africa Karner, G.D. & Driscoll, N.W. 1999. Tectonic and stratigraphic development
break up. Revista Brasileira de Geociências, 30, 017–019. of the West African and eastern Brazilian Margins: insights from quanti-
Fillon, R.H. 2001. Late Mesozoic and Cenozoic deposystem evolution in the tative basin modelling. In: Cameron, N.R., Bate, R.H. & Clure, V.S. (eds)
eastern Gulf of Mexico: implications for hydrocarbon migration. In: Fillon, The Oil and Gas Habitats of the South Atlantic. Geological Society, London,
R.H., Rosen, N., Weimer, P. et al. (eds) Petroleum Systems of Deep-Water Basins: Special Publications, 153, 11–40.
Gulf of Mexico back-arc tectonics 125

Keppie, J.D. 2004. Terranes of Mexico Revisited: A 1.3 Billion Year Odyssey. Pacific margin. Journal of Geophysical Research, 111, B02401,
International Geology Review, 46, 765–794. doi:10.1029/2005JB003705.
Kerr, A., Iturralde-Vinent, M., Saunders, A., Babbs, T. & Tarney, J. 1999. A Monod, O. & Calvet, P.H. 1992. Structural and stratigraphic reinterpretation
geochemical recognaissance of Cuban Mesozoic volcanic rocks: implica- of the Triassic units near Zacatecas, central Mexico: Evidence of a
tions for plate tectonics models of the Caribbean. Geological Society of America Laramide nappe pail. Zentralblatt für Geologie und Paläontologie, Teil I, H.6,
Bulletin, 111, 1581–1599. 1533–1544.
Kerr, A.C., White, R.V., Thompson, P.M.E., Tarney, J. & Saunders, A.D. Moran-Zenteno, D. 1994. Geology of the Southeastern Region of Mexico. In:
2003. No oceanic plateau–no Caribbean plate? The seminal role of an The Geology of the Mexican Republic. American Association of Petroleum
oceanic plateau in Caribbean plate evolution. In: Bartolini, C., Buffler, R.T. Geologists, Studies in Geology, 39, 75–83.
Moretti, I., Tenreyro, T., Linares, E. et al. 2003. Petroleum system of the
& Blickwede, J. (eds) The Circum-Gulf of Mexico and the Caribbean: Hydrocarbon
Cuban northwest offshore zone. In: Bartolini, C., Buffler, R.T. &
habitats, basin formation, and plate tectonics. American Association of Petroleum
Blickwede, J. (eds) The Circum-Gulf of Mexico and the Caribbean: Hydrocarbon
Geologists Memoir, 79, 126–168.
habitats, basin formation, and plate tectonics. American Association of Petroleum
Kharin, G.S., Shirshov, P.P., Arakeljanz, M.M. & Dmitriev, I. 1977. Petrology
Geologists Memoir, 79, 675–696.
and K–Ar age of basaltic rocks, sites 353, 354, and 355, DSDP Leg 39. In:
Müller, R.D., Roest, W.R., Royer, J-Y., Gahagan, L.M. & Sclater, J.G. 1997.
Perch-Nielsen, K., Supko, P.R. et al. (eds) Initial Reports of the Deep Sea
Digital isochrons of the world’s ocean floor. Journal of Geophysical Research,
Drilling Project, 39. U.S. Government Printing Office, Washington, 231–327.
102 (B2), 3211–3214.
Klitgord, K.D. & Schouten, H. 1986. Plate kinematics of the central Atlantic.
Müller, R.D., Cande, S.C., Royer, J.-Y., Roest, W.R. & Maschenkov, S. 1999.
In: Vogt, P.R. & Tucholke, B.E. (eds) The Western North Atlantic Region. The
New constraints on the Late Cretaceous/Tertiary plate tectonic evolution
Geology of North America, Volume M. Geological Society of America,
of the Caribbean. In: Mann, P. (ed.) Caribbean Basins. Sedimentary Basins of
Boulder, Colorado, 351–378.
the World, 4. Elsevier Science, Amsterdam, 39–55.
Knight, K.B., Nomade, S., Renne, P.R., Marzoli, A., Bertrand, H. & Youbi, Paine, W.R. & Meyerhoff, A.A. 1970. Gulf of Mexico Basin: Interactions
N. 2004. The Central Atlantic Magmatic Province at the Triassic–Jurassic among tectonics, sedimentation, and hydrocarbon accumulation. Gulf Coast
boundary: paleomagnetic and 40Ar/39Ar evidence from Morocco for Association of Geological Societies Transactions, 20, 5–44.
brief, episodic volcanism. Earth and Planetary Science Letters, 228 (1–2), Parsons, T., Thompson, G.A. & Sleep, N.A. 1994. Mantle plume influence on
143–160. the Neogene uplift and extension of the US western Cordillera? Geology, 22,
Lasky, S.A. 1938. Newly discovered section of Trinity age in southwestern 83–86.
New Mexico. American Association of Petroleum Geologists Bulletin, 22, 524–540. Phair, R.L. & Buffler, R.T. 1983. Pre-middle Cretaceous geologic history of
Lawton, T.F., Vega, F.J., Giles, K.A. & Rosales-Domínguez, C. 2001. the deep southeastern Gulf of Mexico. In: Bally, A.W. (ed.) Seismic expression
Stratigraphy and origin of the La Popa Basin, Nuevo León and Coahuila, of structural styles — a picture and work atlas. American Association of
Mexico. In: Bartolini, C., Buffler, R.T. & Cantú-Chapa, A. (eds) The western Petroleum Geologists Studies in Geology, 15-2, 141–147.
Gulf of Mexico Basin: tectonics, sedimentary basins, and petroleum systems. American Pindell, J.L. & Barrett, S.F. 1990. Geological evolution of the Caribbean
Association of Petroleum Geologists Memoir, 75, 219–240. region; a plate tectonic perspective. In: Dengo, G. & Case, J.E. (eds) The
Lopez, R.E. 1981. Geologia de Mexico Tomo III, second edition. Universidad Caribbean. Volume H. Decade of North American Geology. Geological
Nacional Autonoma de Mexico, Mexico. Society of America, Boulder, Colorado, 405–432.
Lugo, J. & Mann, P. 1995. Jurassic–Eocene tectonic evolution of Maracaibo Pindell, J.L. & Dewey, J.F. 1982. Permo-Triassic reconstruction of western
Basin, Venezuela. In: Tankard, A.J., Suárez, S.R. & Welsink, H.J. (eds) Pangaea and the evolution of the Gulf of Mexico/Caribbean region.
Petroleum basins of South America. American Association of Petroleum Tectonics, 1, 179–211.
Geologists Memoir, 62, 699–725. Pindell, J. & Kennan, L. 2001a. Kinematic evolution of the Gulf of Mexico
Magoon, L.B., Hudson, T.L. & Cook, H.E. 2001. Pimienta-Tamabra(!)—A and Caribbean. In: Fillon, R.H., Rosen, N., Weimer, P. et al. (eds) Petroleum
giant supercharged petroleum system in the southern Gulf of Mexico, Systems of Deep-Water Basins: Global and Gulf of Mexico Experience. Gulf Coast
onshore and offshore Mexico. In: Bartolini, C., Buffler, R.T. & Section of SEPM (Society for Sedimentary Geology) Publication on
Cantú-Chapa, A. (eds) The western Gulf of Mexico Basin: Tectonics, sedimentary CD-ROM.
basins, and petroleum systems. American Association of Petroleum Geologists Pindell, J.L. & Kennan, L. 2001b. Processes and events in the Terrane
Memoir, 75, 83–125. Assembly of Trinidad and eastern Venezuela. In: Fillon, R.H., Rosen, N.,
Marton, G. & Buffler, R.T. 1994. Jurassic reconstruction of the Gulf of Weimer, P. et al. (eds) Petroleum Systems of Deep-Water Basins: Global and Gulf
Mexico Basin. International Geology Review, 36, 545–586. of Mexico Experience. Gulf Coast Section of SEPM (Society for Sedimentary
Marzoli, A., Renne, P.R., Piccirillo, E.M., Ernesto, M., Bellieni, G. & De Min, Geology) Publication on CD-ROM.
A. 1999. Extensive 200 million-year-old continental flood basalts of the Pindell, J.L. 1985. Alleghenian reconstruction and subsequent evolution of
central Atlantic magmatic province. Science, 284, 616–618. the Gulf of Mexico, Bahamas and proto-Caribbean. Tectonics, 4, 1–39.
McHone, J.G. 2000. Non-plume magmatism and rifting during the opening Pindell, J.L. 1994. Evolution of the Gulf of Mexico and the Caribbean. In:
of the central Atlantic Ocean. Tectonophysics, 316, 287–296. Donovan, S.K. & Jackson, T.A. (eds) Caribbean geology: An introduction.
McHone, J.G., Anderson, D.L., Beutel, E.K. & Fialko, Y.A. 2005. Giant University of the West Indies Publishers Association, Kingston, 13–39.
dikes: patterns and plate tectonics. In: Foulger, G.R., Anderson, D.L., Pockalny, R.A., Larson, R.L., Viso, R.F. & Abrams, L.J. 2002. Bathymetry and
Natland, J.H. & Presnall, D.C. (eds) Plates, Plumes & Paradigms. Geological gravity profiles across a Mid-Cretaceous triple junction trace in the
Society of America Special Paper, 388, 401–420. southwest Pacific Ocean: Implications for accretion of oceanic crust.
McLemore, V.T., Munroe, E.A., Heizler, M.T. & McKee, C. 2000a. Geology Geophysical Research Letters, 29, 1–4.
and evolution of the Copper Flat Porphyry-Copper and associated mineral Pollitz, F.F., Kellogg, L. & Burgmann, R. 2001. Sinking ma?c body in a
deposits in the Hillsboro mining district, Sierra County, New Mexico. In: reactivated lower crust: a mechanism for stress concentration at the New
Cluer, J.K., Price, J.G., Struhsacker, E.M., Hardyman, R.F. & Morris, C.L. Madrid seismic zone. Bulletin of the Seismic Society of America, 91, 1882–1897.
(eds) Geology and Ore Deposits 2000, The Great Basin and Beyond. Geological Pszczolkowski, A. 1999. The exposed passive margin of North America in
Society of Nevada, Symposium Proceedings, 643–659. Western Cuba. In: Mann, P. (ed.) Caribbean basins. Sedimentary basins of
McLemore, V.T., Peters, L. & Heizler, M.T. 2000b. 40Ar/39Ar geochronol- the world series, 4. Elsevier Science, Amsterdam, 93–121.
ogy of igneous rocks in the Lordsburg district, northern Pyramid Moun- Rezak, R. & Tieh, T.T. 1980. Basalt on Louisiana outer continental
tains, Hidalgo County, New Mexico. In: Lawton, T.F., McMillan, N.J., shelf (Abstract 70). EOS, Transactions American Geophysical Union, 61 (46),
McLemore, V.T., Austin, G. & Barker, J.M. (eds) Southwest Passage, A Trip 989.
through the Phanerozoic. New Mexico Geological Society Guidebook, 51, 6–8. Robinson, E. 1994. Jamaica. In: Donovan, S.K. & Jackson, T.A. (eds)
Meneses-Rocha, J.J. 2001. Tectonic evolution of the Ixtapa graben, an Caribbean Geology: an Introduction. University of the West Indies Publishers’
example of a strike-slip basin in southeastern Mexico: Implications for Association, Kingston, Jamaica, 111–127.
regional petroleum systems. In: Bartolini, C., Buffler, R.T. & Cantú-Chapa, Royer, J.-Y., Muller, R.D. Gahagan, L.M., Lawver, L.A., Mayes, C.L.,
A. (eds) The western Gulf of Mexico Basin: Tectonics, sedimentary basins, and Nurnberg, D. & Sclater, J.G. 1992. A Global Ishochron Chart. University
petroleum systems. American Association of Petroleum Geologists Memoir, of Texas Institute for Geophysics Technical Report 117.
75, 183–216. Sager, W.W., Weiss, C.J., Tivey, M.A. & Johnson, H.P. 1998. Geomagnetic
Meschede, M. & Frisch, W. 1998. A plate-tectonic model for the Mesozoic polarity reversal m tow profiles from the Pacific Jurassic Quiet Zone.
and Early Cenozoic history of the Caribbean Plate. Tectonophysics, 296, Journal of Geophysical Research, 103, 5269–5286.
269–291. Salvador, A. 1985. Chronostratigraphic and geochronometric scales. In:
Michaud, F. 1987. Stratigraphie et paléogéographie du Mésozoïque du Chiapas. COSUNA Stratigraphic Correlation Charts of the United States. American
Mémoire de Stratigraphie, 6. Thesis, Academie de Paris, Université Pierre Association of Petroleum Geologists Bulletin, 69 (2), 181–189.
et Marie Curie. Salvador, A. 1987. Late Triassic–Jurassic paleogeography and origin of Gulf
Miller, M.S., Kennett, B.L.N. & Toy, V.G. 2006. Spatial and temporal of Mexico Basin. American Association of Petroleum Geologists Bulletin, 71,
evolution of the subducting Pacific plate structure along the western 419–451.
126 R. H. Fillon

Saunders, J.A. & Harrelson, D.W. 1992. Age and petrology of the Jackson Van Siclen, D.C. 1983. Early Mesozoic tectonics of northern Gulf of Mexico
Dome igneous-volcanic complex, Mississippi: Implications for the tectonic coastal plain. Gulf Coast Association of Geological Societies Transactions, 33,
history of the Mississippi Salt Dome Basin. American Association of Petroleum 231–240.
Geologists Bulletin, 76, 1468. Vogt, P.R. & Jung, W-Y. 2005. Paired basement ridges: spreading axis
Schneider, J.F., Sacks, I.S., Huaco, D., Norabuena, E. & Flores, A. 1988. migration across mantle heterogeneities?. In: Foulger, G.R., Natland, J.H.,
Spatial distribution and B value of intermediate-depth earthquakes beneath Presnall, D.C. & Anderson, D.L. (eds) Plates, Plumes & Paradigms. Geologi-
central Peru. Geophysical Research Letters, 15, 1421–1424. cal Society of America Special Paper, 388, 555–580.
Sedlock, R.L., Ortega-Gutiérrez, F. & Speed, R.C. 1993. Tectonostratigraphic Walker, K.T., Bokelmann, G.H.R., Klemperer, S.L. & Nyblade, A. 2005.
terranes and tectonic evolution of Mexico. Geological Society of America Special Shear wave splitting around hotspots: Evidence for upwelling-related
Paper, 278.
mantle flow?. In: Foulger, G.R., Natland, J.H., Presnall, D.C. & Anderson,
Shaw, A.B. 1964. Time in Stratigraphy. McGraw-Hill Book Co., New York.
Silver, P.G., Russo, R.M. & Lithgow-Bertelloni, C. 1998. Coupling of South D.L. (eds) Plates, Plumes & Paradigms. Geological Society of America Special
American and African plate motion and plate deformation. Science, 279, Paper, 388, 171–192.
60–63. Walper, J.L. 1980. Tectonic evolution of the Gulf of Mexico. In: Pilger, R.H.
Sinton, C.W., Sigurdsson, H. & Duncan, R.A. 2000. Geochronology and Jr (ed.) The origin of the Gulf of Mexico and the early opening of the central North
petrology of the igneous basement at the lower Nicaraguan Rise, Site 1001. Atlantic Ocean. Department of Geology, Louisiana State University and
In: Leckie, R.M., Sigurdsson, H., Acton, G.D. & Draper, G. (eds) Proceedings Louisiana Geological Survey, Proceedings of Symposium, 3–5 March,
of the Ocean Drilling Program, Scientific Results, 165. College Station, Texas, 87–98.
233–236. Watkins, J.S., Ladd, J.W., Buffler, R.T., Shaub, F.J., Houston, M.H. & Worzel,
Smith, W.H.F. & Sandwell, D.T. 1997. Global Sea Floor Topography from J.L. 1978. Occurrence and evolution of salt in deep Gulf of Mexico. In:
Satellite Altimetry and Ship Depth Soundings. Science, 277, 1956–1962. Bouma, A.H., Moore, G.T. & Coleman, J.M. (eds) Framework, Facies, and
Taylor, M.H., Dillon, W.P. & Pecher, I.A. 2000. Trapping and migration of Oil-Trapping Characteristics of the Upper Continental Margin, Chapter 1. The
methane associated with the gas hydrate stability zone at the Blake Ridge Setting. American Association of Petroleum Geologists Studies in Geology,
Diapir: new insights from seismic data. Marine Geology, 164, 79–89. 7, 43–65.
Torsvik, T.H. & Smethurst, M.A. 1999. Plate tectonic modeling: virtual reality White, G.W. 1980. Permian–Triassic continental reconstruction of the Gulf
with GMAP. Computers & Geosciences, 25, 395–402. of Mexico–Caribbean area. Nature, 283, 823–826.
Van Houten, F.B. 1977. Triassic–Liassic Deposits of Morocco and Eastern Withjack, M.O., Schlische, R.W. & Olsen, P.E. 1998. Diachronous rifting,
North America: Comparison. American Association of Petroleum Geologists drifting, and inversion on the Passive Margin of central eastern North
Bulletin, 61, 79–99. America. American Association of Petroleum Geologists Bulletin, 82, 817–835.
van Hunen, J., van den Berg, A.P. & Vlaar, N.J. 2000. A thermo-mechanical Zhong, S. & Davies, G.F. 1999. Effects of plate and slab viscosities on the
model of horizontal subduction below an overriding plate. Earth and geoid. Earth and Planetary Science Letters, 170, 487–496.
Planetary Science Letters, 182, 157–169.

Received 24 June 2006; revised typescript accepted 9 February 2007.

You might also like