Seton Etal Global Plate Model ESR2012
Seton Etal Global Plate Model ESR2012
Seton Etal Global Plate Model ESR2012
Earth-Science Reviews
journal homepage: www.elsevier.com/locate/earscirev
a r t i c l e i n f o a b s t r a c t
Article history: Global plate motion models provide a spatial and temporal framework for geological data and have been
Received 16 March 2011 effective tools for exploring processes occurring at the earth's surface. However, published models either
Accepted 2 March 2012 have insufficient temporal coverage or fail to treat tectonic plates in a self-consistent manner. They usually
Available online 15 March 2012
consider the motions of selected features attached to tectonic plates, such as continents, but generally do
not explicitly account for the continuous evolution of plate boundaries through time. In order to explore
Keywords:
Plate reconstructions
the coupling between the surface and mantle, plate models are required that extend over at least a few
Plate motion model hundred million years and treat plates as dynamic features with dynamically evolving plate boundaries. We
Panthalassa have constructed a new type of global plate motion model consisting of a set of continuously-closing topological
Laurasia plate polygons with associated plate boundaries and plate velocities since the break-up of the supercontinent
Tethys Pangea. Our model is underpinned by plate motions derived from reconstructing the seafloor-spreading history
Gondwana of the ocean basins and motions of the continents and utilizes a hybrid absolute reference frame, based on a
moving hotspot model for the last 100 Ma, and a true-polar wander corrected paleomagnetic model for 200
to 100 Ma. Detailed regional geological and geophysical observations constrain plate boundary inception or
cessation, and time-dependent geometry. Although our plate model is primarily designed as a reference
model for a new generation of geodynamic studies by providing the surface boundary conditions for the deep
earth, it is also useful for studies in disparate fields when a framework is needed for analyzing and interpreting
spatio-temporal data.
© 2012 Elsevier B.V. All rights reserved.
Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
2. Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
2.1. Absolute reference frames . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
2.2. Relative plate motions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
2.3. Geomagnetic polarity timescales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
2.4. Continuously closed plate polygons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
3. Regional continental and ocean floor reconstructions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
3.1. Atlantic and Arctic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
3.1.1. South Atlantic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
3.1.2. Central Atlantic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
3.1.3. Northern Atlantic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
0012-8252/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.earscirev.2012.03.002
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 213
60˚
EUR NAM
J
30˚
ARA
IND PH
AFR PAC CAR
C
0˚
CAP
SOM NAZ
−30˚ AUS SAM
−60˚ ANT
ANT
0˚ 30˚ 60˚ 90˚ 120˚ 150˚ 180˚ −150˚ −120˚ −90˚ −60˚ −30˚ 0˚
Fig. 1. Global gravity anomalies from satellite altimetry (Sandwell and Smith 2009). Red lines denote present day plate boundaries from the plate boundary set presented in this
study. AFR = Africa, ANT = Antarctica, ARA = Arabia, AUS = Australia, C = Cocos, CAP = Capricorn, CAR = Caribbean, EUR = Eurasia, IND = India, NAM = North America, NAZ =
Nazca, PAC = Pacific, PH = Philippine, SAM = South America, SOM = Somalia. (For interpretation of the references to color in this figure legend, the reader is referred to the web
version of this article.)
the seafloor-spreading history of the ocean basins and motions of the continuously closed plate polygons. The continuously closed plate poly-
continents and built around a hybrid absolute reference frame. In recon- gons were created using GPlates software (www.gplates.org).
structing the ocean floor, we use satellite-derived gravity anomalies
(Sandwell and Smith, 2009) (Fig. 1) and an updated set of magnetic 2.1. Absolute reference frames
anomaly identifications to construct seafloor spreading isochrons for all
the major oceanic plates. We use a combination of public and in-house The anchor for any global plate motion model is an absolute refer-
magnetic anomaly data, which were line leveled and then gridded, to ence frame (i.e. how the plates move relative to a fixed reference
produce global magnetic anomaly grids and compare with our seafloor system, such as the spin axis). A comprehensive discussion of abso-
spreading isochrons (Figs. 2, 3, 5–7, 9, 11, 13, 14). We derive a global lute reference frames and the merits of each can be found in Torsvik
set of finite rotations for relative motions between all the major plates. et al. (2008). Our model uses a hybrid reference frame, which merges
In addition, we restore now-subducted oceanic crust for the major plates a moving Indian/Atlantic hotspot reference frame (O'Neill et al.,
following the methodology in Müller et al. (2008b), by using evidence of 2005) back to 100 Ma with a paleomagnetically-derived true polar
subduction, slab windows and anomalous volcanism from onshore geol- wander corrected reference frame (Steinberger and Torsvik, 2008)
ogy and the rules of plate tectonics. We create a set of dynamically closed back to 200 Ma. This reference frame links to the global plate circuit
plate polygons in one million year time intervals, which evolve from a se- through Africa, as Africa has been surrounded by mid-ocean ridges
ries of dynamically evolving plate boundaries (Figs. 18–28). for at least the last 170 Ma and, according to Torsvik et al. (2008),
In building a topological closed plate polygon network, we have Africa has moved less than 500 km over the past 100 Ma.
deliberately excluded many of the smaller tectonic plates and All the major tectonic plates are linked to Africa via the seafloor
micro-plates in order to produce a self-consistent global dataset for spreading or rifting back to 200 Ma, except the Pacific and associated
the community. However, the method of Gurnis et al. (2012) allows plates, such as the Farallon, Izanagi, Phoenix and Kula. The Pacific
for construction of more detailed topological plate polygon networks. plate can only be linked to the plate circuit for times younger than
The data involved in reproducing our models are being made publicly 83.5 Ma, after the establishment of seafloor spreading between the
available enabling researchers to either use our model as a framework Pacific and West Antarctic plates. Prior to this time we switch to a
in which to build upon for their particular area of expertise, input into fixed Pacific hotspot reference frame for the Pacific plate, using a
geodynamic simulations as surface boundary conditions or to under- combination of Wessel and Kroenke (2008) and Wessel et al.
stand the context of regional tectonics. We hope that this paper and (2006). We assume that the Pacific reference frame is fixed relative
the accompanying data will help those researchers from disparate to other hotspots as we have no reliable model for whether the Pacific
fields critically evaluate plate reconstructions, determine areas in mantle plumes moved relative to each other or relative to the Earth's
need of further analysis, use as a basis to further refine models and spin axis before 83.5 Ma, although some authors have invoked motion
explore the limitations and sources of error inherent in plate motion between some hotspots in the Pacific to account for paleo-latitude esti-
models. mates from paleomagnetic data for the Ontong-Java Plateau (Riisager
et al., 2003).
2. Methodology
2.2. Relative plate motions
There are four main components that comprise our plate motion
model: an absolute reference frame, the relative motions between tec- In building our relative plate motion model, we combine pub-
tonic plates linked via a plate circuit, the geomagnetic polarity timescale lished and new magnetic anomaly identifications (magnetic anomaly
and a collection of plate boundaries that combine to form a network of picks) and their associated rotations to construct a global set of
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 215
a a
NW Africa DS
10˚ Greenland
Equatorial Segment BT
60˚ RR
0˚ Lab Sea
RP
South RT
−10˚ America South
Central Segment Africa CG Eurasia
50˚ North
NF
−20˚ America
BB
P-E KT
−30˚ RG
WR
40˚ Iberia
Parana
Colorado
Southern/Austral Segment MM
−40˚
30˚
AB CentralAtlantic
NWAfrica
−50˚ Patagonia
Falkland Segment 20˚
Caribbean CLIP
b
10˚ NW Africa −80˚ −70˚ −60˚ −50˚ −40˚ −30˚ −20˚ −10˚ 0˚ 10˚
Equatorial Segment BT
0˚
b
DS
Greenland
South
−10˚ America South 60˚ RR
Lab Sea
Central Segment Africa
RP
−20˚ RT
CG Eurasia
Parana BB
KT
−40˚ Colorado
Southern/Austral Segment 40˚ Iberia
AB
MM
−50˚ Patagonia 30˚
Falkland Segment CentralAtlantic
NWAfrica
Scotia Sea 20˚
−60˚ Caribbean CLIP
−80˚ −70˚ −60˚ −50˚ −40˚ −30˚ −20˚ −10˚ 0˚ 10˚ 20˚
10˚
LIP Isochrons
Plate Boundary
product SAM-AFR SAM-MAL EquatorialAtlantic
Extinct Ridge South
Non-
AFR-EANT NAM-NWA 0˚ America
LIP Isochrons
Plate Boundary
−50 −40 −30 −20 −10 0 10 20 30 40 50 product SAM-AFR NAM-EUR NAM-IBR
Extinct Ridge
Magnetic Intensity Anomaly [nTesla] Non-
NAM-POR NAM-NWA NAM-GRN
oceanic Tectonic Boundary GRN-EUR BB & OTH
Fig. 2. (a) Gridded magnetic anomalies for the South Atlantic. Seafloor spreading
isochrons used in this study plotted as thin black lines. Due to poor data coverage,
correlations between the gridded data and isochrons are difficult. AB = Agulhas
Basin, BT = Benue Trough, P-E = Parana Flood Basalts, RG = Rio-Grande Rise, WR = −50 −40 −30 −20 −10 0 10 20 30 40 50
Walvis Ridge. (b) Seafloor spreading isochron map colored by spreading system or Magnetic Intensity Anomaly [nTesla]
plate pair. Map abbreviations are same as a. Legend abbreviations are: AFR = South
Africa, BB = Back-arc Basins, EANT = East Antarctica/Antarctica, MAL = Malvinas,
Fig. 3. (a) Gridded magnetic anomalies for the Central and North Atlantic. Seafloor
NWA = Northwest Africa, OTH = Other spreading systems outside area of interest,
spreading isochrons used in this study plotted as thin black lines. BB = Bay of Biscay,
SAM = South America.
CG = Charlie-Gibbs Fracture Zone, CLIP = Caribbean Large Igneous Province, DS =
Davis Strait, JFZ = Jacksonville Fracture Zone, KT = Kings Trough, MM = Morocco Me-
seta, NF = Newfoundland, RR = Rekyjanes Ridge, RP = Rockall Plateau, RT = Rockall
seafloor spreading isochrons (see Section 3: Regional continental and Trough. (b) Seafloor spreading isochron map colored by spreading system or plate pair.
Map abbreviations are same as a. Legend abbreviations are: AFR = Africa, BB = Back-
ocean floor reconstructions for details). This is largely based on the
arc Basins, EUR = Eurasia, GRN = Greenland, IBR = Iberia, NAM = North America,
global plate model presented in Müller et al. (2008a), which builds NWA = Northwest Africa, OTH = Other spreading systems outside area of interest,
upon the present day seafloor agegrid work of Müller et al. (1997) POR = Porcupine, SAM = South America.
and includes a database consisting of over 70,000 magnetic anomaly
identifications, extinct and active spreading ridge locations and bound-
ary locations defining the transition from continental to oceanic 25y (55.9 Ma), 31y (67.7 Ma), 34y (83.5 Ma), M0 (120.4 Ma), M4
crust. Seafloor spreading isochrons were constructed at Chrons 5o (126.7 Ma), M10 (131.9 Ma), M16 (139.6 Ma), M21 (147.7 Ma), and
(10.9 Ma), 6o (20.1 Ma), 13y (33.1 Ma), 18o (40.1 Ma), 21o (47.9 Ma), M25 (154.3 Ma) with more detailed timesteps during major tectonic
216 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
events. A finer set of seafloor spreading isochrons was drawn in back- The early GPTS for the Cenozoic (Heirtzler et al., 1968) and
arc and marginal basins. Quoted ages use Cande and Kent (1995) for Mesozoic (Larson and Pitman, 1972) have been superseded by a
times after 83.5 Ma and Gradstein et al. (1994) for times prior to range of updated timescales. Cande and Kent (1995) (CK95) devel-
83.5 Ma. The letter “y” stands for young end of chron and “o” for old oped a timescale for the Cenozoic (0–83.5 Ma) based on a model of
end of chron. We verify our isochron interpretation by correlating smoothly varying spreading rates in the South Atlantic (Cande and
with the magnetic lineations in the World Digital Magnetic Anomaly Kent, 1992) with the inclusion of astronomical information for the
Map (WDMAM) (Maus et al., 2007), the Earth Magnetic Anomaly Grid past 5.23 Ma. Gradstein et al. (1994) (G94) presented an integrated
(EMAG2) (Maus et al., 2009) and our own preferred magnetic anomaly geomagnetic and stratigraphic Mesozoic timescale, which is com-
compilation (Fig. 2). EMAG2 includes a compilation of both ship-track monly merged with the CK95 timescale to create a hybrid timescale
and long-wavelength satellite magnetic anomaly data with trend- through to the Mesozoic (e.g. (Müller et al., 2008b)). The GTS2004
gridding based on the Müller et al. (2008a) isochrons in most areas, timescale (Gradstein et al., 2004) recalibrated CK95 using alternative
hence WDMAM and our own compilation are preferred for correlation. tie-points from updated radiometric ages and astronomical tuning for
We constrain fracture zone locations using global gravity from satellite the Cenozoic and updated the Mesozoic timescale using the method-
altimetry (Sandwell and Smith, 1997, 2005) (Fig. 1). The boundary ology of Cande and Kent (1992) and additional radiometric age
between oceanic and continental lithosphere was taken from Müller constraints. The most recent GPTS (Gee and Kent, 2007) is a hybrid
et al. (2008a), except where otherwise stated in the text. model, which uses CK95 for the Cenozoic and CENT94 (Channell,
The computation of finite rotations and construction of seafloor 1995) for the Mesozoic and includes sub-chrons from Lowrie and
spreading isochrons are relatively straightforward for areas where Kent (2004). The choice of GPTS (i.e. the ages assigned to each mag-
both flanks of a spreading system are preserved (e.g. Atlantic, SE netic anomaly chron) has major implications for the timing of geolog-
Indian Ridge, Pacific–Antarctic Ridge), but becomes more problematic ical events and the significance of geological processes. For example,
in other settings. When only one flank of a spreading system is the inferred mid-Cretaceous seafloor spreading pulse (Larson, 1995)
preserved (e.g. Pacific–Farallon, Pacific–Kula, Pacific–Izanagi, Pacific– is apparent if using the CK94G95 timescale but diminished if using
Phoenix), we compute half-stage rotations (stage rotation between GTS2004 due to a ~4 million year difference in the age assigned to
adjacent isochrons on one flank) and double the half-stage angle (i.e. M0 (~ 120 Ma) (Seton et al., 2009) (Table 1).
assume that spreading was symmetrical) to create a full stage rotation, The occurrence of magnetic reversals in the so-called Jurassic
following the methodology of Stock and Molnar (1988). This assump- Quiet Zone is not a widely accepted explanation for magnetic anomalies
tion of spreading symmetry is reasonable as the maximum cumulative of ages 157 Ma and older, which are rather modeled as geomagnetic
spreading asymmetry globally is only 10%, on average (Müller et al., intensity variations (Gee and Kent, 2007). Despite this, geomagnetic
1998b). In instances where crust from both flanks has been subducted, timescales based on detailed magnetic anomalies collected closer to
we rely on the onshore geological record (e.g. mapping of major the seafloor (using a deep towed magnetometer) in regions of high
sutures, terrane boundaries and active and ancient magmatic arcs) to seafloor spreading rates (in the Pacific ocean) suggest the existence of a
help define the locations of paleo-plate boundaries and use inferences range of short reversals spanning from M29 to M40 (Sager et al., 1998)
from younger, preserved crust to estimate earlier spreading directions or M29 to M44 (Tivey et al., 2006) (T06). Dating of Jurassic Quiet Zone
and rates. Where continental terranes have crossed ocean basins we based on the timescale of Sager et al. (1998) has been also attempted
use the implied history of mid-ocean ridge evolution and subduction in the Central Atlantic Ocean by Roeser et al. (2002) and Bird et al. (2007).
to create synthetic ocean floor by constructing isochrons based on We ensure that our data, including magnetic anomaly identifica-
assuming spreading symmetry and ensuring triple junction closure. tions, finite rotations and seafloor spreading isochrons are calibrated
The location of mid-ocean ridges as they intersect continents can be to one timescale. We choose the CK95 geomagnetic reversal timescale
further constrained by tracking slab window formation along continen- for the Cenozoic (to Chron 34y; 0–83.5 Ma), G94 for the Mesozoic
tal margins (Thorkelson, 1996) and their correlation to anomalous (Chrons M0–M33; 120.4–158.1 Ma) and T06 for the Jurassic (Chrons
geochemistry and volcanism (Bradley et al., 1993; Sisson and Pavlis, M34–M44; 160.3–169.7 Ma), as our standard. Our continuously closed
1993; Breitsprecher et al., 2003; Madsen et al., 2006), elevated geother- plate polygons can be combined using either timescale.
mal gradients (Bradley et al., 1993; Thorkelson, 1996; Lewis et al., 2000)
and the eruption of massive sulfides (Haeussler et al., 1995; Rosenbaum 2.4. Continuously closed plate polygons
et al., 2005). We do not use arguments for the location subduction based
on mantle tomography as our model is solely underpinned by surface A network of tectonic plates, bounded by a series of plate bound-
constraints. aries, combines to cover the surface of the Earth. Most plate tectonic
Triple junction closure follows the rules set out in McKenzie and models reconstruct features on the surface of the Earth without
Morgan (1969) where we assume that the ridge axes are perpendic- regard to the plate margins and are created in time intervals that are
ular to the spreading direction, transform faults are purely strike- too sparse for current needs. These models are insufficient for studies
slip features, plates are rigid and spreading is symmetrical. We use that couple motions of the plates to other dynamic earth processes,
the finite difference method to compute spreading along the third for example mantle convection and oceanic and atmospheric circula-
arm of a triple junction. In addition, we assume that ridge–ridge– tion. This prompted Gurnis et al. (2012) to develop a novel methodolo-
ridge triple junctions are stable features, but note that there is gy to create a set of dynamically closed plate polygons back in time. The
evidence that fast seafloor spreading rates cause triple junction insta- continuously closing plate (CCP) methodology works by assigning a
bility and complexities in spreading (Bird and Naar, 1994). different Euler pole for each plate boundary that constitutes a plate
polygon, ensuring that the polygon remains topologically closed as a
function of time (Gurnis et al., 2012). The feature is built into the plate
2.3. Geomagnetic polarity timescales reconstruction software GPlates (Boyden et al., 2011).
We use the CCP method and the base set of plate polygons in
Geomagnetic polarity timescales (GPTS) correlate the reversals of Gurnis et al. (2012) to create a new set of dynamically closed plate
the Earth's geomagnetic field, most often the sequence of magnetic polygons based on the plate motion model presented in this study
anomalies recorded on the ocean floor, to those based on biostratigra- for the last 200 Ma. The plate polygons are built using a series of
phy, cyclostratigraphy (which includes Earth's orbital variations), plate boundaries, the location and timing of which have been deter-
absolute ages from radiometric studies and average spreading rates mined by using present day plate boundaries (Bird, 2003), geological
for interpolation. evidence for locations of island arcs, magmatic arcs, sutures and
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 217
reconstructions (see discussions in Eagles, 2007, Torsvik et al., 2009, system during the Cenozoic led to this region being used as a type ex-
and Moulin et al., 2010). ample for calibrating the geomagnetic reversal timescale (Cande and
Rifting prior to seafloor spreading in the southernmost Atlantic Kent, 1992).
(“Falkland segment”) is believed to have occurred in the early Jurassic Recent models have been developed to refine rifting and minimize
(190 Ma) and involved dextral movement between Patagonia and the misfits in the South Atlantic. Although no model accurately restores
Colorado sub-plate until the early Cretaceous (126.7 Ma) (Torsvik et all continental margins without gaps or overlaps, we find that the
al., 2009) (Fig. 2). Opening propagated northward into the “Southern/ model of Torsvik et al. (2009) agrees well with continental stretching
Austral segment” adjacent to the Colorado sub-plate in the late Jurassic rates and conjugate margin rifting episodes. We therefore implement
(around 150 Ma) based on late Jurassic–early Cretaceous sediment fill the model of Torsvik et al. (2009) for the early rifting phase of the
and activation (Nürnberg and Müller, 1991) and the onset of deforma- South Atlantic, including intra-continental deformation in South
tion for a “fit” reconstruction using spreading rate interpolation (Eagles, America and Africa but adjust their rotations to be consistent with
2007) or early Cretaceous (140 Ma) according to Schettino and Scotese the Gradstein et al. (1994) timescale for the Mesozoic. In the early
(2005). The model of Torsvik et al. (2009) suggests that rifting was Jurassic (190 Ma), we follow a plate boundary between Patagonia and
accommodated between the Colorado and Parana subplates, Colorado South Africa connected to the Permian–Triassic to Jurassic rifting in
and Africa, and Parana and Africa from 150 Ma and was associated the Karoo Basin (Banks et al., 1995; Catuneanu et al., 2005) and along
with dextral strike-slip motion between Patagonia/Colorado subplate the Agulhas-Falkland Fracture Zone to the Panthalassic subduction
and Parana (Nürnberg and Müller, 1991; Torsvik et al., 2009). Further zone to the west. The South Atlantic central rift propagated northward,
north, rifting adjacent to the Parana subplate and south of the Walvis with extension between Colorado, Parana and Africa from 150 Ma.
Ridge/Rio Grande Rise is believed to have occurred by about 130 Ma Rifting reached the African continental interior through the West and
(Nürnberg and Müller, 1991), 132 Ma corresponding to the Parana– Central African Rift Zones, along the Central African Shear Zone at
Etendeka magmatic event peak (Torsvik et al., 2009), 134 Ma based 131.7 Ma, connecting with the West and Central African Rift Zones.
on the presence of Anomaly M10 and the GTS2004 timescale (Moulin These continental rift zones encompass the major hydrocarbon-
et al., 2010) or 135 Ma based on dating of the continent–ocean transi- producing Cretaceous basins of the Central and West African rift system
tion (Bradley, 2008). The oldest magnetic anomaly that has been from East Niger to Sudan. We cease rifting in the interior of Africa at
identified is M4 (~127 Ma) (Nürnberg and Müller, 1991; Torsvik et al., about 85 Ma.
2009) adjacent to Falkland and Parana/Chacos basin. Coincident with We use the model of Nürnberg and Müller (1991) for the seafloor
opening along the South Atlantic rift was the activation of the West spreading record but refine the timing of the onset of seafloor spread-
and Central African Rift systems and the Central African Shear Zone ing to 132 Ma to correspond to the peak of magmatism (Torsvik et al.,
(Binks and Fairhead, 1992; Genik, 1992; Guiraud and Maurin, 1992; 2009). In addition, we switch to the updated Cenozoic rotations of
Torsvik et al., 2009). Müller et al. (1999) from Anomaly 34 to the present day. The poles
The “Central” segment of the South Atlantic margin (Fig. 2) is presented in Müller et al. (1999) are similar to those of Shaw and
characterized by widespread Aptian salt basin formation. Rifting con- Cande (1990) but reflect finer scale changes in spreading direction
tinued propagating northward and extended into the African interior, due to the inversion method used for fracture zone interpretation
active in the Benue Trough by at least 118 Ma (Nürnberg and Müller, (Müller et al., 1999). Our seafloor spreading isochrons match well
1991), although earlier extension in the Benue Trough is possible with the magnetic lineations observed in our magnetic anomaly grid
(Torsvik et al., 2009). The onset of seafloor spreading in the “Central” (Fig. 2), although poor data coverage hinders broad scale correlation.
segment is difficult to ascertain because the oceanic crust adjacent to We also incorporate spreading in the Agulhas Basin (southernmost
the margin formed during the Cretaceous Normal Superchron (CNS), South Atlantic) between South America and the Malvinas Plate
however Anomaly M0 has been identified extending to latitude 22°S (LaBrecque and Hayes, 1979; Marks and Stock, 2001) from Anomaly
(Cande et al., 1988; Nürnberg and Müller, 1991; Müller et al., 1999). 34 (83.5 Ma) to Anomaly 30 (~66 Ma) according to the rotations of
Torsvik et al. (2009) used the shape and age of the Aptian salt basins Nürnberg and Müller (1991). The extinct spreading ridge associated
to further refine the opening history in this section of the margin and with this spreading system as well as distinct fracture zone trends are
suggested that seafloor spreading only reached north of the Walvis clearly observed in satellite gravity data (Marks and Stock, 2001)
Ridge–Rio Grande Rise at ~112 Ma, much later than 120.4 Ma sug- (Fig. 1).
gested by previous models.
The “Equatorial” segment of the South Atlantic margin (Fig. 2) was 3.1.2. Central Atlantic
the youngest region of plate break-up. Magnetic anomalies cannot be The Central Atlantic contains the region between North America
interpreted due to equatorial formation of the oceanic crust relative conjugate to Northwest Africa bounded by Pico and Gloria Fracture
to spreading direction. However, Anomaly 33 and fracture zone Zones to the north and the 15° 20′N and Guinean Fracture Zones to
segments are well defined. Seafloor spreading is believed to have the south (Fig. 3). Break-up marked the beginning of Pangea separation
propagated into this area after Anomaly M0 (120.4 Ma) (Nürnberg and involved at least a three-plate system between North America,
and Müller, 1991), ~ 100 Ma (Torsvik et al., 2009), 105 Ma (Moulin Northwest Africa and the Moroccan Meseta (Fig. 3). Rifting was con-
et al., 2010) or 102–96 Ma (Eagles, 2007), corresponding to a subtle trolled by pre-existing structures leading to the formation of a series
bend in the fracture zones in the South Atlantic. Either coincident of rift basins during late Triassic–early Jurassic between North America
or subsequent to the opening of the equatorial segment, the areas and Northwest Africa (Lemoine, 1983; Klitgord and Schouten, 1986),
undergoing continental extension in the African interior ceased but which subsequently filled with salt and became inactive during plate
only after a short-lived compressional phase in the late Cretaceous separation. In addition, transtensional rifting between Northwest Africa
(around 85–80 Ma) observed in folding and faulting across seismic and the Moroccan Meseta formed rift basins along the Atlas rift (Labails
sections (Nürnberg and Müller, 1991; Binks and Fairhead, 1992; et al. 2010). The first stage of Atlas Mountain uplift occurred during the
Schettino and Scotese, 2005). opening of the Central Atlantic (Beauchamp, 1998). Incorporating
The spreading history along the entire length of the South Atlantic motion along the Atlas rift has implications for full-fit reconstructions
from Anomaly 34 (83.5 Ma) onwards is relatively uncomplicated with of the Central Atlantic.
most studies in agreement that largely symmetrical spreading occurred The establishment of seafloor spreading in the Central Atlantic is
after Anomaly 34 to the present day (LaBrecque and Rabinowitz, 1977; debated, with ages ranging from 175 Ma marked by the West African
Shaw and Cande, 1990; Nürnberg and Müller, 1991; Torsvik et al., 2009; Coast Magnetic Anomaly and East Coast Magnetic Anomaly and an
Moulin et al., 2010). The stability and symmetry of this spreading extrapolation of spreading rates (Klitgord and Schouten, 1986;
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 219
120 Ma 120 Ma
40˚
20˚
90 Ma 90 Ma
40˚
20˚
60 Ma 60 Ma
40˚
20˚
30 Ma 30 Ma
40˚
20˚
0 Ma 0 Ma
40˚
20˚
0 20 40 60 80 100 120 140 160 NAM GRN EUR IBR AFR NWA NEA POR
Age of Oceanic Lithosphere (m.yrs) Plate Name
220 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
Müller and Roest, 1992; Müller et al., 1999), 170–171 Ma based on a It includes active and extinct spreading systems, ridge–hotspot inter-
review of global passive margins (Bradley, 2008), diachronous open- actions related to the Iceland plume, volcanic and magma-poor mar-
ing with 200 Ma in the south progressing to 185 Ma in the north gins and microcontinent formation (e.g. Jan Mayen). The Northern
based on dating of post-rift sediment deposition (Withjack et al., Atlantic underwent episodic continental extension in the Permo-
1998) and 200 Ma according to model of Schettino and Turco Triassic, late Jurassic, early and mid Cretaceous, with reactivation
(2009). A recent re-evaluation of the Central Atlantic opening and basin formation largely following pre-existing structures from
(Labails et al., 2010) suggests that the earliest seafloor spreading oc- the closure of the Iapetus Ocean and subsequent Baltica–Laurentia
curred at 190 Ma (maximum at 203 Ma) based on an updated mag- collision (400–450 Ma) (Dore et al., 1999; Silva et al., 2000;
netic anomaly grid and interpretation of salt basins offshore Skogseid et al., 2000; Kimbell et al., 2005). Seafloor spreading propa-
Morocco and North America (Sahabi et al., 2004). In this model, gated from the Central Atlantic starting in the late Cretaceous in six
spreading was initially very slow at half-spreading rates of ~8 mm/ distinct phases: Iberia–Newfoundland, Porcupine–North America,
yr with an increase in spreading rate and direction at 170 Ma to Eurasia–Greenland (conjugate to Rockall), North America–Greenland
~ 17 mm/yr and spreading asymmetry until Anomaly M0 (Labrador Sea), Eurasia–Greenland (Greenland and Norwegian Sea
(120.4 Ma). This is in contrast to previous models (Klitgord and and Jan Mayen), North America–Eurasia (Eurasian Basin, Arctic
Schouten, 1986; Bird et al., 2007) that invoke an early ridge jump at Ocean) (Figs. 3–5).
170 Ma rather than significant spreading asymmetry to account for
increased crustal accretion onto the North American plate. 3.1.3.1. Iberia–Newfoundland. The Iberia–Newfoundland margin is a
Anomalies M25–M0 (~ 154–120 Ma) and 34–30 (~ 84–65 Ma) are type example of a highly extended, magma-poor, rifted continental
well established primarily due to the density of data on the western margin (Boillot et al., 1988; Srivastava et al., 2000; Hopper et al.,
flank (Klitgord and Schouten, 1986; Müller and Roest, 1992; Müller 2004; Peron-Pinvidic et al., 2007) with two main phases of extension.
et al., 1999). The spreading rates in the Central Atlantic in the Cenozo- Extension during the late Triassic to early Jurassic formed large rift
ic are quite slow making identification of magnetic anomalies more basins within the continental lithosphere of both margins (Tucholke
difficult than for the Mesozoic (Klitgord and Schouten, 1986). Anom- and Whitmarsh, 2006) and was followed by a period of quiescence
alies from 25 (~56 Ma) onwards have been identified quite consis- in the early–mid Jurassic marked by subsidence and the accumulation
tently between studies (Klitgord and Schouten, 1986; Müller and of shallow-water carbonates (Tankard and Welsink, 1987). The
Roest, 1992; Müller et al., 1999) with the main difference occurring second phase of deformation, from late Jurassic to early Cretaceous,
between Anomalies 8 and 5 (~ 26–10 Ma) due to finer constraints formed a wide zone of layered basalts, gabbros and serpentinized
on fracture zone trends using the models by Müller and Roest mantle (“transitional” crust) indicative of seafloor spreading and
(1992) and Müller et al. (1999). mantle exhumation (Srivastava et al., 1990; Tucholke and Whitmarsh,
We have implemented the early break-up history of Labails et al. 2006; Peron-Pinvidic et al., 2007; Sibuet et al., 2007).
(2010) to define the Jurassic–early Cretaceous history of the Central The onset and location of normal seafloor spreading are widely
Atlantic as a highly asymmetric, slow spreading system. We initiate debated. The interpretation of low amplitude magnetic anomalies as
the Central Atlantic rift prior to 200 Ma together with a transtensional old as Anomaly M21 (~147 Ma) related to ultraslow seafloor spread-
plate boundary between Northwest Africa and Morocco along the ing within the southern part of the transition zone (Srivastava et al.,
Atlas rift using rotations derived from Labails et al. (2010). The 2000; Sibuet et al., 2007) is the oldest seafloor spreading age assigned
Central Atlantic rift connects to a major transform fault along the to the margin. Other studies have instead suggested younger ages for
Jacksonville Fracture Zone to the south linking with Mesozoic rift basins the onset of seafloor spreading: Anomalies M3–M5 (~ 124–128 Ma)
in the Caribbean (see Section 3.4.1: Caribbean). To the north, the Central based on deep sea drilling and seismic refraction (Whitmarsh and
Atlantic rift extends into the northern Atlantic, where Triassic/Jurassic Miles, 1995; Russell and Whitmarsh, 2003) and late Aptian (~ 112–
rifts are observed (see Section 3.1.3: North Atlantic). Immediately fol- 118 Ma) based on stratigraphic studies (Tucholke et al., 2007).
lowing the initiation of seafloor spreading in the Central Atlantic was Although the earliest timing of seafloor spreading remains controver-
the cessation of transtensional motion along the Atlas rift and the first sial, reconstructions between the Iberia and Newfoundland margin
stage of uplift of the Atlas Mountains (Beauchamp, 1998). from Anomaly M0 (~ 120 Ma) onwards are well established with
We initiate seafloor spreading at 190 Ma (Labails et al. 2010) and changes in spreading rates occurring at Anomaly 25 (~ 56 Ma) coinci-
subsequently use the magnetic anomaly picks from Klitgord and dent with the initiation of spreading further north in the Norwegian–
Schouten (1986) and rotations from Müller et al. (1997) for M25– Greenland Sea (Srivastava and Tapscott, 1986; Srivastava et al., 2000).
M0 (~154–120 Ma). Spreading propagated northward between the Related to the development of the Iberia–Newfoundland margin is
Iberia–Newfoundland margin during Anomaly M20 (~146 Ma) the opening of the Bay of Biscay north of Iberia and the motion of the
(Müller et al., 1997) (Fig. 4). To the south, spreading in the Central Iberia block itself. The Bay of Biscay formed at a ridge–ridge–ridge
Atlantic connected with the Equatorial Atlantic in the late Cretaceous. triple junction (Klitgord and Schouten, 1986) commonly believed
We incorporate the Cenozoic rotations from Müller et al. (1999), to have opened in the late Cretaceous (110–83.5 Ma) according to
which have been updated from those of Müller and Roest (1992) Müller et al. (1997). However, Anomalies M0 to 33 (~120–79 Ma)
and use the isochrons from Müller et al. (2008a). The isochrons have been identified (Sibuet et al., 2004) suggesting that seafloor
match well with the gridded magnetic anomalies (Fig. 3) and fracture spreading initiated in the Bay of Biscay at the same time as an in-
zone identifications from global satellite gravity (Sandwell and Smith, crease in spreading rate and cessation of mantle exhumation along
2009) (Fig. 1). the Iberia–Newfoundland margin (Sibuet et al., 2007). The end of sea-
floor spreading occurred at Anomaly 33 (~79 Ma) (Roest and
3.1.3. Northern Atlantic Srivastava, 1991; Sibuet et al., 2004).
The Northern Atlantic encompasses the area between Newfound- Most models agree that the Iberian continental block was fixed rel-
land–Iberia and the Eurasian Basin in the Arctic Ocean (Figs. 3 and 5). ative to Africa since the start of rifting along the Iberia–Newfoundland
Fig. 4. (a) Agegrid reconstructions of the Central and North Atlantic at 120, 90, 60, 30, 0 Ma highlighting the age–area distribution of oceanic lithosphere at the time of formation and
the extent of continental crust (gray polygons). Plate boundaries from our continuously closing plate polygon dataset are denoted as thick white lines, hotspot locations as yellow
stars, large igneous provinces and flood basalts as brown polygons and coastlines as thin black lines. (b) Reconstructions showing the outlines of the plates in the Central and North
Atlantic for each reconstruction time listed above. Feature descriptions as in panel (a). Abbreviations are: NAM = North American plate, GRN = Greenland plate, EUR = Eurasian
plate, IBR = Iberian plate, AFR = African plate, NWA = Northwest African plate, NEA = Northeast African plate, POR = Porcupine plate. (For interpretation of the references to
color in this figure legend, the reader is referred to the web version of this article.)
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 221
160˚
˚ 140 margin until Anomaly 10 (~28 Ma) (Srivastava and Tapscott, 1986)
180
a ˚
based on geological evidence from the Pyrenees and geophysical data
˚ Eurasia 12 from the Northern Atlantic (Roest and Srivastava, 1991; Sibuet et al.,
60 0˚
−1 2004). The location of the plate boundary is proposed to have been
located north of the Kings Trough from M0 (~120 Ma) to the Eocene
North Slope
(Srivastava et al., 1990), extended along the Kings Trough into the Bay
˚
40
Alaska
10
of Biscay and along the Pyrenees from the Eocene to Anomaly 10
−1
0˚
NR
CR MD
(~28 Ma) (Klitgord and Schouten, 1986; Roest and Srivastava, 1991;
PB
Canada Basin Whitmarsh and Miles, 1995) and was followed by southward ridge
LR
jump along the Azores transform fault and Straits of Gibraltar
−120˚
80˚
MB GR (Klitgord and Schouten, 1986; Roest and Srivastava, 1991).
AR
Eurasian Basin
In our plate kinematic model, we use the boundary between con-
North Barents
Shelf
tinental and oceanic crust interpretation of Todd et al. (1988) for the
−100˚
America
Newfoundland margin and Boillot and Winterer (1988) and
60˚
NS
Srivastava et al. (2000) for the Iberia margin. We take the age given
Eurasia
Svalbard
by Srivastava et al. (2000) for the initiation of ultra-slow seafloor
spreading based on their interpretation of magnetic anomalies back
−8
˚
40
0˚
Alaska
40
0˚
MD
incorporate a southern jump of the plate boundary to the Azores trans-
NR CR
Canada Basin
PB form fault and along the Straits of Gibraltar leading to the capture of
Iberia by the Eurasian plate (Fig. 4).
−120˚
LR
80˚
MB GR
AR
3.1.3.2. Porcupine–North America. The Porcupine Abyssal Plain is
Eurasian Basin
Barents
bounded by the Kings Trough, Labrador Sea and Charlie Gibbs Fracture
Shelf
Zone (Figs. 3 and 4). The existence of the Porcupine Plate as an indepen-
−100˚
60˚
North
America NS
dent plate during the Eocene–Oligocene was first hypothesized by
Svalbard
Eurasia Srivastava and Tapscott (1986) in order to account for overlapping
reconstructed anomalies in the Porcupine Abyssal Plain when using a
single pole of rotation for North Atlantic opening and to explain Eocene
−8
Baffin Bay
40
0˚
Greenland deformation recorded along the north Biscay and Porcupine margins.
MR
The need for a separate Porcupine Plate was challenged by Gerstell
DS
−6 ˚ and Stock (1994) when they computed new rotations for Eurasia–
0˚ KR 20
JM AR North America without overlaps between the magnetic anomalies.
−40 However, these reconstructions were themselves challenged as they
˚ 0˚
could not account for the observed intra-plate deformation recorded
LIP Isochrons both onshore and offshore in the Porcupine Abyssal Plain (Srivastava
Plate Boundary
product EUR-JM2 NAM-EUR EUR-JM1 and Roest, 1996).
Extinct Ridge
NAM-MDR GRN-JM1&2 NAM-NSA
Non- A major phase of rifting occurred from the late Jurassic to early
oceanic Tectonic Boundary GRN-EUR BB & OTH
Cretaceous, marked by the formation of extensional basins along
both margins (Rowley and Lottes, 1988) and the deposition of syn-
rift sediments in the Barremian/late Hauterivian 130–125 Ma (De
−50 −40 −30 −20 −10 0 10 20 30 40 50 Graciansky et al., 1985). Seafloor spreading began by at least the
Magnetic Intensity Anomaly [nTesla] mid–late Albian (110–105 Ma) based on the dating of the sediments
above tholeiitic basalt from DSDP sites 550 and 551 and an Aptian
Fig. 5. (a) Gridded magnetic anomalies for the Arctic. Seafloor spreading isochrons regional unconformity (De Graciansky et al., 1985) and supported
used in this study plotted as thin black lines. AL = Alpha Ridge, AR = Aegir Ridge,
by the interpretation of Anomaly 34 (~84 Ma) seaward of this loca-
CR = Chukchi Ridge, DS = Davis Strait, GR = Gakkel Ridge, JM = Jan Mayen, KR =
Kolbeinsey Ridge, LR = Lomonosov Ridge, MB = Makarov Basin, MD = Mendeleev tion (Srivastava and Tapscott, 1986; Müller and Roest, 1992). Further
Ridge, MR = Mohns Ridge, NR = Northwind Ridge, NS = Nares Strait, PB = Podvodnikov refinement based on magnetic anomalies is not possible as the early
Basin. (b) Seafloor spreading isochron map colored by spreading system or plate pair. part of this crust was formed during the CNS.
Map abbreviations are same as a. Legend abbreviations are: BB = Back-arc Basins, Magnetic anomalies from 34 (~ 84 Ma) are well identified in the
EUR = Eurasia, GRN = Greenland, JAM = Jan Mayen, MDR = Mendeleev, NAM =
North America, NOR = Norway, NSA = North Slope Alaska, OTH = Other spreading sys-
Porcupine Abyssal Plain and initially formed as a continuous spread-
tems outside area of interest. ing ridge to the north and south (i.e. between North America and
Eurasia) (Fig. 4). Magnetic anomalies between 25 and 13 (~56-33 Ma)
222 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
record the motion of the independent Porcupine plate relative to of the Rockall Plateau) in the mid–late Cretaceous, coincident with
Eurasia (Srivastava and Tapscott, 1986; Srivastava and Roest, 1989; rifting in the Porcupine Basin to the south (Fig. 4). The main rift
Müller and Roest, 1992). Spreading in the Porcupine Abyssal Plain was phase then jumped westward between the Rockall Plateau (fixed to
coincident with spreading in the Labrador Sea between Anomalies Greenland) and North America at ~ 85 Ma (Gaina et al., 2002), similar
34–13 (~84–33 Ma). After Anomaly 13 (~33 Ma), the Porcupine plate to previous studies (Rowley and Lottes, 1988). We follow the plate
ceased its independent motion and spreading continued via North boundaries in this area from Srivastava and Tapscott (1986) for the
America–Eurasia motion. earliest part of its history. Rifting progressed to seafloor spreading
We use the rotations of Srivastava and Roest (1989) for the initial by Chron 33o (~ 79 Ma) (Gaina et al., 2002) and propagated into the
rift phase between the Porcupine and North American Plate and in- Labrador Sea (Gaina et al., 2002) (Fig. 4). We follow the plate recon-
corporate the onset of break-up and seafloor spreading at 110 Ma structions of Gaina et al. (2002) whereby spreading initiated between
(Müller et al., 1997), marked by a regional unconformity and dating the Rockall Plateau and Greenland after Chron 25 forming a triple
of sediments at DSDP 550 (De Graciansky et al., 1985). We use our junction between the North American, Greenland and Eurasian plates
preferred rotations from Srivastava and Roest (1989) for the early (Fig. 4). As the pole of rotation describing Eurasia–North America mo-
spreading phase and the initiation of independent motion of the tion accounts for the magnetic anomalies in the area, we do not incor-
Porcupine Plate between Anomalies 25 and 13 (~56–33 Ma) (Fig. 4). porate motion between the Rockall Plateau and Eurasia, as proposed
This results in a small clockwise rotation of Eurasia and counter- by other authors (Srivastava and Roest, 1989; Müller and Roest,
clockwise rotation of Iberia relative to the Porcupine Plate. The cessa- 1992).
tion of independent Porcupine motion coincides with the cessation of Seafloor spreading isochrons were constructed based on the mag-
seafloor spreading in the neighboring Labrador Sea and the establish- netic anomaly identification and finite rotations of Gaina et al. (2002)
ment of a simple two-plate system (North America and Eurasia) to and compared to the several magnetic anomaly datasets (Fig. 3). We
describe the plate motions in the North Atlantic (Fig. 4). From Anomaly find that there is generally good agreement between the gridded
13 (~33 Ma) onwards, we use the rotations of Lawver et al. (1990). A magnetic anomaly data and our seafloor spreading isochrons but
comparison with fracture zone traces and satellite gravity data reveals find interpretation difficult proximal to the spreading axis. This may
a slight mismatch due to the compression inferred from our model be due to the thermal influence of the Iceland hotspot on the mid-
that is supported by the seafloor spreading fabric (Srivastava and ocean ridge together with slow seafloor spreading rates. We find
Roest, 1996). very good agreement between our fracture zone trends and those
expressed in the satellite gravity data (Fig. 1).
3.1.3.3. Rockall–North America/Greenland. The Rockall region in the
North Atlantic encompasses spreading between the Rockall Plateau 3.1.3.4. Labrador Sea and Baffin Bay. The Labrador Sea is located be-
conjugate to North America along its southern arm and conjugate to tween North America and Greenland south of Baffin Bay in the
Greenland along its northern arm (Fig. 3). A failed rift basin in the Canadian Arctic (Fig. 3). Continental stretching in the Labrador
Rockall Trough exists adjacent to the Eurasian margin. Previous au- Sea produced a narrow and symmetrical margin with less than
thors have determined that Rockall behaved as an independent 100 km of extension (Dunbar and Sawyer, 1989) at around 130 Ma
plate throughout part of its history (Srivastava and Roest, 1989; (Umpleby, 1979) based on the dating of pre to early syn-rift sedi-
Müller and Roest, 1992) but recent re-analysis of the magnetic anom- ments. Rifting in the Labrador Sea is believed to have begun only after
alies and satellite gravity data can be explained by Eurasia–North the initiation of seafloor spreading in the Rockall Trough (Srivastava
America and Eurasia–Greenland motion (Gaina et al., 2002). and Tapscott, 1986).
The Rockall Plateau underwent periods of extension in the early The onset of seafloor spreading in the Labrador Sea is quite contro-
Triassic, early and mid-Jurassic and early, mid and late Cretaceous versial. The oldest magnetic anomaly identified in the area is Anomaly
(Knott et al., 1993). The majority of rifting in the Rockall Trough oc- 33 (~ 79 Ma) but spreading is believed to have initiated earlier during
curred in the mid–late Cretaceous, continuing into the Eocene after the CNS around 90–92 Ma (Rowley and Lottes, 1988; Roest and
an earlier Triassic–Jurassic rift phase (Cole and Peachey, 1999). Simul- Srivastava, 1989; Gaina et al., 2002). An analysis of reprocessed seis-
taneous rifting in the Porcupine Abyssal Plain occurred in the Creta- mic data (Chalmers, 1991; Chalmers and Laursen, 1995) suggests
ceous (Srivastava and Tapscott, 1986). Spreading between the that seafloor spreading began much later at Anomaly 27 (~61 Ma)
Rockall Plateau and North America was established at ~ 83 Ma inde- with thin continental crust extending into the region where older
pendent of the Eurasian plate according to the models of Srivastava magnetic anomalies have been interpreted. However, this young age
and Roest (1989) and Müller and Roest (1992) or as part of the is inconsistent with the sedimentary–tectonic history of the basins
Eurasian plate from Anomaly 33 (~79 Ma) based on a reinterpreta- around the Labrador Sea which record post-rift deposition and a
tion of magnetic anomalies and fracture zone locations from satellite phase of thermal subsidence around 100–62 Ma and fault block rota-
gravity data (Gaina et al., 2002) or 83 Ma according to Cole and tion between 80 and 63 Ma. Other estimates for the onset of seafloor
Peachey (1999). Spreading propagated to the northwest into the spreading come from an analysis of global passive margins (Bradley,
Labrador Sea (Srivastava and Tapscott, 1986; Rowley and Lottes, 2008), invoking an age of between 109 Ma and 68 Ma for the initia-
1988; Müller and Roest, 1992; Gaina et al., 2002). tion of spreading.
The establishment of a three-plate system between North America, An interpretation of seafloor spreading anomalies by Roest and
Eurasia/Rockall and Greenland occurred after Anomaly 25 (~56 Ma) Srivastava (1989) produced similar results to Srivastava and
(Srivastava and Tapscott, 1986; Rowley and Lottes, 1988; Gaina et Tapscott (1986) except for a re-identification of Anomaly 25
al., 2002). After the cessation of spreading in the Labrador Sea, the (~56 Ma), which yielded a more symmetrical spreading system im-
system reorganized into a two-plate configuration with spreading be- plying a significant change in spreading direction in the Labrador
tween Rockall/Eurasia and Greenland along the Reykjanes Ridge Sea. The change in spreading direction was linked to the initiation
(Srivastava and Tapscott, 1986) after Anomaly 13 (~33 Ma) to the of the Greenland–Eurasia plate boundary and a change in spreading
present day (Fig. 3). direction experienced in the Central and South Atlantic (Rowley and
In constructing our model for spreading in the Rockall region, we Lottes, 1988). Spreading is believed to have continued to Chron 7
separate the margin into two segments: Rockall Plateau/Eurasia (~25 Ma) (Rowley and Lottes, 1988) or just after Chron 13 (~33 Ma)
relative to North America and Rockall Plateau/Eurasia relative to (Roest and Srivastava, 1989; Gaina et al., 2002).
Greenland. Preceding the opening of the ocean basin between Rockall Northward propagation of the Labrador Sea rift into Baffin Bay
and North America, rifting occurred in the Rockall Trough (landward through the Davis Strait (Fig. 3) has been dated to the late Aptian–
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 223
early Cenomanian (110–100 Ma) by the deposition of fluvial sedi- was still fixed to Greenland) from the MØre and VØring Basin mar-
ments during active rifting and occurred at least 20 Ma after the initi- gins. Spreading along the extinct Aegir Ridge formed magnetic linea-
ation of rifting in the Labrador Sea. Although there are no identifiable tions (fan-shaped from Chron 21) in the Norway Basin until about
magnetic anomalies in Baffin Bay, seismic refraction profiles indicate Anomaly 13 (33–30 Ma) when the spreading ridge jumped west-
that the area is floored by oceanic crust (Chalmers and Pulvertaft, ward, likely as a result of ridge–hotspot interactions and initiated
2001) and is predicted by the Labrador Sea opening model of Roest spreading along the Kolbeinsey Ridge (Gaina et al., 2009). This is in
and Srivastava (1989). The cessation of seafloor spreading in Baffin contrast to a model of simultaneous spreading east and west of Jan
Bay may have been coincident with the termination of spreading in Mayen at Anomaly 13 (~33 Ma), initiation of spreading along the
the Labrador Sea. Kolbeinsey Ridge at Anomaly 7 (~25 Ma) and cessation of spreading
For the Labrador Sea and Baffin Bay, we use a set of rotations that in the Norway basin at Anomaly 7 (~25 Ma) (Talwani and Eldholm,
are based on the model presented in Roest and Srivastava (1989) and 1977; Nunns, 1983). Using new marine geophysical data, Gaina et
Gaina et al. (2002). We model continental extension starting at al. (2009) suggest further complications in the rifting and spreading
135 Ma by extrapolation to match the Mesozoic basins on the North history of the Jan Mayen microcontinent and Faeroe Islands with
American and conjugate Greenland margin. We invoke seafloor numerous triple junctions and ridge propagators leading to significant
spreading at Chron 33 (~79 Ma) and incorporate a major change in continental stretching and the formation of rift-related basins. The
spreading direction between Chrons 31–25 (68–56 Ma), which was Mohns Ridge was connected to the Aegir Ridge from the initiation of
subsequently followed by oblique spreading and eventually cessation spreading at ~55–56 Ma until 30 Ma and the cessation of spreading in
of spreading after Anomaly 13 (33 Ma) (Roest and Srivastava, 1989; the Norway Basin. After the seaward ridge jump, the Mohns Ridge linked
Gaina et al., 2002) (Fig. 4). The extinct ridge matches well with a neg- to the Kolbeinsey Ridge defining the boundary between Greenland and
ative gravity anomaly observed in the satellite gravity data (Sandwell Eurasia.
and Smith, 2009). We infer that the spreading axis in the Labrador Sea We use a combination of magnetic anomaly picks and rotations
and Baffin Bay was joined across the Davis Strait via left-lateral trans- from Gaina et al. (2002) and Gaina et al. (2009) to reconstruct the en-
form faults (Rowley and Lottes, 1988; Roest and Srivastava, 1989) tire Greenland–Eurasia margin. We do not incorporate the complex
from 63 Ma. We model the cessation of spreading in Baffin Bay to spreading (triple junctions and ridge propagators) around the Jan
be coincident with the Labrador Sea at 33 Ma (Fig. 4). Mayen microcontinent implied by the model of Gaina et al. (2009),
We use the magnetic anomaly identifications of Gaina et al. (2002) but envisage that these will be incorporated in a further release. In
to construct seafloor spreading isochrons in the Labrador Sea. The our model, spreading initiates along the entire Greenland–Eurasia
magnetic lineations in this area are not well resolved (Fig. 3) and margin at 56 Ma, initially connecting up to the spreading in the
may be due to a combination of high sedimentation rates, spreading Eurasian Basin to the north and the Greenland–Eurasia–North America
obliquity and data resolution. However, a continuation of magnetic triple junction in the south (Fig. 5). At 33 Ma, spreading between North
lineations from the Rockall segment into the southern Labrador Sea America and Greenland in the Labrador Sea ceased fusing the two
(i.e. the expression of the triple junction) is clearly observed. Al- plates together, shutting down the Greenland–Eurasia–North America
though we agree that oceanic crust floors Baffin Bay, no magnetic lin- triple junction and leading to a change in spreading rate and direction
eations can be resolved from the global gridded magnetic anomaly along the Greenland–Eurasia spreading system. The Jan Mayen micro-
data (Figs. 3 and 5). continent rifted off the Norwegian margin at 56 Ma forming the fan-
shaped Norway Basin along the Aegir Ridge between 56 and 33–30 Ma
3.1.3.5. Greenland–Eurasia and Jan Mayen microcontinent. The separa- (Fig. 5). The Aegir Ridge connected to the Mohns Ridge in the north
tion of Greenland and Eurasia is occurring along the Reykjanes and Reykjanes Ridge in the south via a series of transform faults. Spread-
Ridge adjacent to the Rockall Plateau, through Iceland and along the ing then jumped to the Kolbeinsey Ridge at 30 Ma, connecting with the
Kolbeinsey and Mohns Ridge in the Norwegian and Greenland Seas Mohns Ridge further north and forming the present day plate configura-
(Figs. 3 and 5). The margin has undergone several rift phases since tion (Fig. 5). A comparison between our resultant seafloor spreading
the Triassic primarily during the mid Jurassic–early Cretaceous and isochrons and the magnetic anomaly grids reveals that our trends
late Cretaceous–early Cenozoic (Brekke, 2000). The late Jurassic– match quite well with the magnetic lineations from the gridded dataset.
early Cretaceous rift phase created most of the basin structures
in the hydrocarbon-bearing MØre and VØring Basins, offshore 3.1.3.6. Lomonosov Ridge–Eurasia (Eurasian Basin). The Eurasian Basin
Norway (Skogseid et al., 2000). The final rift phase at the Campanian– is the youngest ocean basin within the Arctic Ocean and was formed
Maastrichtian boundary (~70 Ma) (Skogseid et al., 2000) was followed by spreading between the Lomonosov Ridge and the Barents Shelf
by volcanism (mid Paleocene to early Eocene) and finally to break-up along the Gakkel and Nansen Ridges (Fig. 5). The continental nature
and volcanism prior to Chron 25 (~56 Ma). of the Lomonosov Ridge has been confirmed through seismic reflec-
Traditionally, spreading between Greenland and Eurasia is mod- tion imaging (Jokat et al., 1992) and ACEX drilling (Moran et al.,
eled as a two-plate system with seafloor spreading initiating around 2006). The broad scale early rift phase mimics those of the North At-
55–56 Ma, near the Paleocene–Eocene boundary (Talwani and lantic margin but is less well constrained due to the remoteness of the
Eldholm, 1977; Srivastava and Tapscott, 1986; Rowley and Lottes, region, data quality and persistent ice-coverage. Although the Barents
1988; Peron-Pinvidic et al., 2007). An updated interpretation includ- Shelf is agreed to have formed part of the Eurasian margin, there is
ing new geophysical data suggests that the system underwent several debate in the literature as to whether the Lomonosov Ridge has
plate boundary changes since the inception of seafloor spreading been fixed to the North American plate since at least 80 Ma
around Anomaly 25 (~56 Ma) (Gaina et al., 2009). Fracture zone (Srivastava and Tapscott, 1986; Rowley and Lottes, 1988) or whether
trends mark changes in spreading direction at Chron 21 (~47 Ma) it operated as an independent plate until at least Anomaly 13
and Chron 18 (~ 40 Ma) (Gaina et al., 2009). A major reorganization (~33 Ma) (Jackson and Gunnarsson, 1990; Brozena et al., 2003). The
of the system occurred at Anomaly 13 (~33 Ma) with relative motion lack of evidence for contemporaneous seafloor spreading in other
between Greenland and Eurasia migrating from NW–SE to NE–SW, parts of the Arctic Ocean and the good fit of the magnetic anomalies
leading to the cessation of spreading in the Labrador Sea, the amal- in the Eurasian Basin are cited as reasons for the Lomonosov Ridge
gamation of Greenland with North America and the cessation of being part of the North American Plate. However, a recent compila-
spreading in the Norway Basin. tion of marine geophysical data identified a feature that resembles
Spreading in the Norway Basin (part of the Norwegian Sea) was an extinct spreading ridge near the Lomonosov Ridge, which possibly
initiated at 56 Ma isolating the Jan Mayen microcontinent (which connected spreading in the Eurasian Basin with spreading in the
224 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
Labrador Sea (Brozena et al., 2003), thus requiring independent motion possible. An analysis of rift-related structures and stratigraphy (Grantz
of the Lomonosov Ridge. et al., 1998) reveals that the opening of the Canada Basin could have
The last rifting phase (late Cretaceous) led to break-up and sea- occurred as early as the late Jurassic–earliest Cretaceous. Less well-
floor spreading at 68 Ma (Rowley and Lottes, 1988) or around Anom- accepted models exist to explain the opening of the Canada Basin
aly 25 (~56 Ma) (Srivastava, 1985; Gaina et al., 2002) in the south such as a non-rotational, step-wise late Jurassic–late Cretaceous open-
around Svalbard and at 50 Ma in the Laptev Sea (Rowley and Lottes, ing model (Lane, 1997) and a model involving trapped crust from
1988). There appears to be a consensus in early studies that the oldest Kula–Pacific spreading (Churkin and Trexler, 1980).
magnetic anomaly that can be confidently identified is Anomaly Following the opening of the Canada Basin, Alvey et al. (2008)
25–24 (~ 56–53 Ma) (Srivastava, 1985; Srivastava and Tapscott, postulated that the Mendeleev and Alpha Ridges in the central Arctic
1986; Rowley and Lottes, 1988; Gaina et al., 2002), yet there is formed either: (1) During continental rifting from the Canadian margin
space landward of Anomalies 25–24 (~56–53 Ma) to suggest that sea- in the late Jurassic trapping Jurassic ocean floor in the Marakov/
floor spreading initiated earlier. The early spreading phase was the Podvodnikov Basin (Grantz et al., 1998); (2) During continental rifting
result of transtensional opening (Rowley and Lottes, 1988) producing from the Lomonosov Ridge forming the Marakov/Podvodnikov Basins
slow seafloor spreading rates, strike-slip motion between Svalbard during the late Cretaceous–mid Eocene (Alvey et al., 2008); (3) A hy-
and Greenland (Srivastava and Tapscott, 1986) and displacement brid model which includes an element of Jurassic ocean floor in the
along the Nares Strait (Srivastava, 1985). After Chron 13 (33 Ma), Podvodnikov Basin and a Cenozoic Marakov Basin (Alvey et al., 2008)
true seafloor spreading was established coincident with the major or (4) The ridges formed purely via LIP emplacement related to the
reorganization of the Greenland–Eurasia system and cessation of Iceland plume in the late Cretaceous (Forsyth et al., 1986; Lawver
Labrador Sea spreading. Currently, the Eurasian Basin is undergoing and Müller, 1994; Lawver et al., 2002; Jokat et al., 2003; Dove et al.,
the slowest observed seafloor spreading rates, with a full rate of 2010) overprinting old oceanic crust. Interpretations suggesting a Ce-
~ 10–13 mm/yr. nozoic age for the Marakov Basin match well with the identification of
We have used the magnetic anomaly picks and finite rotations of Anomalies 34–21 (~84–46 Ma; late Cretaceous–mid Eocene) (Taylor et
Gaina et al. (2002) to describe the opening of the Eurasian Basin al., 1981) as well as crustal thickness estimates (Alvey et al., 2008) in
from Anomaly 24 (~ 53 Ma) to the present day. The rotations used the Marakov Basin, but crustal thickness estimates postulate that the
are the same as for North America–Eurasia. We incorporate the Podvodnikov Basin must be floored by older oceanic floor (Alvey et al.,
plate boundary model of Rowley and Lottes (1988) whereby the 2008). The volcanic nature of the Mendeleev and Alpha Ridges has
Gakkel and Nansen Ridges connect to the Baffin Bay ridge axis been confirmed from recovered basalt samples of late Cretaceous age
through the Nares Strait and Mohns Ridge via a major strike-slip (Jokat et al., 2003), an age slightly younger than the predicted location
fault with minor compression between Greenland and Svalbard of the Iceland plume around 130 Ma (Hauterivian/Berremian) (Lawver
(Fig. 5). In our interpretation, we couple the Lomonosov Ridge with and Müller, 1994). However, this does not preclude a continental nature
North America as the rotations of Gaina et al. (2002) to describe for the Mendeleev and Alpha Ridges. Subsequent to the opening of the
North America–Eurasia motion do not result in overlap of the mag- Marakov/Podvodnikov Basins, the locus of spreading jumped to the Eur-
netic anomalies. The seafloor spreading isochrons we implement are asian Basin at ~56 Ma, forming the youngest piece of ocean floor in the
digitized from Gaina et al. (2002) and match well with the magnetic Arctic domain.
anomaly grid (Fig. 5). We have incorporated a model whereby initial rifting occurred be-
tween the North American and Alaskan margins in the early Jurassic
3.1.4. Arctic Basins (~210–200 Ma) followed by the isolation of the Northwind and
The Arctic Ocean encompasses the Eurasian and Amerasia Basins Chukchi Ridges by the earliest late Cretaceous, triggered by the sub-
(divided into the Canada, Makarov and Podvodnikov Basins) as well duction of the Anyui Ocean. We invoke a simple counterclockwise ro-
as numerous continental blocks such as the Lomonosov, Mendeleev, tational model for the opening of the Canada Basin whereby the
Alpha, Northwind and Chukchi Ridges (Fig. 5). The Cenozoic Eurasian North Slope of Alaska starts to rotate at 145 Ma (latest Jurassic)
Basin (see Section 3.1.3.6: Eurasian Basin) has a distinct spreading with seafloor spreading initiating at 142 Ma (Berriasian), with a
history from the late Jurassic–Cretaceous Amerasia Basin. The early much lower spreading rate in the south due to its proximity to the
Mesozoic evolution of the Arctic region involves the closure of the pole of rotation, creating fan-shaped anomalies. The timing is consis-
South Anyui Basin along the North Siberian subduction zone, marked tent with paleomagnetic data from Alaska but is inconsistent with
by the South Anyui suture (Nokleberg et al., 2001; Sokolov et al., previous magnetic anomaly interpretations (Taylor et al., 1981;
2002; Kuzmichev, 2009). This resulted in pre-breakup rifting in the Srivastava and Tapscott, 1986). Cessation of spreading in the Canada
earliest Jurassic, forming the Dinkum and Banks graben systems in Basin and rotation of North Slope occurred at 118 Ma, coincident
Alaska and North America, respectively and the subsequent isolation with a change in the southern North Slope margin from largely
of the Northwind and Chukchi Ridge by the earliest late Cretaceous strike-slip to convergence due to a change in spreading direction in
(Grantz et al., 1998). Panthalassa. We use the finite rotations and seafloor spreading iso-
Rifting and opening of the Canada Basin is believed to have resulted chrons from Model 1 presented in Alvey et al. (2008), however we
from anticlockwise rotation of the North Slope Alaska–Chukotka Block modify the isochrons to extend the interpretation of the Canada
away from the Canadian Arctic Islands, with a possible early strike- Basin over the Alpha Ridge and into the Marakov Basin. The isochrons
slip component, sometime from the late Jurassic to mid Cretaceous are not constrained by magnetic anomaly identifications but rather
(Carey, 1955; Rowley and Lottes, 1988; Grantz et al., 1998; Alvey et are a synthetic interpretation of the timing and orientation of spread-
al., 2008). Although the rotation model is supported by paleomagnetic ing based on the rotation of the North Slope of Alaska. Hence, we do
data (Halgedahl and Jarrard, 1987), the fan-shaped nature of the mag- not expect an exact correlation with the magnetic anomaly grid.
netic lineations (Taylor et al., 1981) and crustal thickness mapping The preferred model presented in Alvey et al. (2008) based on
(Alvey et al., 2008), the exact timing of the rotation of Alaska and forma- crustal thickness estimates, invokes Cenozoic spreading in the Mara-
tion of the Canada Basin is debated. The dating of the magnetic anoma- kov Basin. We do not incorporate a younger Marakov Basin as this
lies in the Canada Basin is difficult due to extensive volcanic would require either a short-lived subduction zone along either the
overprinting, low amplitude signature of the magnetic anomalies Lomonosov or Mendeleev Ridge during the opening of this basin for
and high sedimentation rates. Anomalies M25–M11 (~154–132 Ma) which there is no geological evidence. Instead, we suggest that the
have been tentatively identified (Taylor et al., 1981; Srivastava Alpha and Mendeleev Ridges are predominately LIP-related features
and Tapscott, 1986), but other magnetic anomaly interpretations are associated with the Iceland plume that overprinted the Canada
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 225
Basin in the early Cretaceous (Lawver and Müller, 1994) and not part Engebretson et al. (1985) presented a quantitative plate kinematic
of a rifted Cenozoic continental margin. In our model the Makarov model of the seafloor spreading record focused on the northern
and parts of the Podvodnikov Basin form the northern extent of the Pacific basin for the past 180 Ma and is currently the most compre-
Canada Basin. We do agree with Alvey et al. (2008) that there may hensive and often cited study on Pacific plate reconstructions. This
be a trapped piece of Jurassic ocean floor from the Anyui Basin in study enabled subsequent authors to place their regional tectonic re-
the Podvodnikov Basin, which would explain the anomalous crustal constructions and geological observations into a Pacific-wide tectonic
thickness and would provide a mechanism for the Mendeleev Ridge framework. The model of Engebretson et al. (1985) is based on an ab-
having some continental affinities as continental material may have solute reference frame using fixed Atlantic and fixed Pacific hotspots
been isolated during Jurassic rifting. (Morgan, 1972) with relative plate motions for the Pacific, Farallon,
Izanagi, Kula and Phoenix plates determined by computing the dis-
placements of each plate relative to the absolute reference frame
3.2. Pacific Ocean and Panthalassa rather than via plate circuit closure as is commonly used. Since the
publication of Engebretson et al. (1985), additional data acquisition,
Present day seafloor spreading in the Pacific basin involves nine oce- updated interpretations and more accurate magnetic anomaly time-
anic plates: the Pacific, Antarctic, Nazca, Cocos and Juan De Fuca plates scales have been published, providing improved constraints on
and the smaller Rivera, Galapagos, Easter and Juan Fernandez micro- the Izanagi–Pacific, Phoenix–Pacific, Farallon–Phoenix and Pacific–
plates along the East Pacific Rise (Bird, 2003) (Fig. 1). Additionally, the Antarctic ridges.
Pacific basin seafloor spreading record preserves clear evidence that The Pacific triangle is an area of the western Pacific where three
several now extinct plates (e.g. Farallon, Phoenix, Izanagi, Kula, Aluk Mesozoic magnetic lineation sets (Japanese, Hawaiian and Phoenix
and Bauer plates) existed within the Pacific and proto-Pacific basin lineations) intersect (Fig. 6), recording the birth of the Pacific plate
(Panthalassa) since at least the Jurassic/Cretaceous. In addition, the on- from three “parents”: the Farallon, Izanagi and Phoenix plates. The
shore geological record from the Pacific margins provides evidence for evolution of the three parent plates has influenced the development
the opening and closure of several marginal basins, particularly along of subsequent seafloor spreading systems in the Pacific. The north-
the western North American margin. western (Japanese) lineations represent spreading between the
Previous plate tectonic models of the Pacific have largely focused Pacific and Izanagi plates and young towards the west–northwest,
on identifying magnetic lineations and deriving relative plate motions the easternmost (Hawaiian) lineations represent spreading between
between presently active plates where both sides of the spreading the Pacific and Farallon plates and young towards the east and the
ridge are preserved (e.g. Juan De Fuca–Pacific spreading (Atwater, southernmost (Phoenix) lineations represent spreading between the
1970, 1990; Engebretson et al., 1985; Caress et al., 1988; Stock and Pacific and Phoenix plates and young towards the south (Atwater,
Molnar, 1988; Wilson, 1988; Atwater and Severinghaus, 1990), 1990; Nakanishi et al., 1992) (Fig. 6). These three plates radiated
Pacific–Antarctic spreading (Stock and Molnar, 1987; Cande et al., out from the emerging Pacific plate during the Mesozoic and existed
1998; Larter et al., 2002), the east Pacific Rise (Cande et al., 1982; prior to the establishment of the Pacific plate in a simple ridge–
Tebbens and Cande, 1997) and Cocos and Nazca spreading (Wilson, ridge–ridge configuration. We will present an assessment of the
1996)). Other plate tectonic models have focused on identifying mag- Pacific and Panthalassa by describing each parent plate with their
netic lineations in the older parts of the Pacific, particularly the north associated children.
and western Pacific, where conjugate magnetic lineations no longer
exist as they have been subducted (e.g. Kula–Pacific (Rea and Dixon, 3.2.1. Izanagi plate
1983; Engebretson et al., 1985; Lonsdale, 1988; Mammerickx and The M-sequence Japanese magnetic lineation set found in the
Sharman, 1988; Atwater, 1990), Izanagi–Pacific (Larson et al., 1972; westernmost Pacific represents the last preserved fragments of a
Woods and Davies, 1982; Sager and Pringle, 1987; Handschumacher westward-younging Jurassic–Cretaceous spreading system (Fig. 6).
et al., 1988; Sager et al., 1988; Nakanishi et al., 1992; Nakanishi and Early reconstructions of the area linked the Japanese lineation set to
Winterer, 1998), Farallon–Pacific (Atwater, 1970, 1990; Engebretson the younger, Cenozoic seafloor spreading history of the Pacific–Kula
et al., 1985; Caress et al., 1988; Stock and Molnar, 1988; Wilson, 1988; ridge (Larson et al., 1972). To reconcile the geometry of the preserved
Atwater and Severinghaus, 1989), Phoenix–Pacific spreading (Stock NE–SW trending Japanese lineations with the E–W trending Cenozoic
and Molnar, 1987; Cande et al., 1998; Sutherland and Hollis, 2001; lineations formed by Pacific–Kula spreading, Woods and Davies
Larson et al., 2002; Larter et al., 2002; Viso et al., 2005) and the plates (1982) introduced the idea of an independent Izanagi plate, although
related to the break-up of the Ontong Java–Hikurangi–Manihiki Pla- some models still prefer a single Kula plate (Norton, 2007). Due to
teaus (Taylor, 2006)). Beyond this, few studies have attempted to derive progressive subduction since the Mesozoic, the entire crust that
relative plate rotation models of these now vanished plates (e.g. floored the Izanagi plate as well as the portion of the Pacific plate
Engebretson et al., 1985; Stock and Molnar, 1988) to establish a longer recording the death of the Izanagi has been lost, leaving behind only
tectonic history of the Pacific plate where minimal or no information the Mesozoic fragment of the Pacific plate. This complicates recon-
about the seafloor spreading record exists. structions as few present day constraints exist to tie down tectonic
Another common approach to constrain plate tectonic models of parameters for the evolution of the area. Additionally, there are no
the Pacific has been through the interpretation of the onshore geolo- constraints on the history of the Izanagi plate prior to the birth of
gy, in particular examining anomalous volcanism and geochemistry the Pacific plate.
associated with ridge subduction, crustal shortening rates and events, Magnetic anomalies M33–M0 (~158–120 Ma) of the Japanese line-
accretion of exotic terranes, ophiolite emplacement, large-scale crust- ation set have been confidently identified in the northwest Pacific
al deformation and massive sulfide and other subduction related ore- (Sager and Pringle, 1988; Handschumacher et al., 1988; Sager et al.,
deposit formation (e.g. Bradley et al., 1993; Haeussler et al., 1995; 1988; Atwater, 1989; Nakanishi et al., 1992; Nakanishi and Winterer,
Madsen et al., 2006; Sun et al., 2007). This information is sometimes 1998). A recent deep-tow magnetometer survey over the Pigafetta
translated into a schematic representation of past plate configura- Basin in the vicinity of ODP drill site 801C revealed a low amplitude
tions based purely on the onshore record but these plate reconstruc- magnetic anomaly sequence extending to M44 (~170 Ma), within the
tion schematics are often only snapshots in time rather than evolving Jurassic Quiet Zone (Tivey et al., 2006) with Anomaly M42 (~168 Ma)
and are not quantitatively derived through the seafloor spreading re- corresponding to the location of ODP drill site 801C (Tominaga et al.,
cord. Nevertheless, they are helpful in developing conceptual models 2008). Previous interpretations infer the oldest crust in the Pacific to
for the evolution of now vanished ocean crust. be 175 Ma (Engebretson et al., 1985; Müller et al., 1997) based on
226 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
a b
21 North America 21 North America
21
21 Okhotsk
Okhotsk 50˚ Amur 25
50˚ Amur 25
31
31 M0 33
M4
40˚ M0
M0 33 M1
0
4
M
M4 M1
6
HR
40˚ M0
SR
10
0 M
M1
21
30˚
M
4 16
M
25
M
M
KF
6 21
M1
Z
SR HR M Pacific
10
29
M 25
M
M
21
20˚ MPM
30˚
M
Philippine 29
16 M
25
M 37
M
37
M
KF
M
Z 21
M Pacific M37
10˚
29
25 M29
M
M M25
20˚ ODP801
MPM Caroline M21
Philippine 29 M16 M10
M 0˚ M4
37 OJP
37
M
M0
M
Ellice Basin
M37 Australia
10˚ −10˚
M29 130˚ 140˚ 150˚ 160˚ 170˚ 180˚ −170˚
M25
Caroline M21 LIP Plate Boundary
Isochrons
product PAC-IZA PAC-PHX
M16 M10 Extinct Ridge PAC-FAR PAC-MAN
Non-
0˚ OJP M4 oceanic Tectonic Boundary PAC-KUL BB & OTH
M0
Ellice Basin
Australia
−10˚ −50 −40 −30 −20 −10 0 10 20 30 40 50
130˚ 140˚ 150˚ 160˚ 170˚ 180˚ −170˚ Magnetic Intensity Anomaly [nTesla]
Fig. 6. (a) Gridded magnetic anomalies for the Western Pacific, based on isotropic gridding of a combination of public domain and in-house data. Seafloor spreading isochrons used
in this study plotted as thin black lines. Numbers correspond to magnetic anomaly chron. HR = Hess Rise, OJP = Ontong Java Plateau, SR = Shatsky Rise. (b) Seafloor spreading
isochron map colored by spreading system or plate pair. Legend abbreviations are: BB = Back-arc Basins, FAR = Farallon, IZA = Izanagi, KUL = Kula, MAN = Manihiki, OTH =
Other spreading systems outside area of interest, PAC = Pacific, PHX = Phoenix.
interpolation to the center of the Pacific triangle, but this age ap- M20 (~145 Ma) lineations (Fig. 6). This would suggest no measured
pears to be inconsistent with the recent dating of magnetic anoma- change in spreading direction between at least 145–120 Ma. The ocean-
lies and the dating from ODP site 801C, which is located ~ 750 km ic crust to the north of M0 (~120 Ma) is inferred to have formed during
from the inferred center of the Pacific triangle. After the initiation the CNS and represents the youngest preserved oceanic lithosphere as-
of spreading between the Pacific and Izanagi plates, the ridge under- sociated with Izanagi–Pacific spreading.
went some instability with one or more proposed ridge jumps pos- Previous interpretations have tied the cessation of spreading
tulated to explain the anomalously large distance between the between the Pacific and Izanagi plates to the onset of spreading
adjacent isochrons along a spreading corridor between M33 and between the Kula and Pacific plates (Engebretson et al., 1985), some-
29 (~ 158–156 Ma) (Sager et al., 1998). Analysis of the magnetic time between 83.5 and 70 Ma (Lonsdale, 1988; Atwater, 1989) (see
anomalies and seafloor fabric flanking this proposed ridge jump Section 3.2.2.1: Kula plate). In these models, the orientation of the
has not found an abandoned spreading center. Spreading continued Izanagi–Pacific ridge is depicted as a side-stepping E–W oriented
with relatively high seafloor spreading rates between M29 and 25 ridge perpendicular to the East Asian margin. As the oldest discern-
(~ 156-154 Ma) (Nakanishi et al., 1992) before decreasing to aver- able Japanese magnetic lineation is oriented NE–SW, an E–W oriented
age rates until M21 (~ 147 Ma). mid-ocean ridge requires a major change in spreading direction post-
The fracture zone pattern observed in the satellite gravity data and M0 (~120 Ma). However, there are no fracture zones present in the
mapped via ship track data (Sager et al., 1988; Nakanishi and post-Mesozoic crust of the NW Pacific to suggest a major change in
Winterer, 1998; Sager et al., 1998) indicates a large 24° clockwise ro- spreading direction during the CNS (Fig. 1).
tation of the Izanagi plate relative to the Pacific at M21 (~147 Ma) An alternative approach to constrain the orientation and cessation of
(Sager et al., 1999), particularly evident along the Kashima Fracture the Izanagi–Pacific ridge is through an analysis of the onshore geological
Zone near the Izu-Bonin–Mariana trench (Fig. 6). The change in record in east Asia together with the preserved seafloor spreading
spreading direction from NW–SE to NNW–SSE coincides with the record in the NW Pacific (Whittaker et al., 2007; Seton et al., in
eruption of the Shatsky Rise at the Izanagi–Farallon–Pacific triple preparation). The younging northwestward sequence of magnetic line-
junction (Nakanishi et al., 1999; Sager et al., 1999) followed by the ations and the presence of Indian-type mantle geochemical signatures
progressive reorganization and migration of the triple junction center in various volcanic arcs of the northwest Pacific (Straub et al., 2009) in-
for a period of about 2 million years. The period between Anomalies dicate a ridge subducted under east Asia at some time in the past.
M21 and 20 (~147–145 Ma) also corresponds to changes in spreading Whittaker et al. (2007) assumed no change in spreading direction of
rate and direction in the Pacific, Atlantic and Indian Oceans (Sager et the Pacific–Izanagi from M0 (~120 Ma) onwards as there is no evidence
al., 1988; Nakanishi et al., 1999). The youngest identified Japanese for a major change in spreading direction post-M0 (~120 Ma) resulting
lineation corresponds to M0 (~120 Ma) (Nakanishi et al., 1999; Sager in the mid ocean ridge intersecting the east Asian margin in a sub-
et al., 1999; Tominaga and Sager, 2010) trending similar to the post- parallel fashion. The timing for the intersection of the ridge with the
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 227
margin forming a slab window can be constrained through a number of between M21 (~147 Ma) and M16 (~138 Ma) (Fig. 6). Due to the
geological observations from Japan and Korea. The geology in southern complexity of the triple junction solutions and the lack of preserved
and central Japan records a pulse of volcanism and anomalous heatflow data between the Izanagi and Farallon plates, we incorporate a simple
measurements (Agar et al., 1989; DiTullio, 1993; Sakaguchi, 1996; model whereby the Pacific–Izanagi–Farallon triple junction remains
Lewis and Byrne, 2001) indicative of the presence of a slab window in in a ridge–ridge–ridge configuration during its entire history. As the
the late Cretaceous-early Cenozoic. The cessation of granitic plutonism instability of this triple junction is believed to have existed for only
in Korea suggests that subduction was terminated along east Asia 2 million years (Sager et al., 1999), we believe that our assumption
around 60–50 Ma (Sagong et al., 2005). In addition, seismic tomogra- is reasonable and follows the broad scale development of the area.
phy profiles across east Asia reveal a break in the continuity of slab The fracture zones in the westernmost Pacific do not show a major
material in the mid-mantle (Seton et al., in preparation) possibility indi- change in trend after M20 (~146 Ma) (Fig. 6). No discernable fracture
cating the subduction of a mid-ocean ridge and slab break-off event. zone trends after M0 indicate the direction of motion during the CNS
Based on this model, the cessation of spreading between the Izanagi hence we assume that no change in the direction of motion occurred
and Pacific plates (i.e. the death of the Izanagi plate) occurred around from M20 to the CNS and use a fixed stage rotation pole for this entire
55–50 Ma followed by the complete subduction of the Izanagi plate period. As much of the evidence for the late Cretaceous–early Cenozoic
along the East Asian margin by 40 Ma. In this model, the cessation of history of the Izanagi plate has been lost due to subduction along the
spreading between the Izanagi and Pacific plate is not correlated with east Asian margin, we assume no major change in spreading rate, direc-
the initiation of spreading in the Kula plate, as suggested by previous tion and accretion from M0 (last dated anomaly, ~120 Ma) to the cessa-
studies. tion of spreading along the Izanagi–Pacific ridge.
We model the Mesozoic–early Cenozoic evolution of the Izanagi Finite rotations were computed for Izanagi–Pacific spreading
plate using constraints still preserved on the Pacific plate. We define using the half-stage pole method and assuming spreading symmetry
the onset of spreading between the Pacific and Izanagi plates to and rely heavily on fracture zone traces for direction of motion. For
190 Ma, 15–20 million years earlier than previous interpretations. We younger times when no preserved crust exists, we assume an inter-
base our age estimation, which is a maximum age, on the following: mediate full spreading rate of ~ 80 mm/yr (similar to the spreading
rate in the late Cretaceous), spreading symmetry and a consistent
(1) The location of the oldest identified magnetic anomaly, M44
spreading direction to model the position of the mid-ocean ridge.
(~170 Ma) (Tivey et al., 2006) is over 750 km from the inferred
We find that this results in the Pacific–Izanagi ridge intersecting the
center of the Pacific triangle
east Asian margin around 55–50 Ma in a sub-parallel orientation
(2) ODP site 801C, which lies within M42 (~ 168 Ma) is consistent
and is consistent with geological and seismic tomography observa-
with the dating of microfossils overlying pillow basalts
tions, as explained in Seton et al. (in preparation). Our model suggests
(Lancelot et al., 1990; Tivey et al., 2006)
that spreading continued along the Pacific–Izanagi ridge after the es-
(3) An extrapolation of intermediate seafloor spreading rates
tablishment of the Kula–Pacific ridge to the east, contrary to most
(~30–40 mm/yr) from the location of M44 to the center of
previous models. The preserved seafloor spreading record in the re-
the Pacific triangle suggests an approximate age to be closer
gions adjacent to the Pacific–Izanagi ridge preserves no evidence to
to around 190 Ma.
suggest a readjustment of the plate driving forces due to the merging
(4) A younger age for the initiation of seafloor spreading between
of two major plates (i.e. the death of the Izanagi plate) prior to 55 Ma.
the Izanagi–Pacific, Farallon–Pacific and Phoenix–Pacific would
Instead, we find that spreading between the Kula and Pacific plates
require anomalously high spreading rates or substantial
underwent a major change in spreading rate and direction at Anom-
spreading asymmetry. This cannot be discounted as Tominaga
aly 24 (~55–53 Ma), which resulted in a dramatic doubling of the
et al. (2008) suggest a rapid spreading rate of ~75 mm/yr.
spreading rate of the Kula plate and a counter-clockwise change in
Therefore, we believe an age of 190 Ma for the birth of the
spreading direction from largely N–S to NW–SE. Our model is in
Pacific plate is a maximum age.
stark contrast to the prevailing models for the Izanagi–Pacific and
We have incorporated the Japanese magnetic lineations and frac- Kula–Pacific ridges, but our interpretation is kinematically self-
ture zones of Sager et al. (1988) and Nakanishi et al. (1999) together consistent, matches geological observations and can be linked to the
with fracture zone traces based on satellite gravity anomaly data subduction history as seen in seismic tomography (Seton et al., in
(Sandwell and Smith, 2009) to define the seafloor spreading history preparation).
between the Izanagi and Pacific plates. Our resultant seafloor spread- The birth of the Izanagi plate is far more uncertain. The Izanagi
ing isochrons match well with the magnetic lineations seen in our plate must have existed prior to the birth of the Pacific plate as part
magnetic anomaly grid from M25 (~154 Ma) onwards when the of a three-plate ridge–ridge–ridge triple junction with the Farallon
magnetic anomaly signature is strongest (Fig. 6). Magnetic lineations and Phoenix plates, based on the rules of triple junction closure. How-
prior to M25 (~154 Ma) have larger variability (Tominaga and Sager, ever, there is no crust preserved in the seafloor spreading record
2010) and are not observed in our magnetic anomaly grid (Fig. 6) reflecting this early history as it has been progressively subducted
(see Tominaga and Sager, 2010 for details). The ridge jump prior to under the east Asian margin. We model a simple geometry whereby
M26 (~155 Ma) postulated by Sager et al. (1998) has not been incorpo- the spreading direction between the Izanagi–Farallon plates is con-
rated as we were unable to identify magnetic lineations or an aban- strained by the oldest Pacific–Izanagi and Pacific–Farallon isochrons
doned ridge. In addition, the conjugate ridge flank is absent. via triple junction closure, intermediate spreading rates and spread-
We incorporate the major 24° clockwise change in spreading di- ing symmetry. We constructed the positions of the spreading ridges
rection at M21 (~ 147 Ma) (Sager et al., 1988) primarily constrained by computing small circle arcs between Izanagi–Pacific, Farallon–Pacific
via the Kashima Fracture Zone which shows continuity from at least and Phoenix–Pacific spreading. The spreading direction between the
M28–M10 (~156–130 Ma) (Fig. 6). This major change in spreading Izanagi and Phoenix plates is similarly constrained using triple junction
direction is coincident with the eruption of the southern-end of the closure between the Pacific–Izanagi and Pacific–Phoenix plates and the
Shatsky Rise at the Farallon–Izanagi–Pacific triple junction followed length of the spreading ridges determined by intersection with the
by triple junction instability. According to the model of Sager et al. Pacific margins.
(1988) two simultaneous triple junctions and at least nine small,
short-lived ridge jumps occurred at the Pacific–Farallon–Izanagi junc- 3.2.2. Farallon plate
tion. This led to an 800 km northeast jump in the triple junction Early mapping of magnetic lineations in the western Pacific identi-
center clearly observed in the gridded magnetic anomaly dataset fied a set of NW–SE trending Mesozoic magnetic lineations loosely
228 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
bounded by the Shatsky and Hess Rises and the Mid Pacific Mountains oceanic crust on the Farallon plate subducted under North America
(Figs. 6 and 7). These lineations, termed the Hawaiian lineations, beginning in the late Mesozoic (Bunge and Grand, 2000) and clearly
formed during NE–SW directed spreading between the Pacific and imaged as seismically fast material under central and eastern North
now extinct Farallon plate between at least M29–M0 (~156–120 Ma) America (Bunge and Grand, 2000; Liu et al., 2010). The Hawaiian
(Larson et al., 1972; Atwater and Severinghaus, 1990). The Mesozoic lineations show a clockwise change in spreading direction at M11
(~133 Ma) (Sager et al., 1988; Atwater and Severinghaus, 1990) with
no major change in spreading direction during the early history of
a Pacific–Farallon spreading due to the uniformity of the magnetic linea-
tions (Fig. 6) even though fracture zone traces prior to M25 (~154 Ma)
are absent (Fig. 1). The pole of rotation to describe Mesozoic spreading
was likely located in the south or equatorial Pacific due to the slightly
60˚
fan-shaped nature of the lineations (Figs. 6 and 7).
YTT The Hawaiian lineations form a magnetic bight with the Japanese
lineation set in the north and trace the Pacific–Farallon–Izanagi triple
La
ST
ur
QT
junction (Fig. 6). The Shatsky Rise erupted along the triple junction
en
tia
25 W
center between M21 and 19 (~147–143 Ma), as confirmed by ODP
n
50˚
Ma
31
leg 198 (Mahoney et al., 2005) either as a result of a mantle plume
rg
in
25 21 18 13
31 6 5 J head reaching the surface or decompression melting at a mid-ocean
34
ridge (Mahoney et al., 2005; Sager, 2005). The eruption of the Shatsky
40˚ Mendocino
North America Rise was coincident with an 800 km, nine-stage jump in the location
Pionee
r of the triple junction during which time the triple junction switched
between ridge–ridge–ridge and ridge–ridge–transform configura-
30˚ Murray tions (Nakanishi et al., 1999). The triple junction regained its stability
21
18 13 after the initial eruptive phase followed by waning volcanism forming
34 31 25
Moloka
i
6 5
the Papanin Ridge along the triple junction center until M1 (~ 121–
20˚ R 124 Ma) (Nakanishi et al., 1999).
Pacific
In the south, the Hawaiian lineations disappear beneath the Mid-
Clarion C
5 Pacific Mountains obscuring the trace of the Pacific–Farallon–Phoenix
10˚ triple junction. Further east, the Hawaiian lineations form a complex
−170˚ −160˚ −150˚ −140˚ −130˚ −120˚ −110˚ −100˚ junction with several discrete fan-shaped lineation sets (e.g. Magellan
b and Mid-Pacific Mountain lineation sets) (Tamaki and Larson, 1988)
characteristic of crust that formed during microplate formation at fast-
spreading triple junction centers. These fan-shaped lineations were
60˚ active between M15 and M1 (~138–121 Ma). In addition, a set of
short ENE–WSW trending lineations south of the Mid-Pacific Mountains
YTT
has been identified as M21 (~147 Ma) to M14 (~136 Ma) (Nakanishi
La
ST
ur
25 W
Phoenix plate and the postulated Trinidad plate.
nM
50˚
31
East of the M-anomalies is a wide zone of crust which formed
ar
gin
240˚ 260˚ 280˚ 300˚ 320˚ 340˚ 240˚ 260˚ 280˚ 300˚ 320˚ 340˚
120 Ma 120 Ma
40˚
20˚
0˚
240˚ 260˚ 280˚ 300˚ 320˚ 340˚ 240˚ 260˚ 280˚ 300˚ 320˚ 340˚
100 Ma 100 Ma
40˚
20˚
0˚
240 ˚ 260 ˚ 280 ˚ 300 ˚ 320 ˚ 340 ˚ 240 ˚ 260 ˚ 280 ˚ 300 ˚ 320 ˚ 340 ˚
50 Ma 50 Ma
40˚
20˚
0˚
240˚ 260˚ 280˚ 300˚ 320˚ 340˚ 240˚ 260˚ 280˚ 300˚ 320˚ 340˚
30 Ma 30 Ma
40˚
20˚
0˚
240˚ 260˚ 280˚ 300˚ 320˚ 340˚ 240˚ 260˚ 280˚ 300˚ 320˚ 340˚
10 Ma 10 Ma
40˚
20˚
0˚
240˚ 260˚ 280˚ 300˚ 320˚ 340˚ 240˚ 260˚ 280˚ 300˚ 320˚ 340˚
0 Ma 0 Ma
40˚
20˚
0˚
NAM GRN NAM RIV FAR KUL VAN IBR JDF POR
NAM SAM PAC AFR FAR NAZ VAN COC IZA CAR
0 20 40 60 80 100 120 140 160
Plate Name
Age of Oceanic Lithosphere (m.yrs)
230 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
trench. The most prominent bend observed in all fracture zones in the Atwater and Severinghaus (1990). We use their seafloor spreading
northeast Pacific occurred just prior to Chron 33 (~79 Ma) where isochrons, with adjustments based on Nakanishi et al. (1992), for
spreading changed from roughly E–W to ENE–WSW (Atwater et al., the Mesozoic lineations. Our resultant seafloor spreading isochrons
1993). Atwater et al. (1993) suggested that the inferred continuity match well with our magnetic anomaly grid (Figs. 6 and 7) in the
of the spreading system provides evidence of a simple two-plate sys- north and central sections of the Mesozoic lineations but fail to ac-
tem during this time, negating the need for microplate formation (e.g. count for the fan-shaped lineations in the south. This is a direct con-
Chinook plate). Anomaly 33 (~79 Ma) corresponds to the oldest sequence of our decision to exclude the reconstruction of numerous
clearly identified magnetic anomaly related to Pacific–Kula spreading microplates at the Pacific–Farallon–Phoenix triple junction (e.g.
(Lonsdale, 1988; Atwater, 1989) (see Section 3.2.2.1: Kula plate), Magellan, Mid-Pacific Mountains and Trinidad lineation sets) and in-
marking the minimum timing for the initial break-up of the Farallon stead focus our model the broad-scale development of the area. To
plate. Spreading was reasonably steady between Chrons 32–24 the north, the Hawaiian Mesozoic lineations show a clear magnetic
(~71–53 Ma), connecting with spreading along the Kula–Pacific bight with the Japanese lineations (Figs. 6 and 7), highlighting the
ridge to the north at the Great Magnetic Bight (Fig. 7). Anomaly 24 geometric stability of the Pacific–Izanagi–Farallon triple junction
(~55–53 Ma; late Paleocene–early Eocene) corresponds to a major from M29 to M22 (~156–148 Ma). A major clockwise change in
hemisphere-wide plate reorganization event and is manifested in a spreading direction is recorded in the Japanese lineations and fracture
20° clockwise change in spreading direction between the Pacific and zones at M21 (~147 Ma) leading to a period of instability of the
Farallon plates from WSW–ENE to E–W (Atwater, 1989), a change Pacific–Izanagi–Farallon triple junction (see Section 3.2.1: Izanagi
in spreading direction between Pacific–Kula plates (Lonsdale, 1988) plate). Interestingly, this does not correspond to an adjustment of
and the break-up of the Farallon plate into the Vancouver plate at the Pacific–Farallon relative plate motion suggesting that the adjust-
either Chron 24 (~ 55–53 Ma) (Atwater, 1989) or 23 (~51–52 Ma) ment was related to the Shatsky Rise rather than a regional or global
(Menard, 1978; Rosa and Molnar, 1988) (Fig. 8). The break-up of plate reorganization.
the Farallon plate occurred in between the Pioneer and Murray frac- Finite rotations for the Pacific–Farallon ridge were derived using the
ture zones (Atwater, 1989) (Fig. 7) with oblique compression and half-stage pole method with an assumption of spreading symmetry and
slow relative motion (Rosa and Molnar, 1988). At this time, the average spreading rates. Reconstruction of the Pacific–Izanagi–Farallon,
mid-ocean ridge was located proximal to the subduction zone and Pacific–Phoenix–Farallon and Pacific–Kula–Farallon triple junctions
was followed by a period of complex spreading and/or spreading additionally followed the principles of triple junction closure. Al-
instability forming a “disturbed zone” between Anomalies 19–12 though ridge jumps have been proposed for early CNS spreading
(~41–31 Ma) (Atwater, 1989). Another major change in spreading (Atwater et al., 1993), we have followed a simple model of seafloor
direction is evident in the seafloor spreading record between the spreading throughout the CNS as we cannot identify remnant fea-
Murray and Pioneer fracture zones at Anomaly 10 (~ 28 Ma), forming tures describing the proposed ridge jumps without access to high-
the Monterey and Arguello plates (Atwater, 1989). South of the resolution multibeam bathymetry data. Towards the end of the CNS,
Murray fracture zone, the Guadalupe plate formed between Anoma- constraining the precise timing of the change in spreading direction
lies 7 and 5 (~ 25-10 Ma) (Mammerickx and Klitgord, 1982; Atwater, observed in the Mendocino, Molokai and Clarion fracture zones is
1989). These plates formed progressively as transform faults inter- difficult. We extrapolate using the Müller et al. (2008a) model and
sected with the Farallon subduction zone. After Chron 10 (~28 Ma), suggest that a change in spreading direction between the Pacific
the Vancouver plate is often referred to as the Juan De Fuca plate, co- and Farallon plates occurred at 103 Ma, closely corresponding with
inciding with the establishment of the San Andreas fault no earlier the observed bend in Pacific hotspots at ~ 99 Ma, implied by Veevers
than 30 Ma (Atwater, 1970) (Fig. 8). (2000) and Wessel and Kroenke (2008), based on an updated sea-
Spreading between the Pacific and Farallon plates during the mount dataset.
Mesozoic occurred in the region conjugate to the North American The Shatsky Rise formed at the Izanagi–Farallon–Pacific triple
margin. However, starting in the CNS, the Pacific–Farallon spreading junction and as a consequence, part of the Shatsky Rise must have
extended southward as far south as the Eltanin fracture zone in the erupted onto the Farallon and Izanagi plates. We have modeled the
South Pacific (Figs. 7 and 9). Magnetic anomalies 34 (~ 84 Ma) to 6 conjugate Shatsky Rise (Farallon) and find that it intersects the
(~20 Ma) on the Pacific plate associated with Pacific–Farallon spread- North American margin at 90 Ma, correlating well with the onset of
ing conjugate to the South American margin have been identified the Laramide Orogeny in western North America and a shallow seis-
(Herron, 1972; Cande et al., 1982; Mayes et al., 1990). This is restrict- mically fast region underlying western North America (Liu et al.,
ed to Anomalies 23–6 (~ 52–20 Ma) on the Nazca plate (Cande and 2010). As the geological evidence and seismic tomography images
Haxby, 1991). Seafloor spreading between Anomalies 34–21 (~84– are independent of the plate reconstructions used, our assumption
47 Ma) was reasonably stable until a major reorganization of the of largely symmetrical seafloor spreading and average spreading
spreading system at Chron 21 (~47 Ma), observed in fracture zone rates between Pacific–Farallon appears to be reasonable.
trends in the South Pacific (Mayes et al., 1990). The cessation of spread- After the CNS, we model seafloor spreading based on Atwater and
ing between the Pacific and Farallon plates occurred during break up Severinghaus (1990) for the northeast Pacific but without small-scale
into the Cocos and Nazca plates at 23 Ma (see Section 3.2.2.3: Nazca ridge adjustments associated with plate break-up events (Fig. 8). We
and Cocos plates). concur with the interpretation of Atwater et al. (1993) that the most
Our model for spreading between the Pacific and Farallon plates notable change in spreading direction observed in all northeast Pacific
incorporates spreading initiation at 190 Ma, based on the evidence fracture zones occurred at Chron 33 (~79 Ma). This timing corresponds
presented earlier in the manuscript (see Section 3.2.1: Izanagi to our initiation of seafloor spreading between the Kula and Pacific
plate), even though the oldest Hawaiian lineation identified is M29 plates and establishment of the Pacific–Kula–Farallon triple junction
(~156 Ma). The model we have implemented closely follows that of (see Section 3.2.2.1: Kula plate) (Fig. 8). Further southward, the
Fig. 8. (a) Agegrid reconstructions of the northeast Pacific at 120, 100, 50, 30, 10, 0 Ma highlighting the age–area distribution of oceanic lithosphere at the time of formation and the
extent of continental crust (gray polygons). Plate boundaries from our continuously closing plate polygon dataset are denoted as thick white lines, hotspot locations as yellow stars,
large igneous provinces and flood basalts as brown polygons and coastlines as thin black lines. (b) Reconstructions showing the outlines of the plates in the northeast Pacific for
each reconstruction time listed above. Feature descriptions as in panel (a). Abbreviations are: AFR = African plate, CAR = Caribbean plate, COC = Cocos plate, EUR = Eurasian
plate, FAR = Farallon plate, GRN = Greenland plate, IBR = Iberian plate, IZA = Izanagi plate, JDF = Juan de Fuca plate, KUL = Kula plate, NAM = North American plate,
NAZ = Nazca plate, PAC = Pacific plate, POR = Porcupine plate, RIV = Rivera plate, SAM = South American plate, VAN = Vancouver plate. (For interpretation of the references
to color in this figure legend, the reader is referred to the web version of this article.)
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 231
25 nar
d
6
18
21
−50˚ Elt
6 Antarctica Our model incorporates the Vancouver and Juan De Fuca plates
an 5 13
in
(Fig. 8) but excludes the other proposed microplates, such as the
H 18
5 6
13 18 21 Monterey, Arguello and Guadalupe plates, as no published poles of
25 21
rotation to describe their history are available. Spreading between
−60˚ the Pacific and Farallon plates ceased in the area to the west of
−150˚ −140˚ −130˚ −120˚ −110˚ −100˚ −90˚ −80˚ −70˚ South and Central America at 23 Ma (Chron 6B) as the plate separated
b into the Cocos and Nazca plates.
20˚
N.Am
3.2.2.1. Kula plate. The existence of the Kula plate during the late
10˚ Co 6 Cretaceous to the Paleocene/Eocene has been known since the
5
5 early identification of northward younging, E–W trending magnetic
21 18 13 6 G GR anomalies in the northern Pacific (Rea and Dixon, 1983; Lonsdale,
0˚ 34 31 25
South 1988; Mammerickx and Sharman, 1988; Atwater, 1990) (Fig. 7).
America These magnetic anomalies, located north of the Chinook Trough,
−10˚
B represent only the southern (Pacific) flank of Kula–Pacific spread-
EPR Nazca ing, the remainder having been subducted beneath the Aleutian
−20˚ 21 trench. The initiation of the Pacific–Kula ridge occurred within the
5
18 13 E 6
13 18
Farallon plate and marks the first stage of Farallon plate break-up.
6
−30˚ 5 Additionally, prevailing models of the Pacific (e.g. Engebretson et
34 3125 21
J al., 1985) imply that cessation of spreading along the Izanagi–Pacif-
Pacific ic ridge preceded the establishment of the Pacific–Kula Ridge,
5
−40˚ F
CR 6 therefore suggesting that Pacific–Izanagi and Pacific–Kula spread-
ing was not simultaneous. This assumption has implications for
PAR 5
25
Me 6
the formation of the northern Pacific and plate driving forces in
13
nar
d
18
21
−50˚ Elt
6 Antarctica the area.
an 5 13
in
The oldest well recognized magnetic anomaly associated with
H 18
5 6
13 18 21 Kula–Pacific spreading is either Anomaly 31 (~68 Ma) or possibly
25 21
32 (~71 Ma) (Rea and Dixon, 1983; Lonsdale, 1988) although some
−60˚ authors interpret Anomaly 33 (~79 Ma) (Mammerickx and Sharman,
−150˚ −140˚ −130˚ −120˚ −110˚ −100˚ −90˚ −80˚ −70˚ 1988) and tentatively Anomaly 34 (~83.5 Ma) (Atwater, 1990;
LIP Plate Boundary
Isochrons Norton, 2007). The conventional view is that after the death of the
product PAC-WANT FAR-WANT NAZ-COC
Extinct Ridge
PAC-FAR PAC-RIV NAZ-BAU
Izanagi plate, the locus of rifting and spreading jumped eastward to
Non-
oceanic Tectonic Boundary PAC-NAZ PAC-COC the Chinook Trough where E–W trending magnetic lineations formed
via simple Kula–Pacific spreading. However, Rea and Dixon (1983) pos-
tulated that two spreading ridges formed along existing Pacific–Farallon
−50 −40 −30 −20 −10 0 10 20 30 40 50 fracture zones after a change in spreading direction at ~83.5 Ma form-
Magnetic Intensity Anomaly [nTesla] ing a second plate, the Chinook plate, south of the Chinook Trough.
The Stalemate Fracture Zone delineates the western extent of the
Fig. 9. (a) Gridded magnetic anomalies for the southeast Pacific. Seafloor spreading iso- Kula plate (Fig. 6) and tracks the motion of the Kula plate from N–S
chrons used in this study plotted as thin black lines. Numbers correspond to magnetic
adjacent to Anomalies 34/31 (83.5–71 Ma) to 25 (~56 Ma) to NW
anomaly chron. B = Bauer Microplate, CR = Chile Ridge, E = Easter Microplate, EPR =
East Pacific Rise, F = Friday Microplate, G = Galapagos Microplate, GR = Galapagos from Anomalies 24 (~55–53 Ma) to 20/19 (~ 44–41 Ma). Additionally,
Ridge, J = Juan Fernandez Microplate, PAR = Pacific–Antarctic Ridge. (b) Seafloor spread- Lonsdale (1988) interpreted an extinct spreading ridge adjacent to
ing isochron map colored by spreading system or plate pair. Map abbreviations are same Anomalies 20/19 (~ 44–41 Ma) as well as a short sequence of Anom-
as a. Legend abbreviations are: BAU = Bauer, COC = Cocos, FAR = Farallon, NAZ = Nazca, alies 21–20 (47–44 Ma) on the western side of this extinct ridge.
PAC = Pacific, RIV = Rivera/Guadalope, WANT = West Antarctica/Antarctica.
The study of Lonsdale (1988) therefore suggests a spreading history
for the Kula plate involving N–S spreading from 32 to 25 (~71–
56 Ma) followed by a major change in plate motion by 20–25° at
Chron 24 (~55–53 Ma). The cessation of spreading along the Pacific–
232 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
Kula ridge was initially believed to have occurred at Chron 25 (~56 Ma) The change in spreading direction in the Kula plate identified by
(Byrne, 1979) and later to 43–47 Ma corresponding to the major Pacific Lonsdale (1988) matches with the change in the Pacific plate driven
plate reorganization event (Engebretson et al., 1985). The identification by the subduction of the Izanagi ridge (see Section 3.2.1: Izanagi
of an extinct spreading ridge in the far northwest corner of the plate by plate) and changes that were occurring along the Pacific–Farallon
Lonsdale (1988) further refined the cessation of spreading to around spreading system (see Section 3.2.2: Farallon plate).
Chron 18 (~40 Ma). However, the identification of this ridge was To the east, we model the Kula–Farallon ridge based on triple
based on a small number of ship tracks and seismic profiles. junction closure and the finite difference method resulting in a stable
To the east, the Kula plate is delineated by the Great Magnetic NE–SW orientation of the Kula–Farallon ridge, consistent with previ-
Bight, which traces the Pacific–Kula-Farallon triple junction in a ous studies. The Yellowstone hotspot located offshore the North
ridge–ridge–ridge configuration from Chron 34/31 (~84–71 Ma) to American margin in the Paleocene/Eocene (Fig. 8) was used as a
25 (~56 Ma) (Fig. 7). This is followed by a “T” anomaly corresponding further constraint to guide the position of the NE–SW trending Kula–
to Chron 24 (55–53 Ma), which likely formed during a reorganization Farallon ridge, as mid-ocean ridges are known to preferentially evolve
of the Pacific–Kula–Farallon triple junction (Lonsdale, 1988; Atwater, near hotspots (Müller et al., 1998b). As a result our modeled position
1990). of the Kula–Farallon ridge with respect to the North American margin
The Great Magnetic Bight traces the location of the Kula–Farallon– correlates with onshore geological and geochemical evidence of a
Pacific triple junction (Figs. 7 and 8). Previous models have predicted northward migrating slab window near the northern US/Canadian
the location and orientation of the resultant Kula–Farallon ridge margin (Atwater, 1990; Breitsprecher et al., 2003; Madsen et al.,
(for which there is no preserved evidence in the seafloor spreading 2006). Additionally, our position of the Kula–Farallon ridge is supported
record) based on triple junction closure and tracking evidence of by seismic tomography (Bunge and Grand, 2000).
a slab window beneath western North American margin (e.g.
Engebretson et al., 1985; Atwater, 1990; Breitsprecher et al., 2003;
Madsen et al., 2006). Most models lead to a reasonably consistent re- 3.2.2.2. Vancouver/Juan De Fuca plate. The recognition of a difference in
sult of a NE–SW trending spreading ridge intersecting the North trend by about 11° between the fracture zones north of the Murray
American margin and forming a slab window somewhere near the fracture zone in the northeast Pacific and those to the south (Fig. 1),
present-day Pacific Northwest (Engebretson et al., 1985; Atwater, led Menard (1978) to suggest that the Farallon plate broke into two
1990; Breitsprecher et al., 2003; Madsen et al., 2006). plates around 47–49 Ma Ma (Chrons 22–21). Menard (1978) termed
Our interpretation for the Kula plate closely follows the model the new plate north of the Murray Fracture Zone, the Vancouver
of Lonsdale (1988). However, we have interpreted Anomaly 33 plate. Differential motion between the Vancouver and Farallon plate
(~79 Ma) north of the Chinook Trough as the oldest identified magnetic was confirmed and dated to Chron 21 (~47 Ma) with the spacing of
anomaly based on the interpretation in Atwater (1990) and our own magnetic anomalies in the area between the Murray and Pioneer
analysis of the magnetic anomalies in the area. Most authors have fracture zone possibly indicating either asymmetric spreading or a
only been able to interpret anomalies back to 32 (~71 Ma) as it is the ridge jump between Anomalies 21 (~47 Ma) and 13 (~33 Ma) (Rosa
last clearly identified magnetic anomaly, however the new gridded and Molnar, 1988). The model of Rosa and Molnar (1988) implies
magnetic anomaly datasets such as WDMAM, EMAG2 and our own slow transpressional motion across the plate boundary, which lies be-
gridded compilation (Figs. 6–7) show E–W trending magnetic linea- tween the Murray and Pioneer fracture zones as a “set of curving,
tions south of Anomaly 32 (~71 Ma). There is space south of our inter- tooth-like disjunctures” (Atwater, 1990) clearly seen between Anom-
preted Anomaly 33 (~79 Ma) to accommodate a very small portion of alies 19 and 13 (~41–33 Ma) (Fig. 7) possibly indicative of diffuse
older crust (possibly back to Anomaly 34 (~84 Ma)), but we believe deformation.
that the establishment of the stable Pacific–Kula–Farallon triple junc- The intersection of the Murray transform fault with the North
tion in a ridge–ridge–ridge configuration must have occurred at 33 American subduction zone around 30 Ma led to the establishment
(~79 Ma) and not earlier. Importantly, our model has contemporaneous of the San Andreas Fault and corresponds to the establishment of
Pacific–Izanagi and Pacific–Kula spreading (see Section 3.2.1: Izanagi the Juan De Fuca plate at the expense of the Vancouver plate. The
plate) joined by a NNW–SSE transform. In our model, we have continu- spreading history of the Juan De Fuca plate is very complex (Wilson
ing N–S directed Pacific–Kula spreading until Anomaly 25 (~56 Ma) et al., 1984; Wilson, 1988) most likely due to its proximity to the
followed by an anticlockwise change in spreading direction starting at Cascadia subduction zone. Spreading involved counter-clockwise mo-
Anomaly 24 (~55–53 Ma), as suggested by Lonsdale (1988) and tion followed by progressive clockwise rotation starting at Chron 5D
expressed in the Stalemate Fracture Zone. The magnetic anomaly (~17 Ma) (Atwater, 1990) and a series of propagating rifts and micro-
grids clearly show the NE–SW trending magnetic lineations corre- plate formation (Wilson et al., 1984; Wilson, 1988). Currently, the
sponding to the youngest part of Pacific–Kula spreading (Fig. 6). We fol- Juan De Fuca plate is limited at its southern end by the Mendocino
low the interpretation of Lonsdale (1988) for the cessation of Pacific– Fracture Zone and is subducting slowly along the Cascadia subduction
Kula spreading to be around 41–40 Ma. We compute finite rotations zone (Fig. 7).
based on the half-stage pole method between Chrons 33–22 (~79– Our reconstructions of the Vancouver/Juan De Fuca plates are
49 Ma) as only the Pacific flank of the spreading system is preserved. largely based on the detailed tectonic maps of Atwater and
We use the magnetic lineations of Lonsdale (1988) and the Stalemate Severinghaus (1990) unchanged from the model used by Müller et
Fracture Zone to compute finite rotations between Chrons 21–20 al. (1997). We implement the break-up of the Farallon plate into
(~47–44 Ma) using the traditional method. the Farallon and Vancouver plates along the Pioneer Fracture Zone
The factor leading to the abrupt change in plate motion between at Chron 22 (~50–49 Ma) (Fig. 8). We use the finite rotations from
the Kula and Pacific plates was suggested to be a result of the tempo- Müller et al. (1997) for the Vancouver plate and the rotations in this
rary elimination of northward slab pull when subduction shifted from study for the Farallon plate. Our rotations result in transpressional
the Siberian margin to the Aleutian Trench (Lonsdale, 1988). In our motion along the transform fault connecting the Farallon and
model, we argue that the subduction of the Izanagi–Pacific ridge at Vancouver plates. The Juan de Fuca plate is modeled as a simple
55–50 Ma resulted in the temporary cessation of subduction and two-plate system and we do not include the detailed interpretation
slab break-off along the east Asian margin leading to a change in of Wilson (1988) as there are no rotations associated with the iso-
motion of the Kula plate to the northwest. The intersection of the chrons making it difficult to incorporate into our tectonic model. On
Pacific–Izanagi ridge with subduction under East Asia eliminated the the broad scale, our seafloor spreading isochrons match well with
ridge push force thus enabling the Kula plate to move to the west. the magnetic lineations from our magnetic grid compilation (Fig. 7),
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 233
120 Ma 0˚ 120 Ma
20˚
40˚
60˚
220̊ 240̊ 260̊ 280̊ 300̊ 320̊ 340̊ 220̊ 240̊ 260̊ 280̊ 300̊ 320̊ 340̊
100 Ma 0˚ 100 Ma
20˚
40˚
60˚
220˚ 240˚ 260˚ 280˚ 300˚ 320˚ 340˚ 220˚ 240˚ 260˚ 280˚ 300˚ 320˚ 340˚
80 Ma 0˚ 80 Ma
20˚
40˚
60˚
220˚ 240˚ 260˚ 280˚ 300˚ 320˚ 340˚ 220˚ 240˚ 260˚ 280˚ 300˚ 320˚ 340˚
40 Ma 0˚ 40 Ma
20˚
40˚
60˚
220˚ 240˚ 260˚ 280˚ 300˚ 320˚ 340˚ 220˚ 240˚ 260˚ 280˚ 300˚ 320˚ 340˚
10 Ma 0˚ 10 Ma
20˚
40˚
60˚
220˚ 240˚ 260˚ 280˚ 300˚ 320˚ 340˚ 220˚ 240˚ 260˚ 280˚ 300˚ 320˚ 340˚
0 Ma 0˚ 0 Ma
20˚
40˚
60˚
ANT HIK SND NSC SSC CQL BAU CAZ ESC ESC
NAM SAM PAC AFR FAR NAZ MAN COC IZA CAR
0 10 20 30 40 50 60 70 80 90 100110120130140150160260
Age of Oceanic Lithosphere (m.yrs) Plate Name
234 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
however there are some inconsistencies, particularly approaching the day record, it has been suggested that isolated sections of Nazca–Pacific
trench as we do not include small scale block rotations. spreading have been captured by the Cocos plate to the north and
subsequently subducted under the Middle America trench (Tebbens
3.2.2.3. Nazca and Cocos plates. The East Pacific Rise is currently the and Cande, 1997). Finite rotations and their uncertainties to describe
site of very fast seafloor spreading between the Pacific and Nazca the post break-up phase of Nazca–Pacific and Nazca–Antarctic motion
and Cocos plates and dominates the seafloor of the SE Pacific were computed using a combination of the Hellinger technique
(Fig. 9). Other active seafloor spreading ridges are the Chile Ridge (Tebbens and Cande, 1997), existing rotations (Pardo-Casas and
(active spreading between the Nazca and Antarctic plates) and the Molnar, 1987) and the interpretation of South Pacific magnetic anoma-
Galapagos Spreading Centre Nazca–Cocos spreading) (Fig. 9). The lies (Mayes et al. 1990).
Nazca plate incorporates oceanic crust that formed as a result of A major component of the seafloor spreading history of the Nazca
Pacific–Nazca, Pacific–Farallon, Nazca–Cocos and Nazca–Antarctic plate involves the formation of the Bauer Microplate (Fig. 9). The
spreading as well as the Bauer microplate (Fig. 9). The Cocos plate in- Bauer microplate formed along the northern East Pacific Rise and
cludes oceanic crust that formed as a result of Cocos–Pacific and grew by crustal accretion and counter-clockwise rotation between
Cocos–Nazca as well as spreading in the Rivera and Mathematician Pacific and Nazca spreading (Goff and Cochran, 1996; Eakins and
microplates. Lonsdale, 2003) shortly after a major plate reorganization event at
Both the Nazca and Cocos plates formed as a result of the break-up 20 Ma (Fig. 9). The formation of the Bauer microplate is unlike the
of the southern part of the Farallon plate at approximately 23 Ma step-wise triple junction migration models used to explain the forma-
(Hey, 1977; Lonsdale, 2005) or Chron 6By (~ 23 Ma) (Barckhausen tion of the microplates associated with the Pacific–Nazca–Antarctic
et al., 2008). The break-up of the Farallon plate is believed to have triple junction. Spreading is believed to have initiated at 17 Ma via
been driven by a combination of increased northward pull after the northward propagation of the East Pacific Rise and southward propa-
earlier break-up of the Farallon plate to the north (Lonsdale, 2005), gation of the Galapagos Rise during counter clockwise rotation of the
an increase in slab pull at the Middle America subduction zone due spreading axes (Eakins and Lonsdale, 2003). Rotation and spreading
to an increase in its length (Lonsdale, 2005) and/or the weakening continued about a pole proximal to the spreading axis creating fan-
of the plate along the point of break-up due to the influence of the shaped anomalies until 6 Ma when the spreading ridge realigned
Galapagos hotspot (Hey, 1977; Lonsdale, 2005; Barckhausen et al., with the dominant East Pacific Rise spreading ridge and the Bauer
2008). In addition, plate break-up was preceded by a major plate re- microplate was captured by the Nazca plate (Eakins and Lonsdale,
organization in the Southeast Pacific at 24 Ma leading to a change in 2003) (Fig. 10).
motion of the Farallon plate 1–2 million years before break-up The other smaller microplates within the Nazca/Cocos/Pacific
(Tebbens and Cande, 1997; Lonsdale, 2005; Barckhausen et al., realm are the presently active Easter and Juan Fernandez microplates,
2008). Although the Nazca and Cocos plates are now independent which form small pseudo-circular plates along the actively spreading
plates, an interpretation of their history must consider the evolution East Pacific Rise. These plates are believed to have become active at
of the Farallon plate (see Section 3.2.2: Farallon plate) to understand around Chron 3o (~5 Ma) during a major plate reorganization event
the nature of the oceanic lithosphere in this region older than 23 Ma. in the SE Pacific (Tebbens and Cande, 1997) and have rotated about
The oldest portion of the Nazca plate, adjacent to the South American an axis close to the center of the plate by between 80 and 90°
margin includes the crust that formed due to Farallon–Pacific spreading. (Searle et al., 1993a). The mechanism for the formation of these plates
Magnetic anomalies up to Anomaly 23 (~51 Ma) have been tentatively is believed to be the same process responsible for the development of
identified on the Nazca plate (Cande and Haxby, 1991) but most models the Hudson and Friday microplates related to the northward migrat-
confidently identify magnetic anomalies only back to Anomaly 13 ing Nazca–Pacific–Antarctic ridge (Bird et al., 1998).
(~33 Ma) (Handschumacher, 1976; Pardo-Casas and Molnar, 1987; To the north, the Cocos–Pacific spreading ridge was only estab-
Tebbens and Cande, 1997). Pardo-Cassas and Molnar (1987) and Rosa lished in its present form from Chron 2A (~3 Ma) (Atwater, 1990).
and Molnar (1988) computed finite rotations and their uncertainties Between 23 Ma and Chron 2A (~ 3 Ma), spreading was being accom-
to describe the motion of Pacific–Farallon spreading by assuming modated along the Mathematician and Rivera Ridges to the north
symmetrical spreading where both flanks were not presently preserved. and the Cocos–Pacific to the south (Atwater, 1990; Eakins and
These rotations were used as a basis for the rotation model of Tebbens Lonsdale, 2003). Spreading in this area included many block rotations
and Cande (1997) for the Nazca–Pacific–Antarctic triple junction. A and ridge jumps possibly due to the proximity of the Cocos–Pacific
South Pacific-wide study by Mayes et al. (1990) computed rotations spreading center to the Middle America trench and Galapagos hot-
for the Pacific–Farallon and Pacific–Nazca ridges. spot. The magnetic lineations that formed due to Cocos–Pacific
The crust that formed between the Pacific–Nazca plates subse- spreading are fan-shaped with strongly curved fracture zones ob-
quent to plate break-up at 23 Ma has a complex spreading history. served in the satellite gravity anomalies indicating a pole of rotation
Spreading occurred as a northward “step-wise triple junction migra- close to the northern end of the plate (Figs. 1 and 9).
tion” (see Tebbens and Cande, 1997 for a description of this process) The present day Cocos–Nazca ridge strides the Galapagos hotspot
between the Pacific–Nazca–Antarctic ridges, leaving behind a record and intersects the Middle America convergent margin at the Bulboa
of ridge jumps and microcontinent formation particularly at Anoma- Fracture Zone (Fig. 9). This E–W directed spreading ridge was estab-
lies 6 (~20 Ma) and 5A (~12 Ma) (Tebbens and Cande, 1997) includ- lished around 23 Ma, coinciding with the break-up of the Farallon
ing the Friday microplate south of the Chile Fracture Zone (Fig. 9). plate. The early spreading history is quite complex, requiring several
This complexity in the spreading pattern has hindered the interpreta- ridge jumps during its formation (Barckhausen et al., 2008), the
tion of magnetic anomalies post-Oligocene. Although most of the most significant of which is the Malpelo Ridge, which became extinct
crust created during this spreading phase is preserved in the present around 15–10 Ma (Meschede et al., 1998a). In addition, numerous
Fig. 10. (left) Agegrid reconstructions of the southeast Pacific at 120, 100, 80, 40, 10, 0 Ma highlighting the age–area distribution of oceanic lithosphere at the time of formation and
the extent of continental crust (gray polygons). Plate boundaries from our continuously closing plate polygon dataset are denoted as thick white lines, hotspot locations as yellow
stars, large igneous provinces and flood basalts as brown polygons and coastlines as thin black lines. (right) Reconstructions showing the outlines of the plates in the southeast
Pacific for each reconstruction time listed above. Feature descriptions as in panel (left). Abbreviations are: AFR = African plate, ANT = Antarctic plate, BAU = Bauer plate, CAR
= Caribbean plate, CAZ = Chasca plate, COC = Cocos plate, CQL = Catquil plate, ESC = East Scotia Sea plate, FAR = Farallon plate, HIK = Hikurangi plate, IZA = Izanagi plate,
MAN = Manihiki plate, NAM = North American plate, NAZ = Nazca plate, NSC = North Scotia Sea plate, PAC = Pacific plate, SAM = South American plate, SND = Sandwich
plate, SSC = South Scotia Sea plate. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 235
pseudo-faults indicating rift propagation to the east have been identified spreading migrating microplates of the East Pacific Rise (Tebbens
in the seafloor fabric, the majority in the vicinity of the Galapagos hot- and Cande, 1997).
spot (Atwater, 1990). Further complications occur close to the Middle The ocean floor within the Ellice Basin and directly east of the
America trench where several ridge jumps have isolated spreading sys- Tonga–Kermadec subduction zone is intrinsically linked to the evolu-
tems, particularly in the Panama Basin (Lonsdale and Klitgord, 1978). tion of the Pacific–Phoenix ridge after M0 (~120 Ma) (Figs. 6 and 11).
We incorporate the magnetic anomaly identifications from
Munschy et al. (1996) to derive a set of finite rotations and seafloor
spreading isochrons between the Pacific and Nazca plates and also ex- a 10˚
M29 M21
tend our analysis to include the parts of Pacific–Farallon spreading M16
that are currently preserved on the Nazca plate. The magnetic anom- M10
aly identifications of Munschy et al. (1996) do not extend to the east- 0˚ M4
34 1
ernmost Nazca plate where we would expect to find the oldest OJP M0
preserved oceanic lithosphere corresponding to Pacific–Farallon MP
Pacific
SS EB MP
spreading, mainly due to a lack of data and signal intensity. Instead, −10˚
we predict the age of the oceanic lithosphere in this area by reconstruct- CS
ing the conjugate Pacific–Nazca isochrons. We find that the resultant NLB
NFB
location of isochrons closely corresponds to the interpretation of mag- −20˚ LB
netic anomalies from Cande and Haxby (1991) and matches well with
SFB OT
the magnetic lineations observed on our magnetic anomaly grid
Australia
(Fig. 9). Thus, our model predicts that the oldest ocean floor off South −30˚ HT
LHR 34 31
America corresponds to Anomaly 23 (~51 Ma) (Fig. 10). We derive a
new set of finite rotations to describe Pacific–Nazca spreading largely
based on the rotations of Mayes et al. (1990) to be consistent with our 34
−40˚ HP
magnetic pick compilation. We do not incorporate the detailed triple TS
junction migration model of Tebbens and Cande (1997). CR
together with the finite rotations derived from Eakins and Lonsdale NLB
NFB
(2003) between 17.3 and 11.9 Ma and newly derived finite rotation −20˚ LB
for 23 Ma and 10.9 Ma. We reconstruct the shape and location of the
SFB OT
Cocos Ridge from Meschede et al. (1998b). We model spreading along
Australia
the Galapagos Spreading Centre (Cocos–Nazca) based on the finite dif- −30˚ HT
34 31
ference method. We do not include the small-scale ridge jumps that LHR
Early models predicted that the area formed as part of a simple, con- The plateau break-up model (Taylor, 2006) suggests that the Ontong
tinuous N–S directed spreading system until the end on the CNS Java Plateau, Manihiki and Hikurangi Plateaus were joined at the time of
(Larson and Chase, 1972). However, the anomalously fast seafloor their eruption. This mega-LIP erupted around Aptian time based on the
spreading rates required to populate the region with crust formed dating of sediment overlying pillow basalts (Winterer et al., 1974) and
during the CNS (Atwater, 1990) as well as the identification of tectonic Ar/Ar dating (Mahoney et al., 1993). Taylor (2006) based his interpreta-
structures and seafloor fabric such as the E–W trending Nova Canton tion on recently collected marine geophysical data from the Ellice Basin,
Trough, the E–W trending Osbourn Trough and the N–S directed sea- which he believes was formed during the separation of the Ontong Java
floor fabric and side-stepping fracture zones in the Ellice Basin suggest and Manihiki Plateau and confirmed by Chandler et al. (in press). In the
a more complex history for the area. Based on the interpretation of Taylor (2006) model, the Nova-Canton Trough is interpreted as an ex-
the seafloor spreading structures, two distinct models have been devel- tension of the Clipperton Fracture Zone (Larson et al., 1972; Joseph et
oped to explain the evolution of the Pacific–Phoenix ridge after M1/M0 al., 1990; Taylor, 2006) based on side-scan sonar data (Joseph et al.,
(~123–120 Ma): a successive southward ridge jump model (Winterer, 1992) and not an abandoned spreading ridge. The disturbed “rift
1976; Larson, 1997; Billen and Stock, 2000; Müller et al., 2008b) and a zone” identified by Larson (1997) is instead interpreted as the northern
plateau break-up model (Taylor, 2006). part of an E–W directed spreading system with stair-stepped, large off-
In the successive ridge jump model the Nova Canton Trough, an E–W set E–W trending fracture zones and N–S abyssal hill fabric (Taylor,
gravity low located south and parallel to the Mesozoic lineations (Figs. 6 2006) separating the Ontong Java and Manihiki Plateaus. This model
and 11), is interpreted as an abandoned spreading center associated suggests that after M0 (~120 Ma), the tectonic regime changed from
with Pacific–Phoenix spreading (Rosendahl et al., 1975; Winterer, N–S directed Pacific–Phoenix spreading to E–W directed spreading be-
1976; Müller et al., 2008b). A zone of disrupted seafloor fabric bounded tween the Pacific plate and a new Manihiki plate. Coincidently, N–S
by two prominent E–W trending gravity lows in the northern Ellice directed spreading was occurring between the Manihiki and Hikurangi
Basin observed in satellite gravity data led to the idea of a rift zone plateaus, as suggested in the previous model. The differential motion
associated with N–S directed spreading along the Pacific–Phoenix between the two spreading systems requires a triple junction between
ridge (Larson, 1997). The abandoned ridge/rift zone model implies the Pacific, Manihiki and Hikurangi plates (Taylor, 2006). The timing of
that the Pacific–Phoenix ridge either became extinct shortly after M0 plateau break-up is unconstrained from the seafloor spreading record
(~120 Ma) or that the spreading ridge jumped to another location, like- as no magnetic anomalies can be interpreted. However, rift structures
ly to the south subsequent to M0 (~120 Ma), during a regional plate re- on the eastern side of the Ontong Java plateau and western margin
organization. The timing is constrained by the identification of magnetic of the Manihiki plateau suggest that this occurred around 120 Ma,
anomaly M0 (~120 Ma) just north of the Nova-Canton Trough (Larson, matching well with the dated break-up of the Manihiki and Hikurangi
1997; Nakanishi and Winterer, 1998). The southern ridge jump model is plateaus. Further supporting the common origin of the Ontong Java,
supported by the identification of the E–W trending Osbourn Trough Manihiki and Hikurangi Plateaus is similar geochemical compositions
(located to the east of the Tonga–Kermadec Trench and north of the between the three plateaus suggesting a related source (Mahoney et
Louisville Seamount Chain) as an extinct spreading ridge of Cretaceous al., 1993; Hoernle et al., 2010).
age (Lonsdale, 1997; Billen and Stock, 2000) (Fig. 11) rather than a The other main feature on the seafloor attributed to Pacific–Phoenix
late stage crack in the Pacific plate (Small and Abbott, 1998). spreading is the Tongareva triple junction trace in the SW Pacific
The seafloor spreading morphology in the vicinity of the Osbourn (Larson et al., 2002; Pockalny et al., 2002; Viso et al., 2005). The
Trough confirms roughly north–south spreading along a slow- Tongareva triple junction trace is a roughly NNW–SSE linear feature
intermediate spreading center (Worthington et al., 2006; Downey which starts at the northeastern corner of the Manihiki Plateau in the
et al., 2007) whereas the early motion appears to be parallel to the Pernyn Basin and extends to west of the Cook Islands before it changes
Wishbone Ridge (Figs. 1 and 2g). Spreading along the Osbourn trend to NW–SE until it reaches spreading associated with Pacific–
Trough is believed to have initiated right after M0 (~120 Ma) (Davy Antarctic Ridge (Fig. 11). The western side of the triple junction trace
et al., 2008) leading to the separation of the Manihiki and Hikurangi consists of ENE trending abyssal hill topography and directly east, the
Plateaus. The timing cannot be constrained from the seafloor spread- morphology is NNW–SSE trending (Larson et al., 2002; Pockalny et al.,
ing record as the early crust would have formed during the CNS. 2002). This lineament is believed to record the migration of a ridge–
Instead, the timing for the initiation of spreading is constrained ridge–ridge triple junction between the Pacific–Farallon–Phoenix plates
from the dating of rift-related structures on the southern side of the (Larson et al., 2002) whereas more detailed analysis revealed that the
Manihiki Plateau (e.g. Nassau-Suwarrow Scarp) and the northern triple junction likely flipped between ridge–ridge–ridge and ridge–
side of the Hikurangi Plateau (e.g. Rapuhia Scarp) (Lonsdale, 1997; ridge–transform configurations throughout its evolution (Pockalny
Billen and Stock, 2000; Sutherland and Hollis, 2001; Davy et al., et al., 2002). Sutherland and Hollis (2001) suggested that this lineament
2008). The cessation of spreading is poorly constrained but most was a rift but this has been refuted by subsequent studies (e.g. (Larson
authors tie the termination of spreading along the Osbourn Trough et al., 2002). The eastern margin of the Manihiki Plateau comprises a
with the docking of the Hikurangi Plateau to the Chatham Rise. Unfor- dramatic transtensional scarp (Winterer et al., 1974; Stock et al.,
tunately, the timing of collision between the Hikurangi Plateau and 1998) suggesting that the easternmost portion of a presumably larger
Chatham Rise is also ill constrained. Some authors favor collision at Manihiki Plateau was rifted off the margin and was controlled by the
105–100 Ma (Lonsdale, 1997; Sutherland and Hollis, 2001; Davy et plate motions related to the triple junction. Larson et al. (2002) hypoth-
al., 2008) based on geological observations and the onset of extension esized that a piece traveled across Panthalassa on the Farallon plate and
in New Zealand whereas others favor collision around 80–86 Ma another piece rifted to the south with the Phoenix plate. The timing for
(Billen and Stock, 2000; Worthington et al., 2006). The youngest mag- activity along the triple junction is poorly constrained. Spreading is
netic anomalies associated with the Osbourn Trough are as young as believed to have initiated around 120 Ma, based on the dating of
Anomalies 33 (~79 Ma) or 32 (~71 Ma) (Billen and Stock, 2000) or carbonate sedimentation on the Manihiki Plateau (Larson et al., 2002)
Anomaly 34 (~84 Ma) but prior to ~ 87 Ma (Downey et al., 2007). with termination around 84 Ma (Larson et al., 2002).
Based on the age range allowed from the magnetic anomaly interpre- Our model for the evolution of the Phoenix plate incorporates
tation and the age constraints on the initiation of spreading between simple N–S directed spreading in the Mesozoic followed by a major
the Pacific and Antarctic plate to the south, Müller et al. (2008b) sug- plate reorganization at ~ 120 Ma (M0) coincident with the eruption
gested that the spreading along the Osbourn Trough ceased at 85 Ma, of the Ontong Java–Manihiki–Hikurangi plateau as one mega-LIP, as
leading to a final jump in the plate boundary to the south along the suggested by Taylor (2006) and Chandler et al. (in press) (Fig. 10).
present day Pacific–Antarctic ridge. This spreading system shuts down at 86 Ma, after which spreading
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 237
120˚ 140˚ 160˚ 180˚ 200˚ 220˚ 120˚ 140˚ 160˚ 180˚ 200˚ 220˚
was accommodated along the Pacific–Farallon and Pacific–Antarctic
140 Ma 140 Ma Ridges (Figs. 10 and 12).
20˚ The Mesozoic lineations are constrained by magnetic anomaly
identification from Munschy et al. (1996), with geophysical data
(including satellite derived gravity data) constraining the location of
40˚
the Osbourn Trough, Nova Canton Trough and Tongareva triple junction
trace. Our seafloor spreading isochrons match the magnetic anomaly grid
quite well for the central and western part of the Mesozoic lineations but
60˚ there is a poor match to the east corresponding to the fan-shaped Magel-
lan and Mid-Pacific Mountain lineations (Figs. 6 and 11). We do not re-
120˚ 140˚ 160˚ 180˚ 200˚ 220˚ 120˚ 140˚ 160˚ 180˚ 200˚ 220˚ construct these complex lineation sets due to a lack of age constrains
on initiation and cessation of the microplates at the Pacific–Phoenix–Far-
120 Ma 120 Ma
20˚
allon triple junction that would have formed these lineations. In addition,
our aim is to model the broad scale development and the evolution of the
larger plates in the area rather than the smaller scale microplates. Finite
40˚ rotations are derived for the E–W trending M-series anomalies by using
the half-stage pole methodology and following the fracture zones traced
from satellite gravity data (Sandwell and Smith 2009).
60˚ Our reconstructions are based on the model of Taylor (2006) and
Chandler et al. (in press) with roughly E–W directed spreading form-
120˚ 140˚ 160˚ 180˚ 200˚ 220˚ 120˚ 140˚ 160˚ 180˚ 200˚ 220˚ ing the crust underlying the Ellice Basin between the Ontong Java and
Manihiki Plateaus and simultaneous rifting of the Manihiki and
80 Ma 80 Ma Hikurangi plateaus from a N–S directed spreading system along the
20˚
Osbourn Trough. We initiate this spreading system at 120 Ma, corre-
sponding to the timing of the LIP eruption and the dating of rift-
40˚ related sequences along the margin. The oceanic crust between
these plateaus formed during the CNS so no correlations can be ob-
served in the magnetic anomaly grids (Fig. 11). However, the satellite
60˚ derived gravity data indicates fracture zone trends and limited abys-
sal hill fabric. We derive our own finite rotations for the opening of
120˚ 140˚ 160˚ 180˚ 200˚ 220˚ 120˚ 140˚ 160˚ 180˚ 200˚ 220˚ the Osbourn Trough region by following fracture zone traces. The sep-
aration of the mega-LIP requires that a triple junction was active ac-
40 Ma 40 Ma commodating motion between the Ontong Java and Hikurangi
20˚ Plateaus during its formation. We reconstruct the arm of the triple
junction based on the finite difference method.
40˚ We suggest a further two triple junctions were located to the east
of the Manihiki Plateau, one of which formed the Tongareva triple
junction trace. However, unlike previous interpretations (Larson et
al., 2002; Viso et al., 2005), we suggest that the triple junction repre-
60˚
sented spreading between the Manihiki, Hikurangi and a new plate
we term the Chasca plate to the east of the Manihiki and Hikurangi
120˚ 140˚ 160˚ 180˚ 200˚ 220˚ 120˚ 140˚ 160˚ 180˚ 200˚ 220˚
plates (Fig. 10). The Chasca plate, which was located off the South
20 Ma a American margin, is named after the Incan goddess of dawn and twi-
20˚ light. Our finite rotations were derived by using a combination of
fracture zone and triple junction traces and the finite difference
method. A second triple junction between the Hikurangi, Manihiki
40˚
and a new plate we term the Catequil plate was required to account
for the trends in the seafloor fabric to the west of the Tongareva triple
junction trace. The Catequil plate is named after the Incan god of
60˚ thunder and lightning.
The fracture zone traces between the Manihiki and Hikurangi
120˚ 140˚ 160˚ 180˚ 200˚ 220˚ 120˚ 140˚ 160˚ 180˚ 200˚ 220˚ Plateau show a change in direction but this change has never been
0 Ma 0 Ma
20˚ Fig. 12. (left) Agegrid reconstructions of the southwest Pacific at 140, 120, 80, 40, 20,
0 Ma highlighting the age–area distribution of oceanic lithosphere at the time of forma-
tion and the extent of continental crust (gray polygons). Plate boundaries from our
40˚ continuously closing plate polygon dataset are denoted as thick white lines, hotspot lo-
cations as yellow stars, large igneous provinces and flood basalts as brown polygons
and coastlines as thin black lines. (right) Reconstructions showing the outlines of the
plates in the southwest Pacific for each reconstruction time listed above. Feature de-
60˚ scriptions as in panel (left). Abbreviations are: ANT = Antarctic plate, AUS = Austra-
lian plate, CAR = Caroline plate, ENK = East Norfolk Basin plate, EUR = Eurasian
plate, HIK = Hikurangi plate, IZA = Izanagi plate, JUN = Junction plate, LAU = Lau
SLY AUS NTY LAU PHL NFB ENK WNK SOL CAR
Basin plate, LHR = Lord Howe Rise plate, NBR = New Britain plate, NFB = North Fiji
ANT AUS PAC LHR EUR PHX JUN HIK IZA NBR Basin plate, NTY = Neo-Tethys plate, PAC = Pacific plate, PHL = Philippine Sea
0 20 40 60 80 100 120 140 160 plate, PHX = Phoenix plate, SLY = South Loyalty Basin plate, SOL = Solomon Sea
Age of Oceanic Lithosphere (m.yrs)
Plate Name plate, WNK = West Norfolk Basin plate. (For interpretation of the references to color
in this figure legend, the reader is referred to the web version of this article.)
238 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
Schist (Vry et al., 2004) and the seafloor spreading constraints pre- Pacific
SEIR
sented in Billen and Stock (2000). The docking led to the shut-down EM
of the seafloor spreading system in the South Pacific and a change CP Australia
Scotia
the Pacific and Antarctic plates started earlier than observed or that
there were two contemporaneous spreading ridges located in close WS
proximity in the South Pacific. It would also require fast seafloor- RLS
spreading rates between the Manihiki and Hikurangi plateaus not sup- GR
ported by the seafloor morphology. To the east, the Pacific–Farallon
Ridge extended to the south connecting up with the Pacific–Antarctic
EB
Ridge at the Pacific–Antarctic–Farallon triple junction.
Antarctica
sociated ocean floor dominate the South Pacific (Fig. 13) and form a
crucial link in the global plate circuit. Early reconstructions of the
South Pacific recognized that spreading between the Pacific and the AT
Pacific
et al., 2004a,b). Rifting between the Chatham Rise and Antarctica/Marie SEIR
Byrd Land is believed to have occurred at 90 Ma (Larter et al., 2002; EM
Eagles et al., 2004a) with the initiation of spreading between the CP Australia
et al., 1995; Larter et al., 2002; Croon et al., 2008) or 80 Ma for EANT-AFR EANT-SAM EANT-IND
and 6C (~24 Ma) and at Chron 3a (~6 Ma) (Cande et al., 1995; Croon et a
al., 2008). A recent update of the seafloor spreading history between the
Pacific and Antarctic plates (Croon et al., 2008) is in general agreement 20˚ Arabia
with the model of Cande et al. (1995) for times 61 Ma to 12.3 Ma, but Africa
India
the model and rotations differ slightly for younger times. SR
10˚
To construct our seafloor spreading isochrons between the Pacific CR Sundaland
and Antarctic plates, we used the magnetic anomaly pick identifica-
0˚ Somalia
tions and finite rotations of Cande et al. (1995) for times from
61 Ma to the present day, which are also used in the model of WSB
MR WB
Eagles et al. (2004a,b). Croon et al. (2008) provide updated rotations −10˚ EFR
MB
for times younger than 12.3 Ma but they are not incorporated into our CIR A
model. As noted by Croon et al. (2008) the effect of using these rota- −20˚
Australia G
tions can be identified adjacent to the Campbell Plateau and clearly 10˚
CR Sundaland
reflect a clockwise change in spreading direction between Anomalies
31 (~68 Ma) and 25 (~56 Ma) consistent with our isochrons. 0˚ Somalia
WSB
MR WB
−10˚ EFR
fuse deformation in the East Indian Ocean (Demets et al., 1994; Royer
and Gordon, 1997; Weissel et al., 1980) (Figs. 1 and 14). Several smaller −40˚
Antarctica SEIR
plates exist along the East African margin associated with continental
AAB
rifting and diffuse deformation, including the proposed Nubian and
KP
Lake Victoria plates (Lemaux et al., 2002; Bird, 2003). Prior to Gondwa- −50˚
na break-up and the opening of the Indian Ocean, a now entirely van- LIP Plate Boundary
Isochrons
product
ished ocean basin, the Tethys Ocean, existed between Gondwana and Extinct Ridge
EANT-AFR AUS-AFR ARA-SOM
IND-AFR AFR-MAD END-EANT
Laurasia. The evidence for this ocean basin is primarily preserved in Non-
oceanic Tectonic Boundary AUS-EANT BB & OTH EANT-IND
the terranes and ophiolite complexes along southern Eurasia and the AUS-IND AUS-GIN/WB MAD-SEY
Karoo Volcanics during the early Jurassic (around 185–180 Ma) (Cox, Ridge where highly oblique, ultra-slow seafloor spreading is occur-
1992; Forster, 1975; Jourdan et al., 2005; Reeves, 2000; Storey et al., ring (Patriat and Ségoufin, 1988; Royer et al., 1988).
2001) (Fig. 14). The cessation of volcanism along the Karoo Rift led The final break-up of Gondwana continental blocks occurred with
to a seaward jump in the locus of rifting, initiating contemporaneous- the separation of Madagascar and India forming the Mascarene Basin.
ly between Africa and Antarctica in the Mozambique Basin and Riiser- Previous interpretations of the area suggest that rifting initiated in the
Larson Sea (Simpson et al., 1979; Marks and Tikku, 2001; Eagles and late Cretaceous (Norton and Sclater 1979; Masson 1984; (Bernard and
König, 2008) and Africa and Madagascar in the West Somali Basin Munschy, 2000) with the oldest magnetic anomaly identified being
(Smith and Hallam, 1970; Hankel, 1994) and either contemporane- Anomaly 34 (~84 Ma) or 33 (~79 Ma). A major change to NE–SW
ously or earlier between Africa and West Antarctica in the Weddell spreading is recorded in the fracture zones and magnetic lineations
Sea (Livermore and Hunter, 1996; König and Jokat, 2006). around Anomaly 31 (~68 Ma) (Bernard and Munschy, 2000). Part of
Separation between Africa and Antarctica/Madagascar forming the Mascarene Ridge jumped northward isolating the Seychelles micro-
the Mozambique Basin, Riiser-Larson Sea and West Somali Basin is continent (Masson, 1984). The model of Bernard and Munschy (2000)
believed to have initiated in the early–mid Jurassic supported by the suggests contemporaneous spreading between the easternmost part
stratigraphy and pre-rift structures along the conjugate margins of the Mascarene Basin and spreading to the north between the
(Smith and Hallam, 1970; Bunce and Molnar, 1977; Norton and Seychelles and Laxmi Ridge, implying a cessation of spreading in the
Sclater, 1979; Ségoufin and Patriat, 1980; Scrutton et al., 1981; Mascarene Basin as late as Anomaly 27 (~61 Ma). The oldest identified
Coffin and Rabinowitz, 1987; Lawver and Scotese, 1987; Reeves, magnetic lineation between the Seychelles and Laxmi Ridge in the East
2000; Müller et al., 2008b). The transition from continental rifting Somali and West Arabian Basin is Anomaly 28 (~63 Ma) (Masson, 1984;
to seafloor spreading is believed to have occurred either at Collier et al., 2008) based on the dating of syn-rift volcanics offshore
183–177 Ma based on Eagles and Konig (2008) full-fit reconstruction, from the Seychelles (Collier et al., 2008) or Anomaly 27 (~61 Ma)
170 Ma (Müller et al., 1997; Reeves and De Wit, 2000), 167 Ma (König (Chaubey et al., 1998) defining the initiation of spreading along the
and Jokat, 2006) or 165 Ma based on matching tectonic sequences in Carlsberg Ridge.
Africa and East Antarctica (Coffin and Rabinowitz, 1987; Livermore We have adopted a model for East Africa whereby pre-breakup
and Hunter, 1996; Marks and Tikku, 2001). Early full-fit reconstruc- margin extension was initiated at 180 Ma as a response to thermal
tions place Madagascar west of the Gunnerus Ridge (Royer and weakening by the eruption of the Karoo flood basalts. We initiate sea-
Coffin, 1992) whereas most recent studies place Madagascar to the floor spreading at 160 Ma along the entire East Africa margin after the
east (Marks and Tikku, 2001; Eagles and König, 2008) thereby elimi- cessation of rifting in the Karoo Rift, about 5 million years before the last
nating overlap issues between Antarctica and Madagascar. confidently dated magnetic anomaly, M25 (~154 Ma) (Fig. 15). We con-
The oldest identified magnetic anomalies interpreted in the nect the rift to the mid-ocean ridge that developed between Patagonia
Mozambique and West Somali Basins and Riiser-Larson Sea are and Southern Africa (Torsvik et al., 2009) and Weddell Sea to the south-
Anomalies M25–M24 (~ 154–152 Ma) (Ségoufin and Patriat, 1980; west and to a transform in the Tethys to the northeast (Figs. 14 and 15).
Rabinowitz et al., 1983; Coffin and Rabinowitz, 1987; Roeser et al., The identification of magnetic anomalies and fracture zone trends is dif-
1996; Marks and Tikku, 2001; Jokat et al., 2003). However, some ficult in the area due to thick sediment cover and volcanic overprinting.
have inferred Jurassic Quiet Zone crust between the oldest magnetic Weakly trending magnetic lineations observed in the magnetic anomaly
anomalies and the continental slope (Coffin and Rabinowitz, 1987) grid confirm the N–S directed spreading direction (Fig. 14). We adopt
possibly as old as M40 (~166 Ma) (Gaina et al., 2010). Spreading in the model for the cessation of spreading in the West Somali Basin short-
all basins was directed N–S for most of the opening history, confirmed ly after M0 (~120 Ma) and not at M10 (~131 Ma) as suggested by Eagles
through the interpretation of fracture zones (Heirtzler and Burroughs, and Konig (2008). The cessation of spreading at M10 (~131 Ma) results
1971), but a NNE–SSW direction can also be seen in the older oceanic in the position of Africa relative to Madagascar and Antarctica that is in-
crust fabric. Paleomagnetic (McElhinny et al., 1976), seismic and compatible with newly interpreted aeromagnetic data in the area
gravity anomaly data (e.g. (Rabinowitz, 1971; Bunce and Molnar, (König and Jokat, 2010). After the cessation of spreading, we implement
1977; Coffin and Rabinowitz, 1987, 1988; Storey et al., 1995) support a southward ridge jump towards the site of Madagascar Ridge and
the southward motion of Madagascar relative to Africa during the Ju- Conrad Rise eruption. Our model implies that Madagascar operated as
rassic and Early Cretaceous. an independent plate from 144 to 115 Ma, based on our interpretation
The spreading histories of the Mozambique/Riiser-Larson Sea and of the West Somali Basin. Spreading in the Mozambique/Riiser-Larson
the West Somali Basin diverge at about M10 (~130–132 Ma). Spread- Sea continued unabated throughout the Mesozoic and along the South-
ing in the West Somali Basin ceased either at M10 (~130–132 Ma) west Indian Ridge to the present day.
(Rabinowitz et al., 1983; Coffin and Rabinowitz, 1987; Eagles and Our model for the separation of Madagascar and India is similar to
König, 2008) or M0 (~ 120 Ma) (Ségoufin and Patriat, 1980; that presented in Masson (1984) and Müller et al. (1997). Although the
Cochran, 1988; Müller et al., 1997; Marks and Tikku, 2001; Müller oldest magnetic anomaly identified is Anomaly 34 (~84 Ma), we initiate
et al., 2008a) depending on the magnetic anomaly identification rifting at 87 Ma, preceded by a period of strike-slip motion between India
used. After the cessation of spreading, the mid-ocean ridge jumped and Madagascar. A major change in spreading direction occurred at
southward initiating spreading in between Madagascar and Antarcti- Anomaly 31 (~68 Ma) to NE–SW spreading based on an interpretation
ca. The timing of the southern ridge jump and seafloor spreading his- of the fracture zone trends in the basin. Spreading in the Mascarene
tory in the surrounding Enderby Basin and Weddell Sea has major Basin ceased at 64 Ma resulting in a northward ridge jump and initiation
implications for the plate boundary configurations in the Mesozoic In- of spreading between India and the Seychelles microcontinent forming
dian Ocean. For example, the model of Eagles and Konig (2008) infers the crust in the East Somali and West Arabian Basins. However, spreading
a southward ridge jump from the West Somali Basin at M10 (~ 130– may have continued to at least Anomaly 27 (~61 Ma) in the eastern
132 Ma) transferred Madagascar to the African plate and initiated Mascarene Basin (Bernard and Munschy, 2000). The spreading ridge be-
spreading in the Enderby Basin. In this model Madagascar did not tween the Seychelles to the south and Laxmi ridge to the north (Carlsberg
act as an independent plate throughout any of its Mesozoic–Cenozoic Ridge) is modeled based on triple junction closure with India and Arabia.
history. Other models propose that Madagascar must have acted in- The Carlsberg Ridge connected with the Central Indian Ridge to the
dependently, at least for part of its history (e.g. Marks and Tikku, southeast and the Sheba Ridge via a series of large offset transform faults
2001). The mid-ocean ridge which formed the Mesozoic magnetic to the northwest. The Sheba Ridge separates Arabia from Africa/Somalia,
lineations in the Mozambique Basin/Riiser-Larson Sea continued which we initiate at 20 Ma to coincide with the initiation of the East
throughout the Cenozoic eventually becoming the Southwest Indian African Rift. The Sheba Ridge propagated into the Red Sea at 15 Ma.
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 241
20˚ 20˚
40˚ 40˚
20˚ 40˚ 60˚ 20˚ 40˚ 60˚ 20˚ 40˚ 60˚ 20˚ 40˚ 60˚
120 Ma 120 Ma 80 Ma 80 Ma
0˚ 0˚
20˚ 20˚
40˚ 40˚
20˚ 40˚ 60˚ 20˚ 40˚ 60˚ 20˚ 40˚ 60˚ 20˚ 40˚ 60˚
40 Ma 40 Ma 0 Ma 0 Ma
0˚ 0˚
20˚ 20˚
40˚ 40˚
0 20 40 60 80 100 120 140 160 ANT WGD NWA SOM AFR IND NEA EGD NTY
Fig. 15. Agegrid reconstructions of the east African basins at 160, 140, 120, 80, 40, 0 Ma highlighting the age–area distribution of oceanic lithosphere at the time of formation and the
extent of continental crust (grey polygons) and reconstructions showing the outlines of the plates in the east African basins for each reconstruction time listed above. Plate boundaries
from our continuously closing plate polygon dataset are denoted as thick white lines, hotspot locations as yellow stars, large igneous provinces and flood basalts as brown polygons and
coastlines as thin black lines. Abbreviations are: AFR = African plate, ANT = Antarctic plate, EGD = east Gondwana plate, IND = Indian plate, NEA = northeast African plate, NTY =
Neo-Tethys plate, NWA = northwest African plate, SOM = Somali plate, WGD = west Gondwana plate.
3.3.2. Antarctic margin spreading is believed to have occurred at ~ 167 Ma (König and Jokat,
The Antarctic margin bordering the Indian Ocean involves at least 2006), 165 Ma (Livermore and Hunter, 1996; Marks and Tikku,
four distinct spreading phases, including (from west to east): the 2001) and 160 Ma (Ghidella et al., 2002; Müller et al., 2008a). The
Weddell Sea opening between West Antarctica and South America, M-series magnetic anomalies are difficult to identify but a recent
the Riiser-Larson Sea between Antarctica and Africa (conjugate to study by Konig and Jokat (2006) identified magnetic anomalies as
the Mozambique Basin), the Enderby Basin between Antarctica and old as M17 (~ 140 Ma) with seafloor spreading believed to have initi-
India/Elan Bank and the Southern Ocean between Antarctica and ated around M20 (~146 Ma), suggesting 15–20 Ma of rifting and
Australia (Figs. 13 and 14). continental stretching before the establishment of seafloor spreading.
The opening of the Weddell Sea is believed to have initiated as a Seafloor spreading was initially very slow, directed north–south (König
three-plate system between Antarctica, South America and Africa and Jokat, 2006). The Cenozoic magnetic anomalies are well-identified
(Marks and Tikku, 2001), or initially as a two-plate system with N–S (LaBrecque and Barker, 1981; Kovacs et al., 2002) eventually leading to
directed spreading between South America and Antarctica (Kovacs the establishment of the American–Antarctic Ridge (Fig. 13). Due to sub-
et al., 2002). The transition from seafloor spreading to incipient duction starting in the Cretaceous, the entire northern plate involved in
242 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
Weddell Sea spreading has been subducted including parts of the early seafloor spreading (Anomalies 34–27; ~ 84-61 Ma) via NW–SE
Cenozoic crust from the Antarctic (southern) plate. directed spreading. Spreading developed into a N–S configuration
Spreading in the Riiser-Larson Sea (conjugate to Mozambique and has continued to the present day with a dramatic increase in
Basin, west of the Gunnerus Ridge), has been dated with a well- spreading rate from Anomaly 13 (~33 Ma) (Tikku and Cande, 1999).
defined sequence from at least M24 (∼ 152–153 Ma) (Roeser et al., We adopt a model for the Antarctic margins, which suggests con-
1996; Jokat et al., 2003), although a recent reinterpretation of mag- temporaneous rifting in the Weddell Sea, Riiser-Larson Sea and the
netic anomalies suggests that magnetic anomalies as old as M40 East African margins starting in the late Jurassic, at 180 Ma, after the
(~166 Ma) exist in both the Riiser-Larson and Mozambique Basins cessation of Karoo volcanism and seaward jump in the locus of rifting.
(Gaina et al., 2010). The spreading system here continued into the We model the opening of the Weddell Sea based on Konig and Jokat
Cenozoic to the west and north of the Conrad Rise where Anomalies (2006), with M20 (~ 146 Ma) corresponding to the oldest oceanic
34 (∼83.5 Ma) to 28 (∼63 Ma) have been identified (Goslin and crust in the area. Comparison of our seafloor spreading isochrons
Schlich, 1976; Royer and Coffin, 1992). This spreading ridge developed with our magnetic anomaly compilation is difficult (Figs. 13 and 14)
into the ultra-slow Southwest Indian Ridge (Patriat and Ségoufin, due to the lack of data coverage and weak magnetic anomaly signa-
1988). tures. Spreading continued until the end of the CNS (83.5 Ma) when
East of the Gunnerus Ridge and west of the Bruce Rise lies the there was a reorganization of the spreading ridge system leading to
Enderby Basin (Fig. 13) recording the opening and seafloor spreading the establishment of spreading along the American–Antarctic Ridge.
history between Antarctica and India. The paucity of data in the area This ultra-slow spreading system is currently intersecting the Sand-
and the identification of magnetic anomaly sequences on the conju- wich subduction zone, one of the few regions of the world where an
gate Indian side in the Bay of Bengal and south of Sri Lanka have led active mid ocean ridge is intersecting a subduction zone. The Mesozoic
to two alternative theories for the break-up of Antarctica and India: Weddell spreading center connected with spreading in the Riiser-
(1) Break up and seafloor spreading during the CNS (Royer and Larson Sea/Mozambique Basin in a triple junction configuration.
Coffin, 1992; Banerjee et al., 1995; Müller et al., 2000; Jokat et al., Further east, we initiate rifting between Antarctica and India in
2010), or (2) Break-up and seafloor spreading in the Mesozoic at the Enderby Basin (central and eastern parts) at 160 Ma to coincide
135 Ma with the oldest identified magnetic anomaly being M11 with the initiation of rifting between Australia and Antarctic, which
(~132 Ma) (Ramana et al., 1994, 2001; Desa et al., 2006) or M9 has been well dated. We adopt the Mesozoic seafloor spreading
(~129 Ma) (Gaina et al., 2007). The model of Marks and Tikku model in Gaina et al. (2007) using the finite rotations that describe
(2001) tentatively identified anomalies M10Ny–M1 (~ 132–121 Ma) motion between Antarctica and the Elan Bank from Gaina et al.
in the West Enderby Basin, whereas the most recent model of Jokat (2003) for the central and eastern Enderby Basin. Here, seafloor
et al. (2010) for the West Enderby Basin suggests break-up between spreading initiated at 132 Ma with M9 (~ 129 Ma) corresponding to
India and Antarctica during the CNS (~90–118 Ma). the oldest identified magnetic anomaly. The initiation of spreading
The Mesozoic spreading model implies contemporaneous opening in the Enderby Basin results in strike-slip motion between India and
with the well-documented M-sequence anomalies (M10–M0; ~ 132– Madagascar of over 1000 km. A ridge jump isolating the Elan Bank
120 Ma) off the Perth Abyssal Plain (Powell et al., 1988; Müller et al., microcontinent occurred at 120 Ma coincident with the eruption of
1998a). The model of Gaina et al. (2007) further incorporates micro- the Kerguelen Plateau. For the Western Enderby Basin, we initiate
continent formation (Elan Bank) due to one or several ridge jumps as- break-up during the CNS at around 118 Ma, consistent with the
sociated with the Kerguelen Plume (Müller et al., 2000; Gaina et al., model of Jokat et al. (2010).
2003). We model a simple scenario for the rifting, break-up and seafloor
The area east of the Bruce Rise and Vincennes Fracture Zone and spreading history between Australia and Antarctica with rifting initi-
south of Australia involves rifting, break-up and seafloor spreading ating at 165 Ma based on the evidence presented in Totterdell et al.
between Antarctica and Australia forming the Southern Ocean (2000) and break-up at 99 Ma (Müller et al., 2000; Müller et al.,
(Figs. 13 and 14). The conjugate Australia and Antarctic margins con- 2008a). The rift boundary extended into the Enderby Basin from
sist of a wide zone of highly extended continental crust adjacent to a 165 Ma and extended eastward to connect with the Western
narrow zone of incipient oceanic crust formed by slow to ultra-slow Panthalassic subduction zone along eastern Australia. We incorporate
seafloor spreading. Continental rifting is believed to have initiated at the oldest magnetic anomaly as Anomaly 34 (~83.5 Ma) based on the
165 Ma based on the dating of syn-rift sedimentary sequences within model of Tikku and Cande (2000) with a N–S direction of spreading.
the Australian rift basins and increased tectonic subsidence rates We do not incorporate the NW–SE early separation motion of Australia
(Totterdell et al., 2000) or 160 Ma (Powell et al., 1988). However, and Antarctica (Whittaker et al., 2007) but anticipate that this will be
the nature of break-up and transition to true seafloor spreading incorporated in a future model. We use the rotations and magnetic
along the margin remains controversial (Tikku and Cande, 1999; anomaly identifications of Müller et al. (1997) for Anomalies 31–18
Sayers et al., 2001). The timing of break-up is inferred to be around (~68–40 Ma) and Royer and Chang (1991) from Anomaly 18
100 Ma based on the identification of seafloor spreading magnetic (~40 Ma) to the present day. Our resultant seafloor spreading isochrons
anomalies adjacent to the margin (Cande and Mutter, 1982) or by ex- match very well with the trends observed in our magnetic anomaly grid
trapolation of the spreading rate (Veevers et al., 1990), 135–125 Ma (Figs. 13 and 14).
based on the relationship between continental margin sequences
and the oceanic crust from seismic data (Stagg and Willcox, 1992) 3.3.3. West Australian margins
or 83.5 Ma based on the dating of the oldest magnetic anomaly The West Australian continental margin is an old, sediment-starved
(Tikku and Cande, 1999; Whittaker et al., 2007), depending on how volcanic continental margin, which formed as a result of multistage rift-
the crust in the transition zone is defined. The oldest magnetic anom- ing and seafloor-spreading during a late Paleozoic and early Mesozoic
aly that can be identified is Anomaly 34 (~84 Ma) (Cande and Mutter, phase of East Gondwana break-up (Baillie and Jacobson, 1995;
1982; Tikku and Cande, 1999; Whittaker et al., 2007) but Anomalies Bradshaw et al., 1988; Veevers, 1988). The area can be separated into
34 (~84 Ma) and 33 (~ 79 Ma) are located in a zone of transitional four distinct zones: the Argo Abyssal Plain, alongside the Browse and
crust (i.e. morphology not typical of abyssal hill fabric), therefore Roebuck (former offshore Canning Basin) basins, the Gascoyne Abyssal
Anomaly 32 (71 Ma) is often quoted as the oldest magnetic anomaly Plain, alongside the Exmouth Plateau and the Northern Carnarvon
to indicate true seafloor spreading. The direction of spreading has Basin, the Cuvier Abyssal Plain delimited by the Cape Range Fracture
previously been modeled as N–S, however a recent reanalysis of grav- Zone (CRFZ) and Wallaby-Zenith Fracture Zones (WZFZ), and includes
ity and magnetic anomaly profiles (Whittaker et al., 2007) suggests the Southern Carnarvon Basin, the Exmouth Sub-basin and the Wallaby
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 243
and Zenith plateaus and the Perth Abyssal Plain extending from the
150 Ma 150 Ma
WZFZ to the Naturaliste Plateau in the south (Fig. 14). 20˚
Rifting in the Argo Abyssal Plain started around 230 Ma (e.g.
Müller et al., 2005) eventually leading to the separation of the West
Burma block/Argoland from the Australian continental margin. The
transition from rifting to seafloor spreading has been constrained by 40˚
the dating of magnetic anomalies in the Argo Abyssal Plain and
through tectonic subsidence analysis along the margin. The interpre-
tation of magnetic lineations resolves that seafloor spreading initiated 60˚ 80˚ 100˚ 120˚ 60˚ 80 ˚ 100˚ 120˚
immediately prior to Anomaly M26 (~ 155 Ma) (Fullerton et al., 1989; 130 Ma 130 Ma
Sager et al., 1992; Müller et al., 1998a; Heine and Müller, 2005) with 20˚
NW–SE directed spreading. Previous models have invoked a south-
ward propagating ridge along the Western Australian margin, which
started in the Argo Abyssal Plain progressing southward. Spreading
40˚
in the Gascoyne and Cuvier Abyssal Plains initiated at M10
(~132 Ma) (Johnson et al., 1976, 1980; Larson, 1977; Falvey and
Mutter, 1981; Powell et al., 1988; Fullerton et al., 1989; Sager et
60˚ 80˚ 100˚ 120˚ 60˚ 80˚ 100˚ 120˚
al., 1992; Müller et al., 1998a) and marked the break-up between
Australia and Greater India. The model for the opening of the Argo 100 Ma 100 Ma
Abyssal Plain presented in Heine and Müller (2005) differs from pre- 20˚
vious models and that of Robb et al. (2005) in that spreading between
the Argo and Gascoyne Abyssal Plains initiated almost simultaneously
with the same orientation. The model also invoked a landward ridge 40˚
jump at M13 (~ 136 Ma). Further southward, spreading in the Perth
Abyssal Plain which records break-up between Australia and India oc-
curred around 132 Ma based on the mapping of magnetic anomalies 60˚ 80˚ 100˚ 120˚ 60˚ 80˚ 100 ˚ 120 ˚
(Veevers et al., 1985; Müller et al., 1998a) and involved several sea-
80 Ma 80 Ma
ward ridge jumps towards the Kerguelan plume (Müller et al.,
20˚
2000). However, the majority of the crust may have formed during
the CNS.
We adopt the model for the formation of the Argo and Gascoyne
Abyssal Plains following Heine and Müller (2005) which involves 40˚
NW–SE oriented rifting of West Burma from the northwestern mar-
gin of Australia at around 156 Ma (Fig. 16). The continent–ocean
boundary along Australia's western margin is from Heine and 60˚ 80˚ 100˚ 120˚ 60˚ 80˚ 100˚ 120˚
Müller (2005). Spreading continued until a landward ridge jump at 50 Ma 50 Ma
M13 (~136 Ma). We infer that the plate boundary connected with a 20˚
Tethyan spreading ridge located to the north of India/Greater India
to the west and a transform fault to the north (Fig. 16). Our model in-
vokes a southward propagating ridge into the Cuvier and Perth Abys-
40˚
sal Plain at 132 Ma following the models presented in Müller et al.
(1998a) and Müller et al. (2000). The mid-ocean ridge associated
with spreading in the Perth Abyssal Plain formed a triple junction
60˚ 80˚ 100˚ 120˚ 60˚ 80˚ 100˚ 120˚
with mid-ocean ridge opening the Enderby Basin (between East Ant-
arctica and India) (e.g. Gaina et al., 2007) and the Australia–Antarctic 0 Ma 0 Ma
mid-ocean ridge (Fig. 16). The NW-–SE directed spreading along the 20˚
Western Australian margin persisted until around 99 Ma. The fracture
zones record a dramatic change in trend from NW–SE to roughly N–S
at around 99 Ma (Müller et al., 1998a). The change to N–S spreading 40˚
forms the oldest crust associated with the Wharton Ridge/Wharton
Basin. Seafloor spreading in the Wharton Basin ceased at 43 Ma
(Singh et al., 2010).
ANT AUS CAP NEA EUR IND JUN EGD NTY NJU
3.3.4. Tethys Ocean 0 20 40 60 80 100 120 140 160 ANT AUS CAP SOM EUR IND JUN EGD NTY NJU
The Tethys Ocean represents a now largely subducted ocean basin Age of Oceanic Lithosphere (m.yrs) Plate Name
that existed between Gondwanaland and Laurasia and involves a histo-
Fig. 16. (left) Agegrid reconstructions of the west Australian margin at 150, 130, 100, 80, 50,
ry of successive continental rifting events along the northern Gondwana
0 Ma highlighting the age–area distribution of oceanic lithosphere at the time of formation
margin, oceanic basin formation and accretion of Gondwana-derived and the extent of continental crust (gray polygons). Plate boundaries from our continuously
continental blocks onto the southern Laurasian margin and Indochina/ closing plate polygon dataset are denoted as thick white lines, hotspot locations as yellow
SE Asia. The majority of Tethyan oceanic crust no longer exists due to stars, large igneous provinces and flood basalts as brown polygons and coastlines as thin
long-lived subduction along the southern Eurasian margin, except in black lines. (right) Reconstructions showing the outlines of the plates in the west Australian
margins for each reconstruction time listed above. Feature descriptions as in panel (left).
the Argo Abyssal Plain off NW Australia where a fragment of in-situ Abbreviations are: ANT = Antarctic plate, AUS = Australian plate, CAP = Capricorn plate,
oceanic crust recording the youngest Tethyan spreading system is pre- EGD = east Gondwana plate, EUR = Eurasian plate, IND = Indian plate, JUN = Junction
served (Fullerton et al., 1989; Heine and Müller, 2005). In addition, the plate, NEA = northeast African plate, NJU = north Junction plate, NTY = Neo-Tethys
Ionian Sea and several basins in the eastern Mediterranean (e.g. Levant plate, NWA = northwest African plate, SOM = Somali plate. (For interpretation of the ref-
erences to color in this figure legend, the reader is referred to the web version of this article.)
Basin) may be floored by Mesozoic Tethyan oceanic crust (Stampfli and
244 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
Borel, 2002; Müller et al., 2008b), however identification of magnetic Hassanzadeh et al., 2008), the collision of Sibumasu/Malaya to Indo-
anomalies is difficult. The limited amount of preserved in-situ oceanic china by 250–220 Ma (Metcalfe, 1999; Stampfli and Borel, 2002;
crust of Tethyan origin hampers our knowledge and understanding of Golonka, 2007) and 200–160 Ma for other elements including Qiang-
the evolution and structure of the Tethys Ocean. Instead we primarily tang (North Tibet) and Helmand (Stampfli and Borel, 2002). The ac-
rely on the accreted terranes and sutures in SE Asia, southern Eurasia, cretion of South Tibet varies from 200 to 160 Ma (Stampfli and
Arabia and throughout the Mediterranean and southern and central Borel, 2002), 150 Ma (Golonka et al., 2006) and 120 Ma related to a
Europe (e.g. Sengor, 1987; Metcalfe, 1996; Stampfli and Borel, 2002) separate episode of accretion (Metcalfe, 1996).
as they record the timing of continental block collision, ophiolite em- Following closure of the paleo-Tethys and accretion of the Cimme-
placement, back-arc basin development and provide paleo-latitudinal rian terrane, several back-arc basins opened as a response to slab-pull
estimates of continental material derived from the northern Gondwana forces along the Tethyan subduction zone. The major back-arc
margin. complexes include the Pindos, Maliac, Meliata, Küre, Sangpan, Kudi,
Successive rifting events from the Gondwana margin have led to the Vardar (Stampfli and Borel, 2002) and the early Cretaceous Taurus,
subdivision of the Tethys Ocean into several oceanic domains: the Troodos, Hatay and Baer-Bassit ophiolite complexes (Whitechurch
paleo- and neo-Tethys (e.g. Stampfli and Borel, 2002) or the paleo-, et al., 1984). The closure of these back-arc basins varied along the
meso- and neo-Tethys (Metcalfe, 1996; Heine et al., 2004) (Fig. 3a–d). margin from Triassic to Cenozoic (Stampfli and Borel, 2002), with a
The additional subdivision by Metcalfe (1996) and Heine et al. (2004) date of ~ 70–65 Ma for the obduction of the Taurus, Troodos, Hatay
stems from an alternative rift history for crust that formed after the and Baer-Bassit ophiolite complexes (Whitechurch et al., 1984) and
paleo-Tethys, which affects whether the Argo Abyssal Plain is classified an early Cenozoic age of obduction for the Pindos and Vardar back-
as part of the Tethys or Indian Ocean domains. arc basins (Stampfli and Borel, 2002). As these back-arc basins
The paleo-Tethys formed after the initiation of rifting and seafloor opened and closed, inferred NE–SW directed spreading continued in
spreading between the European and Asian Hunic superterrane (e.g. the meso-Tethys (or neo-Tethys ocean) orthogonal to the Gondwana
North China, Indochina, Tarim, Serindia, Bohemia) and the northern rifted margin (Stampfli and Borel, 2002). The cessation of spreading in
Gondwana margin (Metcalfe, 1996; Stampfli and Borel, 2002; the meso- or neo-Tethys is difficult to ascertain. However, Stampfli
Blakely, 2008). The timing of passive margin formation is dependent and Borel (2002) postulate that the subduction of the mid-ocean ridge
on the margin segment and ranges from Ordovician/Silurian based diachronously across the margin can be tied to the initiation of rifting
on subsidence analysis in the western Tethys (Stampfli, 2000; of the Argoland Block/West Burma from the northwest shelf of Australia,
Stampfli and Borel, 2002) or the late/early Devonian based on the thus timing the cessation of spreading in the meso-/neo-Tethys ocean.
Gondwana affinity of Devonian vertebrate faunas in the Hun super- A third phase of rifting along the northern Gondwana margin in the
terrane (Metcalfe, 1996), Devonian to Triassic passive margin se- northwest Australian shelf initiated in the late Triassic (Müller et al.,
quences along the southern margin of South China (Metcalfe, 1996) 2005). The models of Metcalfe (1996) and Heine et al. (2004) label the
and the dating of oceanic deep-marine ribbon bedded cherts in the resultant ocean basin as the neo-Tethys as their models extend the
Chang-Rai region of Thailand (Sashida et al., 1993; Metcalfe, 1996). Argo Abyssal Plain mid-ocean ridge north of Greater India. Hence, this
The direction of spreading is uncertain due to the lack of in-situ pre- ocean basin forms part of the Tethys ocean domain. However, most
served crust, however the seafloor spreading model of Stampfli and other studies associate the Argo Abyssal Plain with the Indian Ocean be-
Borel (2002) invokes NE–SW directed spreading orthogonal to the in- cause they follow the Argo spreading ridge southward between India
ferred margin. The passage of the Hunic superterrane from south to and Australia, thus representing earliest Indian Ocean spreading. The
north was facilitated by northward-dipping subduction along the preserved seafloor spreading record in the Argo Abyssal Plain confirms
southern Eurasian margin. The Hunic superterrane accreted to the that spreading initiated around 156 Ma leading to the separation of the
southern Laurasian margin diachronously in the Carboniferous– West Burma Block from the northwest Australian margin (Heine and
Permian (Stampfli and Borel, 2002). The cessation of spreading in Müller, 2005). The model of Metcalfe (1996) suggests that Lhasa
the paleo-Tethys is difficult to establish, however most modelers (South Tibet) also rifted off the northern margin of Greater India at the
agree the paleo-Tethys spreading ridge jumped southward along time. Spreading in the Argo Abyssal Plain is described in the Indian
the northern Gondwana margin and initiated the rifting of a new con- Ocean section of this paper. The West Burma Block was carried north-
tinental sliver from the Gondwana margin (e.g. Metcalfe, 1996; ward due to continuing subduction along the northern Tethyan margin
Stampfli and Borel, 2002; Blakely, 2008) after the accretion of the and sutured to Sibumasu in the Cretaceous around 80 Ma (Lee and
Hunic superterrane. Lawver, 1995; Metcalfe, 1996; Heine and Müller, 2005).
The second main phase of rifting isolated the Cimmerian terrane The termination of spreading in the Tethys Ocean is controversial.
from the Gondwana margin some time in the Pennsylvanian–early The model of Stampfli and Borel (2002) suggests cessation of spread-
Permian (Metcalfe, 1996; Stampfli and Borel, 2002), constrained by ing in the early Cretaceous when the meso- or neo-Tethys spreading
changes in biota (Shi and Archbold, 1998) and evidence of rifting on ridge intersected the Tethyan subduction zone. However, other
the northwest shelf of Australia (Falvey and Mutter, 1981; Müller et models (Metcalfe, 1996; Heine et al., 2004; Heine and Müller, 2005)
al., 2005), northern Pakistan and Afghanistan (Boulin, 1988; Pogue suggest that neo-Tethyan spreading continued through the Creta-
et al., 1992) and Iran (Stocklin, 1974). The Cimmerian terrane com- ceous, merging into the Wharton Basin spreading ridge from the
prises elements including Sibumasu (Sino-Burma-Malaya-Sumatra end of the CNS to 43 Ma (Heine et al., 2004). The final closure of the
continental sliver), Qiangtang (North Tibet), Helmand (Afghanistan), Tethys Ocean started with the collision of Greater India to the southern
Iran and possibly Lhasa/South Tibet (Fig. 18a). The ocean basin that Eurasian margin either around 55 Ma (Lee and Lawver, 1995) or 35 Ma
formed between the Gondwana margin to the south and the Cimme- (Van der Voo et al., 1999b; Hafkenscheid et al., 2001; Aitchison et al.,
rian terrane is labeled as the meso-Tethys in the models of Metcalfe 2007) marked by the Indus-Tsangpo Suture zone and ended with the
(1996) and Heine et al. (2004) but the neo-Tethys for most other closure of the Tethyan seaway between Arabia and Iran forming the
models. Continued northward-dipping subduction of paleo-Tethys Zagros Mountains (Hessami et al., 2001). Several fragments of Tethyan
oceanic lithosphere along southern Laurasia carried the Cimmerian ocean floor are postulated to underlay some of the basins in the eastern
terrane northward, leading to its accretion and closure of the paleo- Mediterranean (see Müller et al., 2008a).
Tethys ocean starting in the late Triassic (Metcalfe, 1996; Stampfli In the Mediterranean region, several Cenozoic back-arc basins formed
and Borel, 2002; Golonka et al., 2006; Blakely, 2008). Accretion is due to the convergence between Eurasia and Africa (Rosenbaum et al.,
constrained by the Cimmerian orogeny in present-day Iran, which 2002). The Liguro–Provençal basin opened from around early Oligocene
initiated in the late Triassic (Sengor, 1987; Stampfli and Borel, 2002; (~35 Ma) due to the eastward rollback of Apennines subduction (e.g.
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 245
Carminati et al., 2004) and the rotation of Corscia and Sardinia (Speranza in the eastern Mediterranean (e.g. Levant basin and Ionian Sea) repre-
et al., 2002) and the accretion of the Kabylies blocks to the African mar- sent the oldest preserved in-situ ocean floor, ranging in age from about
gin (e.g. Rosenbaum et al., 2002). Additional extensional basins such as 270 Ma (Late Permian) to 230 Ma (Middle Triassic) according to our
the Pannonian basin were associated with Africa–Eurasia collision and model. The Cenozoic basins in the western Mediterranean (e.g. Liguro–
associated with the Carpathian, Ionian and Hellenic subduction zones Provençal Basin) are reconstructed based on the tectonic model and
(Faccenna et al., 2001). rotations from Speranza et al. (2002), describing a Miocene counter-
Our model for the evolution of the Tethys Ocean closely follows clockwise rotation of Corsica–Sardinia relative to Iberia and France,
that of Heine et al. (2004), which is largely based on Stampfli and thereby creating accommodation space for back-arc opening.
Borel (2002) except in the Jurassic–Cretaceous. We agree with the
separation of the Tethys into three oceanic domains, as first suggested 3.4. Marginal and back-arc basins
by Metcalfe (1996) and adopted by Heine et al. (2004). We define the
paleo-Tethys as the ocean basin that formed after the separation of The present day distribution of the continents and oceans includes
the Hunic superterrane from the northern Gondwana margin, the many smaller ocean basins that formed either in a back-arc setting
meso-Tethys as the ocean basin that formed after the separation of behind a retreating subduction zone (Karig, 1971; Sleep and Toksoz,
the Cimmerian terrane from the northern Gondwana margin and 1971; Uyeda and Kanamori, 1979; Taylor and Karner, 1983;
the neo-Tethys as the ocean basin that formed when West Burma/ Faccenna et al., 2001; Sdrolias and Müller, 2006) or as a result of con-
Argoland separated from northwest Australia. Finite rotations de- tinental rifting without the influence of a subduction zone forming
scribing the opening of all three basins as well as associated seafloor marginal seas. The presence of ophiolites embedded within accreted
spreading isochrons are mostly derived by following the model of terranes provides evidence for the opening and closing of marginal
Stampfli and Borel (2002) and Heine et al. (2004). seas and back-arc basins in the past, most notably along the Tethyan
We follow a Devonian opening model for the paleo-Tethys (Metcalfe, margin and in the western North American margin. We have modeled
1996) but do not discount that opening may have been diachronous and some of the major marginal and back-arc basins observed in the sea-
occurred as early as the Silurian (Stampfli and Borel, 2002) in the west- floor spreading record today. We have also modeled the opening of
ern Tethys. As the reconstructions presented in this paper do not extend three critical marginal and back-arc basins that existed in the past
beyond 200 Ma, we will not describe the accretionary history of the but have been subsequently destroyed. These include the Mongol–
Hunic superterrane. We agree with Stampfli and Borel (2002) that the Okhotsk Ocean in Central Asia, the marginal basins that formed in
cessation of spreading in the paleo-Tethys led to southern ridge jump, the Caribbean, off the coast of western North America and the
initiating opening of the meso-Tethys around 280 Ma, coincident with proto-South China Sea. We also model the opening of the Caribbean,
the collision of the Hunic terrane to the southern Laurasian margin which includes a combination of marginal seas and back-arc basins.
and the initiation of rifting of the Cimmerian terrane from the northern
Gondwana margin in the early–mid Permian (Metcalfe, 1996). We in- 3.4.1. Caribbean
voke NE–SW directed spreading for the meso-Tethys consistent with The Caribbean resides between the North American and South
Stampfli and Borel (2002). The accretion of the Cimmerian terrane to American plates and contains Jurassic–Cretaceous ocean floor in the
the southern Laurasian margin also marks the closure of the paleo- Gulf of Mexico and Venezuela Basin, Cenozoic ocean basins such as
Tethys ocean. We broadly follow the timing of accretion based on the Cayman Trough, Gernada and Yucatan Basins, numerous conti-
Golonka et al. (2006) and Golonka (2007). The uncertainty in the south- nental blocks, accreted terranes, volcanic arcs and the Caribbean
ern extent of the Laurasian margin means that the timing of accretion Large Igneous Province (CLIP) (Figs. 3 and 9). The sedimentary basins
may change significantly depending on the southern extent of the surrounding the Gulf of Mexico are some of the world's most produc-
Laurasian continental margin. Following the closure of the paleo- tive hydrocarbon bearing basins, prompting quite detailed studies of
Tethys, a margin-wide episode of back-arc opening occurred along the the tectonic evolution of the region (Pindell, 1987; Burke, 1988;
southern Eurasian margin—from China to western Europe. This back- Ross and Scotese, 1988; Pindell and Kennan, 2009). The development
arc system was responsible for the crust that now forms part of the of the Caribbean is tied to break-up of Pangea and rifting in the
Cretaceous aged ophiolite complexes through southern Europe, Cyprus Central Atlantic, which extended into the Caribbean during the Triassic
(Troodos), Iran and Oman. Although these basins are known to have to earliest Cretaceous. This early phase formed rift basins, stretched
existed after the closure of the paleo-Tethys, we do not include their for- continental crust and salt basins in areas such as the South Florida
mation (e.g. Whitechurch et al., 1984; Robertson, 2000; Stampfli and Basin, Great Bank of the Bahamas, Yucatan and along northern South
Borel, 2002) as we focused on the broad-scale development of the Te- America (Pindell and Kennan, 2009). To the west, a continuous subduc-
thys Ocean. However, these back-arc basins have played a vital role in tion zone along the eastern margin on Panthalassa was consuming oce-
the development of the region and we anticipate that a thorough re- anic lithosphere beneath the western margin of the proto-Caribbean/
view of ophiolite complexes and back-arc basin correlatives will be in- trans-American region.
cluded in the next generation of the plate motion model. The Gulf of Mexico is bounded by predominately Triassic–Jurassic
Our model invokes continuous seafloor spreading in the meso-Tethys syn-rift structures and salt bearing basins and is partly floored by
from 280 Ma to 145–140 Ma. The neo-Tethys ocean forms with rifting Jurassic–Cretaceous oceanic crust. The timing of seafloor spreading
and seafloor spreading in the Argo Abyssal Plain, following the model of in the Gulf of Mexico is not well constrained with ages ranging from
Heine and Müller (2005), isolating the West Burma Block from the Gond- 158 to 170 Ma based on the timing of salt deposition and regional
wana margin. We initiate seafloor spreading at 156 Ma and extend the changes in structural trend and block rotations (Buffler and Sawyer,
mid-ocean ridge westward, north of Greater India where it intersects 1983; Ross and Scotese, 1988; Pindell and Kennan, 2009). The cessa-
with a Tethyan transform fault. The accretion of West Burma to Sibumasu tion of extensional faulting in the SE Gulf of Mexico and the dating of
occurred at 80 Ma, following Heine and Müller (2005). Seafloor spreading a post-rift unconformity (Ross and Scotese, 1988; Marton and Buffler,
in the neo-Tethyan ocean continued unabated eventually transforming 1999; Pindell and Kennan, 2009), places the cessation of seafloor
into the Wharton basin spreading ridge system in the eastern Indian spreading in the latest Jurassic-earliest Cretaceous between 145 and
Ocean until 43 Ma (Singh et al., 2010). 135 Ma. The opening of the Gulf of Mexico led to a two-stage anti-
In the western Mediterranean, we reconstruct the continental clockwise rotation of the Yucatan Block away from North America
blocks that comprise southern Europe and the Middle East in the into its present day location (Pindell and Kennan, 2009).
same manner as in Müller et al. (2008a). The basins floored by ocean- The existence of a proto-Caribbean Basin has been hypothesized
ic crust in the Mediterranean fall into two types. The Mesozoic basins based on the accommodation space created by the relative motion
246 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
between the North and South American plates. The development of (van der Lelij et al., 2010). The new tectonic regime led to opening of
this basin (its orientation and timing) is therefore purely dependent the Yucatan and Grenada–Tobago Basins in the Paleogene, Cayman
on the chosen plate tectonic model. Opening of the basin was either Trough since the Eocene (Ross and Scotese, 1988; Pindell and Kennan,
coincident with spreading in the Gulf of Mexico (Meschede and Frisch, 2009) and the Puerto Rico Basin in the Oligocene (Ross and Scotese,
1998; Pindell and Kennan, 2009) or initiated only after a southward 1988).
ridge jump in the early Cretaceous (Ross and Scotese, 1988). Models The Yucatan Basin currently resides between Cuba and the
that propose the encroachment of proto-Pacific oceanic lithosphere Cayman Ridge and is believed to have formed prior to the collision
into the Caribbean (e.g. Ross and Scotese, 1988; Pindell and Kennan, of the Caribbean Arc as a passive response to the northwestward roll-
2009) imply that all evidence of the proto-Caribbean Basin was sub- back of the trench (Pindell et al., 2006). The cessation of spreading is
ducted by the late-Cretaceous–early Cenozoic, whereas models that do correlated with the docking of the arc terranes along Cuba and the
not invoke an advancing trench relate NE–SW trending magnetic Bahaman Platform. The Grenada–Tobago Basin formed as a back-arc
lineations in the Venezuela Basin (Ghosh et al., 1984) to the proto- between the Aves Ridge and Lesser Antilles Ridge due to the eastward
Caribbean Basin (Meschede and Frisch, 1998). rollback of the Lesser Antilles Trench. The timing of spreading is
One of the major features that controlled the broad-scale develop- unconstrained by magnetic anomaly interpretations but initiation is
ment of the Caribbean is the nature of the plate boundary between believed to have occurred sometime in the Paleogene based on the
the Caribbean and Panthalassa/Pacific Ocean. Most models agree cessation of plutonism on the Aves Ridge (Pindell et al., 1988) and
that east-dipping trans-America subduction was consuming proto- from seismic stratigraphy and heatflow measurements within the
Pacific oceanic lithosphere during the Triassic–Cretaceous (Ross and basin (Speed, 1985; Pindell and Kennan, 2009). Spreading is believed
Scotese, 1988; Meschede and Frisch, 1998; Pindell and Kennan, to have ceased in the Oligocene coincident with the collision of the Less-
2009). However, models subsequently diverge into either “Pacific or- er Antilles forearc with the Venezuelan margin (Pindell and Kennan,
igin” (Malfait and Dinkelman, 1972; Burke, 1988; Ross and Scotese, 2009). The Cayman Trough formed as a left-lateral pull-apart basin be-
1988; Pindell and Kennan, 2009) or “intra-American origin” scenarios tween two major transform faults starting at Chron 19 (~41 Ma)
(Meschede and Frisch, 1998; James, 2006). “Pacific-origin” scenarios (Rosencrantz et al., 1988; Ross and Scotese, 1988) based on the inter-
propose a switch in the polarity of the trans-American plate boundary pretation of magnetic anomalies. The Puerto Rico Basin opened in the
from east-dipping to southwest-dipping in the late Cretaceous along Oligocene–early Miocene as a result of relative motion between Hispan-
the Caribbean/Greater Antilles Arc, causing the subduction of the iola and the Caribbean plate (Ross and Scotese, 1988).
proto-Caribbean Basin and encroachment of oceanic lithosphere Our model largely follows the hierarchical model of Ross and
from the Pacific domain into the Caribbean. The timing of this polarity Scotese (1988) (with an updated timescale) and elements of
flip is believed to be around 100–90 Ma (Ross and Scotese, 1988; Pindell and Kennan (2009), with minor adjustments based on recent
Pindell and Kennan, 2009) and constrained to 90 Ma in the south on geological information and an updated spreading model in the
Aruba and within the Bonaire Block (van der Lelij et al., 2010). Con- Central and Equatorial Atlantic. Rifting in the Caribbean since the
tinued northeastward rollback of the subduction hinge eventually Triassic connected to the Central Atlantic rift zone through Florida
caused collision with Yucatan and accretion of the arc along the and Gulf of Mexico and extended westward to the trans-America
Bahamas Platform. In the model of Ross and Scotese (1988) this ac- subduction zone, which was actively consuming Panthalassic ocean
cretion led to a jump in the locus of subduction westward, initiating floor. In our model, we follow the initiation of spreading in the Gulf
subduction along the Panama–Costa Rica Arc around 60 Ma. Howev- of Mexico at 170 Ma based on Ross and Scotese (1988) coincident
er, other models place the initiation of Panama–Costa Rica Arc to with accelerated seafloor spreading rates in the Central Atlantic
80–88 Ma (Pindell and Kennan, 2009) before the accretion of the (Labails et al., 2010) (Fig. 8). We update the cessation of spreading
Caribbean Arc to the Bahamas Platform. Recent tectonostratigraphic to 145 Ma based on evidence presented in Pindell and Kennan
and geochemical data from exposed rocks in southern Costa Rica (2009). After the cessation of spreading in the Gulf of Mexico, we
and western Panama indicate protoarc initiation on top of CLIP base- model a ridge jump to the south initiating the opening of the
ment occurred between 75 and 73 Ma (Buchs et al., 2010). Irrespec- proto-Caribbean Basin within the accomodation space created by
tive of timing, in the “Pacific origin” model, the initiation of the the relative motion between the North and South American plates
Panama–Costa Rica Arc trapped Pacific-derived oceanic lithosphere (Fig. 8). Spreading was NW–SE directed and initiated around
(now underlying the Venezuela Basin) as well as the CLIP onto the 145 Ma forming a triple junction to the east between the mid
Caribbean plate. “Intra-American origin” models assume a continuous ocean ridge of the Central Atlantic and rift axis of the Equatorial/
trans-America east-dipping subduction zone, which provided a per- South Atlantic. To the west, the mid ocean ridge of the proto-
manent barrier between the Pacific/Panthalassa and Caribbean. Con- Caribbean Basin formed a ridge–ridge–transform triple junction
currently, southwest-dipping subduction to the east of the proto- with the spreading ridge of the Andean back-arc basin and the
Caribbean Basin led to the docking of tectonic elements along the trans-American subduction zone.
Bahaman Platform. In the “intra-American” model, the origin of the We favor the “Pacific-origin” model for the formation of the
oceanic lithosphere underlying the Venezuela Basin and the CLIP are Caribbean plate with a subduction polarity flip of the trans-America
both derived in-situ. This model implies that the Panama–Costa Rica subduction zone to west-dipping along the eastern boundary of the
Arc was built upon a much older arc sequence. Caribbean Arc at 100 Ma (Fig. 8). The rollback of this subduction
After ~60 Ma, most models for the Caribbean are largely similar on a zone led to the consumption of the actively spreading proto-
broad scale. After the establishment of subduction along the Panama– Caribbean ocean floor and encroachment of the Farallon plate into
Costa Rica Arc, the Caribbean plate became a stationary feature the Caribbean domain. Our model predicts that the oceanic litho-
influenced only by the relative motions between the North and South sphere intruding into the Caribbean (and currently underlying the
American plates (Ross and Scotese, 1988). The southern margin of the Venezuela Basin) formed along the Pacific–Farallon ridge between
Bahaman platform changed from convergence to sinistral strike-slip Chrons M16-M4 (~139–127 Ma) at a latitude of around 10–15°S,
after the accretion of arc terranes with E–W transform faults dominat- agreeing well with paleomagnetic constraints, which suggest an
ing the region. To the east, west-dipping subduction and arc volcanism equatorial formation for the oceanic crust of the Nicoya Complex
along the Aves Ridge were still occurring. To the south, thermochrono- (Duncan and Hargraves, 1984). The continued roll-back of the
logical and sedimentological analyses suggest that the Bonaire Block Caribbean Arc subduction zone led to the formation of the Yucatan
collided with the South American margin at ~50 Ma thereby constrain- Basin as a back-arc in the late Cretaceous with cessation of spreading
ing the change from convergence to strike-slip along South America occurring at 70 Ma when the Caribbean Arc accreted to the Bahaman
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 247
Platform. The accretion led to a jump in the locus of subduction west- Golonka et al., 2006; Golonka, 2007) based on collision followed by
ward along the newly developed Panama–Costa Rica to accommodate folding and intrusion of granitic batholiths in Mongolia and the trans-
the continued eastward motion of the Farallon plate, trapping Faral- Baikal area (Golonka et al., 2006) and the formation of the Mongol–
lon oceanic lithosphere onto the Caribbean plate in the process. The Okhotsk Suture (Tomurtogoo et al., 2005). Complete closure may have
eruption of the Caribbean flood basalt province occurred around ended as late as the early Cretaceous (Zorin, 1999) based on the cessa-
90 Ma on top of the oceanic lithosphere that now underlies much of tion of compression in the area (Zorin, 1999). Alternative models exist
the Caribbean ocean floor (Sinton et al., 1998). The Caribbean flood that predict an older initial closure age of late Carboniferous (Badarch
basalt province (or CLIP) has been suggested to be the product of et al., 2002; Cocks and Torsvik, 2007), but again, this may be due to a dif-
the Galapagos hotspot (Pindell and Kennan, 2009), however in our ference in the definition of the Mongol–Okhotsk Ocean.
model the CLIP erupted on Farallon oceanic lithosphere over We have modeled the opening of the Mongol–Okhotsk Basin in
2000 km away from the present day position of the Galapagos hot- the late Carboniferous to account for the zircon data of Tomurtogoo
spot precluding this as a source, even assuming the motion of hot- et al. (2005), followed by the onset of subduction along the Siberian
spots relative to each other (Fig. 10). margin in the late Permian. We continue seafloor spreading in the
Coincident with subduction along the proto-middle America Mongol–Okhotsk Basin until the Permo-Triassic boundary (250 Ma).
trench was west-dipping subduction to the east along the Aves/Lesser Based on our initiation and termination of spreading, we suggest
Antilles Ridge, consuming Atlantic Ocean floor (Fig. 8). The rollback of that the Mongol–Okhotsk Ocean had a maximum width of about
this subduction zone led to the formation of the Grenada Basin be- 4000 km. We model the closure of the Mongol–Okhotsk Basin to
tween the Aves and Lesser Antilles Arcs in the Paleogene. In the mid- 150 Ma (late Jurassic) based on the overwhelming evidence in the lit-
dle Eocene (41 Ma), relative motion between North America and erature for the dating of the Mongol–Okhotsk Suture.
Caribbean began to form the Cayman Trough along sinistral faults
that later merge with the Lesser Antilles trench. In the early Miocene 3.4.3. North American margins
(20 Ma), the Cayman Trough continued to expand and develop, and The western North American margin is characterized by the accre-
the Chortis Block moved over the Yucatan promontory. Westward tion of native and exotic terranes throughout the late Paleozoic and
motion of the North American plate relative to the slow moving Mesozoic. The timing of formation of the numerous terranes with is-
Caribbean plate was accommodating the opening of the Cayman land arc affinities, their accretion onto the continental margin and
Trough. The Puerto Rico Basin formed in the Oligocene–early Miocene other subduction-related structures provide constraints for the age,
due to a similar process. Currently, opening is continuing within the orientation and tectonics associated with the oceanic basins that
Cayman Trough accommodated by the motion along the bounding formed adjacent to the margin. The Laurentian peri-continental mar-
transforms. Active subduction of Atlantic oceanic lithosphere is oc- gin was a passive Atlantic-style margin until the early Mesozoic
curring along the Lesser Antilles Trench, which connects up to the (Nokleberg et al., 2001). Many accretion events have been recorded
Mid-Atlantic Ridge along the Researcher Ridge and Royal Trough along this margin but we simplify them into three main sectors: the
(Müller et al., 1999). Yukon-Tanana/Quesnellia/Stikina terrane, the East Klamath terrane
and the Wrangellia superterrane separated by major fault systems.
3.4.2. Mongol–Okhotsk Basin There are many alternative interpretations for the source of the ter-
The Mongol–Okhotsk Basin is a Mesozoic ocean basin that existed ranes, their age of formation, timing and location of accretion and
between the Siberian craton to the north and the Amuria/Mongolia their field relationships. Our model relies heavily on the reconstruc-
block to the south. The Mongol–Okhotsk suture zone defines basin tions represented in Nokleberg et al. (2001) and Colpron et al.
closure (Apel et al., 2006; Golonka et al., 2006; Cocks and Torsvik, (2007) but note that other alternative scenarios exist.
2007). Evidence for the existence of the Mongol–Okhotsk Basin is Arc magmatism occurred along the western Laurentian margin
found in a series of remnant island arc volcanics and ophiolites ~390–380 Ma forming many of the rocks of the Yukon-Tanana Terrane
adjacent to the suture zone as well as a large area of seismically fast (YTT) and western Kootenay terranes (Nokleberg et al., 2001) currently
material in the lower mantle underlying Siberia imaged in seismic located in Yukon and southern Alaska (Fig. 7). The base of the YTT has
tomography (Van der Voo et al., 1999a). isotopic, geochemical characteristics indicating a Laurentian source for
The opening of the Mongol–Okhotsk Basin is not well constrained, the terrane (Nokleberg et al., 2001). Following a period of arc magma-
ages range from 610 to 570 Ma (Sengör et al., 1993), Ordovician tism was a period with coeval rift-related magmatism leading to the
(Cocks and Torsvik, 2007), Cambrian (Harland et al., 1990) and rifting of the YTT from the Laurentian margin around 360–320 Ma
Permian (Zorin, 1999; Kravchinsky et al., 2002). The large age range (Mortensen, 1992; Nokleberg et al., 2001; Colpron et al., 2002; Nelson
stems from the associations made between geological units in the et al., 2006). The separation of the YTT was driven by N–NE dipping
Siberia, Mongolia and North China realm and the definition of the subduction and led to the opening of the Slide Mountain Ocean. The
ocean basins that existed between these geological units. A zircon Slide Mountain ophiolite, which is currently emplaced onto the YTT
age of 325 Ma from a leucogabbro pegmatite has been associated and Cassier Terranes (Nokleberg et al., 2001) preserves evidence of
with oceanic crust from the Mongol–Okhotsk Ocean (Tomurtogoo et this paleo-ocean basin. The Slide Mountain Ocean is less commonly re-
al., 2005) indicating that seafloor spreading was active from at least ferred to as the Anvil Ocean (Hansen, 1990). Some of the rocks related
the late Carboniferous. In addition, paleomagnetic data suggests that to arc magmatism were left on the margin (in the parautochthonous
Siberia and Mongolia were separated by 10–15° (Zorin, 1999) by rocks of east–central Alaska and the Kootenay terrane) before the open-
the Permian. The presence of continental volcano-sedimentary se- ing of the Slide Mountain Ocean while the majority of the YTT formed
quences and granitoid magmatism proximal to the suture zone indi- the base of the frontal arc (Nokleberg et al., 2001).
cates that the basin was being subducted northward during the The Slide Mountain Ocean opened due to west–southwest slab
Permian (Zorin, 1999), Triassic and Jurassic (Stampfli and Borel, roll-back, reaching a maximum width in the early Permian (Nelson
2002; Golonka et al., 2006). It is difficult to ascertain when seafloor et al., 2006) of around 1300 km (Nokleberg et al., 2001). Spreading
spreading ceased in the Mongol–Okhotsk Basin. Triassic MORB basalts in the back-arc basin ceased at around 280–260 Ma coincident with
in the eastern part of the Mongol–Okhotsk belt (Golonka et al., 2006) a subduction polarity reversal (Mortensen, 1992; Nokleberg et al.,
provide a minimum age for seafloor spreading. Continued subduction 2001) recorded in west-facing coveal calc-alkalic and alkalic plutons
along the Siberian margin led to initial closure of the Mongol– (Nokleberg et al., 2001). The subduction polarity reversal led to the
Okhotsk Ocean sometime in the Jurassic (Van der Voo et al., 1999a; formation of two adjacent arcs, the Stikinia and Quesnellia Arcs, over-
Zorin, 1999; Kravchinsky et al., 2002; Stampfli and Borel, 2002; lying the YTT via a southwest-dipping subduction zone along the
248 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
eastern side of the YTT. This subduction led to the closure of the Slide spreading in the Cache Creek Ocean simultaneous with the accretion
Mountain Ocean and the accretion of the YTT/Quesnellia Arc to the of the YTT along the Laurentian margin at 230 Ma. This is followed by
Laurentian margin by the middle Triassic (240–230 Ma) (Hansen, the subduction of the Cache Creek Ocean behind a rapidly retreating
1990; Nokleberg et al., 2001; Nelson et al., 2006). The Stikinia Arc west-dipping subduction zone along the eastern side of the Stikinia
was still intraoceanic when the YTT/Quesnellia Arc accreted to the Arc and East Klamath (Fig. 17). The Stikinia Arc and East Klamath ac-
margin as it trends outboard of the Cache Creek Terrane (Fig. 7). crete to the North American margin at 172 Ma (Fig. 17), resulting in
The Cache Creek Terrane is a mid-Paleozoic to mid Jurassic oceanic the emplacement of the Cache Creek ophiolite between the Stikinia
terrane with exotic Permian Tethyan faunas in limestone blocks and Arc and the Quesnellia Arc. We accrete the Wrangellia superterrane
long-lived island edifices (Nelson and Mihalynuk, 1993; Mihalynuk to the margin at 140 Ma following the northern accretion model.
et al., 1994). The Cache Creek Terrane, which is very distinct from The accretion of the Wrangellia Terrane marks the true establishment
the Slide Mountain Terrane implies that another ocean basin, the of the boundary between North America and the Pacific.
Cache Creek Ocean, formed in between the Stikinia Arc to the west
and the rapidly retreating YTT/Quesnellia Arc to the east. Based on 3.4.4. Proto-South China Sea
trend-surface analysis of the distribution of Permian coral genera, A Mesozoic–Cenozoic back-arc basin situated adjacent to the
taxonomic diversity and paleomagnetic data, Belasky and Runnegar Eurasian passive margin, named the proto-South China Sea, is incor-
(1994) predict that the Stikinia Arc was located up to 6700 km from porated into many regional models of SE Asia (Hamilton, 1979;
the Laurentian margin in the early Permian and that the Eastern Holloway, 1982; Williams et al., 1988; Hutchison, 1989; Lee and
Klamath terrane was located proximal to the Stikinia Arc. Lawver, 1994; Hall, 2002). Rifting is believed to have initiated along
To address the field relationships of the YTT, Quesnellia Arc, Cache the South China margin in the late Cretaceous (Ru and Pigott, 1986;
Creek Terrane and Stikinia Arc, Colpron et al. (2007) invoke an “oroclinal” Lee and Lawver, 1994) although a rift-related unconformity is dated
model whereby the Stikinia Arc segment rotated counterclockwise con- to the early Cretaceous (Lee and Lawver, 1994). This rift event led
suming the Cache Creek Ocean along a west–southwest-dipping sub- to the separation of northern Borneo from the South China margin
duction zone. The rotation of the Stikinia Arc may have initiated as resulting in the formation of NE–SW trending structures and sedi-
early as ~230 Ma. The timing of accretion of the Stikinia Arc to the mentary basins (Lee and Lawver, 1994). The provenance of ophiolitic
North American margin and therefore the closure of the Cache Creek igneous rocks in northwest Borneo from late Jurassic–late Cretaceous
Ocean is tightly constrained to around 172–174 Ma (Colpron et al., (based on the dating of sediments overlying pillow basalts) is tied to
2007 and references therein). However, collision may have started in the proto-South China Sea (Hutchison, 2005), further constraining
the early Jurassic coincident with a phase of cooling (Nokleberg et al., the timing of formation of the basin.
2001). The cessation of spreading in the proto-South China Sea and its
The next major event to affect the margin was the accretion of the lateral extent is unknown. Most models invoke the initiation of clo-
exotic Wrangellia superterrane. The basement of the Wrangellia super- sure in the early Cenozoic/early Neogene beneath Kalimantan/north-
terrane consists of Triassic flood basalts (285–297 Ma) that formed at ern Borneo and Palawan (Ludwig et al., 1979; Williams et al., 1988;
equatorial latitudes and overlain by a carbonate platform (Richards Lee and Lawver, 1994; Hall, 2002). The closure is believed to have
et al., 1991; Greene et al., 2008). Although recent data suggests initial been triggered either by the counterclockwise rotation of Borneo
collision with the North American margin at about 175 Ma (Gehrels, (Hall, 2002) or by the southeast extrusion of Indochina (Lee and
2001, 2002; Colpron et al., 2007), the main accretion event occurred Lawver, 1994).
at 145–130 Ma (Nokleberg et al., 2001; Trop et al., 2002). There is We model the opening of the proto-South China Sea during rifting
controversy over whether the allochthonous terranes (including between the stable Eurasian margin and northern Borneo during the
Wrangellia) of southern Alaska and western Canada were originally ac- late Cretaceous (~90 Ma) with spreading orthogonal to the Eurasian
creted (1) ≤1000 km of their existing location, offshore present day margin. The cessation of spreading occurred at 50 Ma coincident with
British Columbia, Oregon, and Washington, during the late Mesozoic the clockwise rotation of the neighboring Philippine Sea plate. The dra-
and early Cenozoic or (2) were located 1000–5000 km along the west- matic change in motion of the Philippine Sea plate reorganized the plate
ern coast of the North American Craton and subsequently transported boundaries in the area leading to the establishment of a subduction
northwards during the Late Cretaceous and Cenozoic (Keppie and zone between Palawan and the proto-South China Sea, which began ac-
Dostal, 2001; Stamatakos et al., 2001). After collision, the Wrangellia tively consuming the proto-South China Sea since 50 Ma with an in-
terrane underwent margin-parallel dextral motion but the amount of crease in convergence rate from 25 Ma. We model complete closure of
dextral motion is a matter of debate. the proto-South China Sea at around 10 Ma behind a subduction zone
We model the evolution of the marginal and back-arc basins that located along Palawan and the north Borneo/Kalimantan margin.
formed along the western North American margin as described
above. We create a set of synthetic seafloor spreading isochrons to de- 3.4.5. Western Pacific and SE Asian back-arc basins
pict the opening of the Slide Mountain Ocean starting at 340 Ma The continental blocks and basins in SE Asia comprise one of the
based on a margin parallel opening and a maximum opening width most complex regions in the world. Most models focus on the Cenozoic
of 1300 km, suggested by Nokleberg et al. (2001). Break-up may interpretation of onshore geology, including: Rangin et al. (1990), Lee
have been at least partially driven by a mantle plume as our recon- and Lawver (1995), and Hall (2002). Other models couple the seafloor
structions show that the plume associated with the present day spreading history in the back-arc basins of both SE Asia and the Western
Azores hotspot closely corresponds to the break-up location. Osmium Pacific for a continent and ocean basin evolution (Gaina and Müller,
isotopes suggest that Azores has a deep origin (Schaefer et al., 2002) 2007). The model we use in our reconstructions is based on Gaina and
suggesting that this plume may have been long-lived but whether Müller (2007) and additionally incorporate the rotation of the
hotspots are active and can be traced as far back as 340 Ma remains Philippine Sea plate based on Hall et al. (1995) and the seafloor spread-
open to debate. We terminate spreading in the Slide Mountain ing model of Sdrolias et al. (2003b) for spreading in the Parece Vela and
Ocean at 280 Ma followed by a subduction polarity flip along the Shikoku Basins. For further details of the model, we refer to Sdrolias et
YTT and the establishment of an eastward retreating subduction al. (2003b) and Gaina and Müller (2007).
zone. Subduction led to the consumption of the Slide Mountain
Ocean along this southwest–west dipping subduction zone. 3.4.6. SW Pacific Back-arc basins and marginal seas
We form the Cache Creek Ocean in between the retreating YTT The SW Pacific is characterized by a series of marginal basins
and the Stikinia Arc and East Klamath at 280 Ma with a cessation of (Tasman and Coral Seas), submerged continental slivers (Lord Howe
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 249
60°
200 Ma 200 Ma
40°
20°
0°
240° 260° 280° 300° 320° 340° 240° 260° 280° 300° 320° 340°
60°
180 Ma 180 Ma
40°
20°
0°
240° 260° 280° 300° 320° 340° 240° 260° 280° 300° 320° 340°
60°
170 Ma 170 Ma
40°
20°
0°
240° 260° 280° 300° 320° 340° 240° 260° 280° 300° 320° 340°
60°
150 Ma 150 Ma
40°
20°
0°
240° 260° 280° 300° 320° 340° 240° 260° 280° 300° 320° 340°
60°
140 Ma 140 Ma
40°
20°
0°
Fig. 17. (left) Agegrid reconstructions of Mesozoic North America at 200, 180, 170, 150, 140 Ma highlighting the age–area distribution of oceanic lithosphere at the time of forma-
tion and the extent of continental crust (gray polygons). Plate boundaries from our continuously closing plate polygon dataset are denoted as thick white lines, hotspot locations as
yellow stars, large igneous provinces and flood basalts as brown polygons and coastlines as thin black lines. (right) Reconstructions showing the outlines of the plates around Me-
sozoic North America for each reconstruction time listed above. Feature descriptions as in panel (left). Abbreviations are: CAR = Caribbean plate, EUR = Eurasian plate, FAR =
Farallon plate, IZA = Izanagi plate, NAM = North American plate, PAC = Pacific plate, SAM = South American plate. (For interpretation of the references to color in this figure
legend, the reader is referred to the web version of this article.)
250 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
Rise, Mellish Rise, Louisiade, Papuan, Kenn, Dampier and Chesterfield Pacific, Phoenix–Farallon, Phoenix–Pacific, Farallon–Pacific). Initially
Plateaus), island arcs (Norfolk, Three-Kings, Loyalty, New Hebrides, spreading along the Pacific ridges was slow/moderate (70–80 mm/yr)
Vitiaz and Lau-Colville Ridges), back-arc basins (South Loyalty, with a progressive increase in spreading rates to a peak in the mid
North Loyalty, Norfolk, South Fiji, North Fiji and Lau Basins and Cretaceous. In northeast Panthalassa, closure of the Cache Creek Ocean
Havre Trough) as well as numerous features with an uncertain origin (back-arc basin which formed between the Yukun-Tanana Terrane
(e.g. D'Entrecasteaux Zone and Basin and Rennell Trough and Basin) and the Stikinia Arc) was occurring along a southwest dipping subduc-
(Fig. 11). In a broad sense, these features developed behind the east- tion zone on the eastern side of the Stikinia Arc. In northwestern
ward migrating Australia–Pacific plate boundary from the late Meso- Panthalassa, the Mongol–Okhotsk Ocean (an ancient ocean basin
zoic to the present day (Karig, 1971; Symonds et al., 1996; Müller et which formed between Amuria and Siberia) continued its closure via
al., 2000; Crawford et al., 2002; Sdrolias et al., 2003a). Our plate mo- northeast directed subduction along the southern Siberia margin. This
tion model incorporates the opening model for the Tasman and Coral Mongol–Okhotsk subduction zone connected with the landward-
Seas based on Gaina et al. (1998) and Gaina et al. (1999). We incorpo- facing northern Panthalassic subduction zone to its northeast and the
rate the model of Sdrolias et al. (2003a) and Sdrolias et al. (2004) for Tethyan subduction zone to its southwest.
the formation of the back-arc basin and island arc systems seaward of In the Tethys Ocean, the remnant paleo-Tethys was separated
the Lord Howe Rise. For further details, we refer to the abovemen- from the actively spreading meso-Tethys ocean by the continental
tioned publications. blocks of the Cimmerian terrane (e.g. Iran, Afghanistan, Pakistan,
South Tibet, Sibumasu). The Tethyan subduction zone located along
4. Global plate reconstructions the southern Laurasian margin was driving the opening of the
Meso-Tethys and consumption of the paleo-Tethys ocean. Active rift-
Our regional kinematic models fit within a hierarchical global plate ing was occurring along the Argo Abyssal Plain (NW Australia) that
circuit tied to a hybrid moving hotspot/true polar wander corrected ab- we suggest extended to the north of Greater India and westward to
solute reference frame through Africa. We create a set of dynamic plate the East Africa/Karoo Rift, marking the break-up of Gondwanaland
polygons since the time of Pangea break-up with the assumption that into West Gondwana (including South America, most of Africa and
the plates themselves are rigid. The birth of a plate (the establishment Arabia) and East Gondwana (including Antarctica, Australia, India,
of relative motion after a break in the lithosphere), can be defined in eastern Africa, Madagascar). We continue the Karoo Rift southward to
two ways: either the initiation of rifting due to weakening of the litho- connect with extension along the Agulhas-Falkland transform. This
sphere by basal heating forming a series of faults and rift-related struc- plate boundary between West Gondwana and Patagonia connected
tures (sometimes called incipient spreading), or the initiation of with east-dipping subduction along the South American/Panthalassa
seafloor spreading, when there is a complete break of the lithosphere margin.
and extrusion of the mantle. Our plate boundary set distinguishes be- An extensive seaway between the Tethys Ocean and Panthalassa
tween the two modes via a continental/oceanic rift or mid-ocean existed in the mid-Mesozoic. We envisage that the confluence of
ridge coding of the plate boundaries, which allows for the construction these two oceanic domains occurred north of Australia at the so-
of a plate polygon dataset using either mode. The plate polygons pre- called Junction region/plate (Seton and Müller, 2008). The differential
sented in this study follow the former definition but an ancillary set can motion between the meso-Tethys and Izanagi plates results in con-
be produced to follow the later definition. Below we describe tectonic vergence and we model the subduction of Izanagi lithosphere be-
events every 20 million years with accompanying maps (Figs. 18–28) neath a westward verging subduction zone.
and also provide the plate polygon and plate boundary files. These files
can be directly loaded into GPlates software for reconstructions in one 4.2. 180–160 Ma (Figs. 19 and 20)
million year time intervals.
At 180 Ma, early opening by ultra-slow seafloor spreading contin-
4.1. 200–180 Ma (Figs. 18 and 19) ued in the Central Atlantic with ongoing rifting in the northern
Atlantic and Caribbean. A readjustment of the plate-mantle system
Prior to the Mesozoic, the continents were amalgamated into one occurred at 170 Ma, coincident with a doubling of seafloor spreading
big supercontinent, Pangea, surrounded by two ancient oceans, rates in the Central Atlantic (Labails et al., 2010) and the establish-
Panthalassa and the smaller Tethys Ocean. By the early–mid Mesozo- ment of seafloor spreading in the Gulf of Mexico. Evidence for
ic, Pangea was undergoing slow continental break-up centered along changes in plate motion and accretion events in the Tethys Ocean
a rift zone extending from the Arctic, North Atlantic (adjacent to the and Panthalassa at 170 Ma (see below) may indicate a global plate
Norwegian shelf and Iberia–Newfoundland margins), Central Atlantic reorganization event at this time.
and along the Jacksonville Fracture Zone through Florida and the Gulf This time period saw the accelerated growth of the Pacific plate at
of Mexico in the Caribbean region. The Caribbean rift zone, defined the expense of the Izanagi, Farallon and Phoenix plates. In northeast
by a series of Mesozoic rift basins, connected with east-dipping Panthalassa, closure of the Cache Creek Ocean, obduction of the
trans-America subduction, which was consuming oceanic lithosphere Cache Creek Terrane and accretion of the Stikinia Arc occurred
from Panthalassa. At 190 Ma, there was a change from rift to drift along the Laurentian margin between 175 and 172 Ma. The accretion
along the early Atlantic rift, restricted to the Central Atlantic. Contem- of the Stikinia Arc forced a jump in the locus of subduction and rever-
poraneously, dextral motion was occurring along the early Atlas Rift, sal of subduction polarity from southwest to northeast along the new
isolating Morocco. Laurentian margin, establishing the Farallon subduction zone. The
The Panthalassic Ocean was entirely surrounded by subduction northwest Panthalassa margin interacted with the Mongol–Okhotsk
during the mid–early Mesozoic. We model seafloor spreading as a Ocean, which continued its closure along the southern Siberia sub-
simple three-plate system between the Izanagi, Farallon and Phoenix duction zone.
plates. The three arms of the triple junction extended outward inter- Rifting continued along the southern Tethyan margin, adjacent to
secting with the circum-Panthalassic margins with minor margin mi- Argoland/West Burma and northern Greater India to the east African
gration: east of Australia (Izanagi–Phoenix ridge), along the Amurian rifts. In the western Tethys, volcanism ceased along the Karoo Rift at
margin (Izanagi–Farallon ridge) and southern North America (Farallon– 180 Ma leading to a jump in the locus of rifting from the Karoo Rift
Phoenix ridge). At 190 Ma, the birth of the Pacific plate established a to the area between Africa and Madagascar/Antarctica, later forming
more complex spreading ridge system involving three triple junctions the Weddell and Riiser-Larson Sea and Mozambique and West Somali
and six spreading centers (Izanagi–Farallon, Izanagi–Phoenix, Izanagi– Basins. Incipient spreading in the Mozambique and West Somali
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 251
a
60˚N Laurentia Amuria
Laurasia
IZA
30˚N
FAR IZA
NMT
0˚
SMT
30˚S Gondwanaland
PHX PHX
60˚S
200 Ma
0 20 40 60 80 100 120 140 160 180 200 220 240 260 280
Age of Oceanic Lithosphere [m.y.]
Amuria SMT
FAR NMT
Laurasia PHX
Laurentia
Gondwanaland IZA
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170 180 190 200 210 220 230 240 280
Age of Oceanic Lithosphere [m.y.]
Fig. 18. Global plate reconstructions from 200 Ma to the present day in 20 million year time intervals. Basemap shows the age–area distribution of oceanic lithosphere at the time of
formation. Red lines denote subduction zones, black lines denote mid-ocean ridges and transform faults. Brown polygons indicate products of plume-related excessive volcanism.
Yellow stars are present day hotspot locations. Absolute plate velocity vectors are denoted as black arrows. Abbreviations for the plates are the same as in previous figures. Addi-
tional abbreviations include: ALA = Alaska, CA = Central Atlantic, CAP = Capricorn, CAR = Caribbean, CAT = Catequil, CCO = Cache Creek Ocean, COL = Colorado, CS = Caroline
Sea, JUN = Junction, MOO = Mongol–Okhotsk Ocean, NL = North Loyalty Basin, NMT = North Meso-Tethys, NNT = North Neo-Tethys, PAR = Parana, PAT = Patagonia, PS =
Philippine Sea, PSC = Proto-South China Sea, SCO = Scotia Sea, SLB = South Loyalty Basin, SMT = South Meso-Tethys, TS = Tasman Sea.
Basins connected with both the Weddell Sea rift and the Agulhas- 4.3. 160–140 Ma (Figs. 20 and 21)
Falkland transform in the south. In the northern Tethys, closure of
the paleo-Tethys and accretion of the Cimmerian terrane occurred The Central Atlantic continued spreading between 160 and
along the southern Laurasian margin at 170 Ma. Spreading in the 140 Ma, connecting with the Gulf of Mexico ridge system to the
meso-Tethys continued with an acceleration in spreading rate after south. After the cessation of spreading in the Gulf of Mexico, the
the complete accretion of the Cimmerian terrane at 170 Ma. At mid-ocean ridge jumped southward initiating the opening of the
165 Ma, rifting extended southward from Argoland to the area be- proto-Caribbean Basin through the accommodation space created
tween Australia and India (adjacent to the Gascoyne, Cuvier and due to the relative motion between the North and South American
Perth Abyssal Plains) thereby initiating a plate boundary between plates. Spreading was NW–SE directed and initiated around 145 Ma
India and Australia. This connected with the newly established rift forming a triple junction to the east between the mid ocean ridge of
margin between Australia and Antarctica at 165 Ma and extended the Central Atlantic and rift axis of the Equatorial/South Atlantic. To
into the Enderby Basin from 165 Ma to the west connected with the the west, the spreading ridge of the proto-Caribbean Basin formed a
Western Panthalassic subduction zone along eastern Australia to the ridge–ridge–transform triple junction with the spreading ridge of
east. the Andean back-arc basin and the trans-American subduction zone.
252 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
a
60˚N Amuria
Laurentia
IZA Laurasia
30˚N
FAR IZA
NMT
0˚
West Gondwana
PAC
SMT JUN
30˚S
PHX PHX
FLK
East Gondwana
60˚S
180 Ma
0 20 40 60 80 100 120 140 160 180 200 220 240 260 280
Age of Oceanic Lithosphere [m.y.]
180 Ma
IZA West Gondwana
FAR
FLK
East Gondwana
Amuria SMT
FAR NMT
Laurasia PHX
PAC
JUN
Laurentia
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170 180 190 200 210 220 230 240 280
Age of Oceanic Lithosphere [m.y.]
In the South Atlantic, extension began within continental South formation of the neo-Tethys ocean. Spreading in the meso-Tethys con-
America at 150 Ma, partitioning the southern part of the continent tinued the meso-Tethys ridge intersected the Tethyan subduction
into the Parana and Colorado subplates and inducing a rift zone be- zone around 140–145 Ma resulting in a southern ridge jump and con-
tween South America and Africa, which connected to the Agulhas- tinuation of seafloor spreading in the meso-Tethys.
Falkland transform to the south. Spreading and growth of the Pacific plate continued in Panthalassa,
The Agulhas-Falkland transform extended eastward connecting to with a gradual increase in spreading rate. The eruption of the Shatsky
the mid-ocean ridge in the Weddell Sea, which was established at Rise at the Pacific–Izanagi–Farallon triple junction led to a major read-
160 Ma. The Weddell Sea ridge joined with mid-ocean ridges along justment of the triple junction center and was coincident with a major
East Africa, including between Africa and Antarctica in the Mozambique clockwise change in spreading direction, by 24°, between the Pacific
Basin/Riiser-Larson Sea and Africa and Madagascar in the West Somali and Izanagi plates at M21 (~147 Ma). This resulted in an increased clock-
Basin. This newly established ridge system led to an acceleration of wise rotation and a change in configuration of the Pacific–Izanagi, Iza-
break-up between East and West Gondwana. From 144 Ma onwards, nagi–Phoenix and Izanagi–Farallon ridges. The Mongol–Okhotsk Ocean
Madagascar operated as an independent plate. In the eastern Tethys, closed at 150 Ma forming the Mongol–Okhotsk Suture.
rifting extended along the Argo Gascoyne, Cuvier and Perth Abyssal In the Arctic Ocean, the Canada Basin initiated opening at 145 Ma via
Plains forming a triple junction between the Australia/Antarctic rift counterclockwise rotation of North Slope of Alaska with seafloor
margin and the Enderby rift. By 156 Ma, NW–SE oriented seafloor spreading starting at 142 Ma. The Canada Basin spreading ridge con-
spreading begun in the Argo Abyssal Plain, rifting West Burma/Argoland nected with the North Atlantic rift zone, which extended as far south
and establishing the mid-ocean ridge system that resulted in the as the Kings Trough adjacent to the Newfoundland/Iberia margin. The
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 253
a
60˚N
Laurasia
Laurentia Amuria
30˚N IZA IZA
FAR
CA NMT
0˚
JUN
PAC West
SMT
Gondwana
30˚S
60˚S FLK
Antarctica
160 Ma
0 20 40 60 80 100 120 140 160 180 200 220 240 260 280
Age of Oceanic Lithosphere [m.y.]
160 Ma
IZA West Gondwana
FAR FLK
Antarctica
Amuria SMT
CA
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170 180 190 200 210 220 230 240 280
Age of Oceanic Lithosphere [m.y.]
plate boundary follows the Kings Trough through the Pyrenees connect- break the African continent into three discrete plates: South, NW and
ing with the northern Tethyan subduction zone and to the south con- NE Africa. Seafloor spreading between Madagascar and the East African
nects with the Central Atlantic mid-ocean ridge. margin ceased around 120 Ma. In the South Atlantic, seafloor spreading
propagated northward to the central segment of this ocean by 125 Ma.
4.4. 140–120 Ma (Figs. 21 and 22) The early–mid Cretaceous marks a significant increase in seafloor
spreading rates in Panthalassa corresponding to the mid-Cretaceous
The Central Atlantic and Iberia–Newfoundland spreading ridge con- seafloor spreading pulse. Spreading was occurring between the Pacific,
tinued and connected via a series of rift zones to the Canada Basin in Farallon, Izanagi and Phoenix plates. In northern Panthalassa, North
the Arctic and to the South Atlantic spreading center to the south. In ad- Slope of Alaska was continuing its counterclockwise rotation and open-
dition, rifting between North America and Greenland initiated around ing of the Canada Basin.
135 Ma, establishing Greenland as an independent plate and marking The southwest Panthalassic margin, along eastern Australia in-
the end of the Laurentian continental landmass. The proto-Caribbean volved the opening of the South Loyalty Basin, due to roll-back of
Sea continued its growth via differential motion between South and the southwest Panthalassic subduction zone from 140 Ma. The
North America. Seafloor spreading initiated in the southern South Atlantic South Loyalty Basin was actively opening until 120 Ma until a major
by 132 Ma coinciding with a peak in magmatism (Parana–Etendeka Large change in the plate configurations in the SW Panthalassic Ocean.
Igneous Province) and the initiation of rifting in the African continental Seafloor spreading in the meso-Tethys continued after its southern
interior via the West and Central African rift zones. At this time, we ridge jump at 140 Ma. Coincidently, spreading along the neo-Tethys
254 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
a ALA
60˚N
Laurentia Eurasia
IZA
IZA
30˚N FAR
PC
0˚
PAC NNT
West JUN
30˚S Gondwana
PHX PHX
COL East Gondwana
60˚S PAT
Antarctica
140 Ma
0 20 40 60 80 100 120 140 160 180 200 220 240 260 280
COL
FAR PAT
Antarctica
NNT NNT
FAR East
Eurasia
PHX Gondwana
PAC
Laurentia
JUN
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170 180 190 200 210 220 230 240 280
Age of Oceanic Lithosphere [m.y.]
ridge extending from the Argo Abyssal Plain to north of Greater India. along the South Atlantic ridge with a northward propagation leading
After a landward ridge jump of the neo-Tethys ridge at 135 Ma, the to seafloor spreading in the “Central” segment by 120 Ma and in the
mid-ocean ridge propagated southward to open the Gascoyne, Cuvier “Equatorial” segment by 110 Ma. Extension along the West and
and Perth Abyssal Plains between India and Australia. The West Central African rifts, including the Benue Trough continued during this
Australian spreading ridge system joined with the Enderby Basin spread- time period. Further north, spreading between Iberia and Newfoundland
ing ridge, separating Antarctica from the Elan Bank/India, to the west and connected to a rift zone adjacent to the Rockall and Porcupine Plateaus
to the rift between Australia and Antarctica to the east. The initiation and continued to the Labrador Sea/Baffin Bay (between Greenland and
of seafloor spreading in the Enderby Basin accommodated strike-slip North America) and between Greenland and Eurasia. Break-up between
motion between India and Madagascar of over 1000 km and connected Porcupine and North America occurred from 110 Ma. These North
to the West Somali Basin spreading ridge. The East African and Weddell Atlantic rift zones connected with the Canada Basin spreading center
Sea spreading ridges were active during this time period and connected until about 118 Ma when spreading ceased in the Canada Basin. Spread-
to the South Atlantic via the Agulhas-Falkland transform. ing terminated when the rotation of North Slope Alaska ceased, coinci-
dent with a change in the southern North Slope margin from largely
4.5. 120–100 Ma (Figs. 22 and 23) strike-slip to convergence due to a change in spreading direction in
Panthalassa.
Spreading along the Central Atlantic ridge continued into the Ultra fast seafloor spreading rates were occurring in Panthalassa to-
proto-Caribbean Sea until 100 Ma. Spreading extended southward gether with the eruption of a suite of Large Igneous Provinces, most
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 255
a ALA
60˚N
GRN
Eurasia
North
IZA America
30˚N IZA
FAR
NWA
NEA
0˚
0 20 40 60 80 100 120 140 160 180 200 220 240 260 280
Age of Oceanic Lithosphere [m.y.]
b North Pole South Pole
NWA
120 Ma
South NEA
America
PAC IZA
SAF
FAR
CHA Antarctica
CAT India
ALA India
FAR
Eurasia MAN
HIK Australia
GRN
North PAC
America
JUN
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170 180 190 200 210 220 230 240 280
Age of Oceanic Lithosphere [m.y.]
notably the eruption of the Ontong-Java, Manihiki and Hikurangi Pla- In the Tethys Ocean, spreading was continuing along the western
teaus at 120 Ma. The eruption of this mega-LIP led directly to the Australian margin, connecting to spreading in the Enderby Basin and
break-up of the Phoenix plate into four plates: the Hikurangi, Manihiki, rifting between Australia and Antarctica. A ridge jump at 120 Ma
Chasca and Catequil plates. The separation occurred at 120 Ma in an E– isolated the Elan Bank microcontinent, roughly coincident with the
W direction in the Ellice Basin between the Ontong Java and Manihiki eruption of the Kerguelen Plateau. A strike-slip margin between
Plateaus with simultaneous rifting of the Manihiki and Hikurangi pla- India and Madagascar joined to a transform in the Tethys Ocean and
teaus from a N–S directed spreading system along the Osbourn Trough. not to the West Somali Basin spreading ridge which had become ex-
An additional two triple junctions were active in the region leading to tinct at 120 Ma. Spreading continued in the Mozambique Basin/Riiser
the break-up of the Eastern Manihiki Plateau and the development of Larson Sea and continued to the Weddell Sea and north to the South
the Tongareva triple junction. The eastern triple junction represented Atlantic spreading ridge.
spreading between the Manihiki, Phoenix and Chasca plate and the
southern triple junction represented spreading between the Hikurangi, 4.6. 100–80 Ma (Figs. 23 and 24)
Catequil and Manihiki plates. The initiation of the Pacific–Manihiki–
Hikurangi triple junction led to change in the tectonic regime along east- The Mid and South Atlantic Ridges were well established from
ern Australia. Prior to 120 Ma, the Phoenix plate was subducting beneath 100 Ma. As spreading occurred, rifting in the interior of Africa ceased
the east Australia margin, which changed to the Hikurangi plate and a at about 85 Ma. The Mid-Atlantic ridge propagated northward to be-
small portion of the Catequil plate but with a decreased rate of conver- tween the Porcupine margin and between North America and the
gence after 120 Ma. Rockall margin at 50 Ma. Rifts were still active surrounding Greenland.
256 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
a
60˚N
GRN
North Eurasia
America
IZA
30˚N IZA
FAR
NWA
NEA
0˚
PAC
South
America India JUN
30˚S SAF
PAC
CHA
MAN
Australia
60˚S HIK
CAT
Antarctica
100 Ma
0 20 40 60 80 100 120 140 160 180 200 220 240 260 280
Age of Oceanic Lithosphere [m.y.]
b North Pole South Pole
NWA
100 Ma
South NEA
America
IZA SAF
PAC
FAR
CHA
Antarctica
CAT India
Eurasia India
FAR
MAN
North Australia
GRN HIK
America
PAC JUN
SAM
NWA NEA
IZA
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170 180 190 200 210 220 230 240 280
Age of Oceanic Lithosphere [m.y.]
The south of the Mid Atlantic Ridge connected to the actively opening with an observed bend in the hotspot trails on the Pacific plate, suggest-
proto-Caribbean Sea along a major transform fault. The western margin ing a plate reorganization at this time. In addition, we model a clockwise
of the Caribbean plate underwent a change in subduction polarity from change in spreading direction in the Osbourn Trough region based on our
east-dipping to west-dipping at 100 Ma. The rollback of this subduction age estimate for a bend in observed fracture zones between the Manihiki
zone along the Caribbean Arc led to the consumption of the actively and Hikurangi plateaus. The change in spreading direction modified the
spreading proto-Caribbean ocean floor and encroachment of the nature of the boundary east of Australia from convergence to dominantly
Farallon plate into the Caribbean domain (Fig. 9). The continued roll- strike-slip. At 86 Ma, we model the docking of the Hikurangi Plateau
back of the Caribbean Arc subduction zone led to the formation of the with the Chatham Rise triggering a cessation in spreading associated
Yucatan Basin as a back-arc in the late Cretaceous. The eruption of the the Ontong-Java, Manihiki and Hikurangi plateaus. After the cessation
Caribbean flood basalt province occurred around 90 Ma overlying oce- of spreading along these ridges axes, the locus of extension jumped
anic lithosphere that formed on the Farallon plate and later migrated southward between Antarctica and the Chatham Rise, establishing the
to the Caribbean region. Pacific–Antarctic spreading ridge. To the east, the Pacific–Farallon Ridge
In Panthalassa, spreading was occurring along the Pacific–Izanagi, extended to the south connecting with the Pacific–Antarctic Ridge at
Pacific–Farallon, Farallon–Izanagi and along the ridges associated with the Pacific–Antarctic–Farallon triple junction.
the plateau break-up region. A change in spreading direction is recorded After the cessation of the spreading centers associated with the LIP
in the Mendocino, Molokai and Clarion fracture zones (associated with break-up, the Pacific plate became the dominant plate in Panthalassa
Pacific–Farallon spreading), which we date to 103–100 Ma coincident and it is at this time that we switch to the Pacific Ocean. In the
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 257
a
GRN
60˚N
IZA North
America Eurasia
IBR IZA
30˚N
FAR
0˚ PAC
AFR PAC
South India
America JUN
30˚S
Australia
LHR
60˚S
Antarctica
80 Ma
0 20 40 60 80 100 120 140 160 180 200 220 240 260 280
Age of Oceanic Lithosphere [m.y.]
b
North Pole South Pole
80 Ma
South AFR
America
IZA
PAC
FAR India
FAR Antarctica
Eurasia
North Australia
India
America GRN
LHR
JUN
IBR
SAM PAC IZA
AFR
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170 180 190 200 210 220 230 240 280
Age of Oceanic Lithosphere [m.y.]
western Pacific, the Tasman Sea was opening from 84 Ma leading to Atlantic at 83.5 Ma and the American–Antarctic ridge (established
the establishment of the Lord Howe Rise plate. Further north, the after the cessation of spreading in the Weddell Sea). The West Burma
proto-South China Sea initiated its opening between the South continental sliver reached the Eurasian margin and accreted starting
China margin and Borneo/Kalimantan. at 87 Ma and sutured to Sibumasu at 73 Ma.
In the Tethys/Indian Ocean, spreading was occurring along the West
Australian margins continuing the separation of India and West Burma 4.7. 80–60 Ma (Figs. 24 and 25)
from Australia. A major change direction is recorded in the fracture zone
trends at 99 Ma, led to a change in the motion of the Indian plate. The South and Mid-Atlantic ridges continued spreading. The Mid-
Spreading became dominantly N–S directed establishing spreading in Atlantic Ridge propagated northward into the North Atlantic with the
the Wharton Basin. The West Australian mid ocean ridge system formed initiation of seafloor spreading in the Labrador Sea (between North
a triple junction with the Australian–Antarctic ridge at 99 Ma (initiation America and Greenland) and between Rockall and Greenland at
of ultra-slow seafloor spreading) and spreading between India and 79 Ma. Spreading propagated from the Labrador Sea to Baffin Bay by
Antarctica north of Elan Bank. The Indian–Antarctic ridge (or Southeast 63 Ma across the Davis Straits via left-lateral transform faults and
Indian Ridge) connected with the African–Antarctic ridge (or South- connected to the Arctic via the Nares Strait. In the Caribbean, spreading
west Indian Ridge) from 100 Ma. Rifting between India and Madagascar in the proto-Caribbean Sea ceased at 80 Ma whereas the Caribbean Arc
in the Mascarene Basin initiated at 87 Ma. The Southwest Indian Ridge subduction zone continued its northeastward rollback. The Yucatan
connected with spreading in the Malvinas plate in the southernmost Basin opened as a back-arc in the late Cretaceous with cessation
258 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
a
GRN
60˚N
IZA North Eurasia IZA
America
IBR
30˚N
FAR CAR
0 20 40 60 80 100 120 140 160 180 200 220 240 260 280
Age of Oceanic Lithosphere [m.y.]
b North Pole South Pole
60 Ma
PAC AFR
South
America
IZA
FAR India
FAR Antarctica
Eurasia
North Australia
India
America GRN
CAR LHR
JUN
IBR
SAM PAC
AFR
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170 180 190 200 210 220 230 240 280
Age of Oceanic Lithosphere [m.y.]
occurring at 70 Ma when the Caribbean Arc accreted to the Bahaman ridge, extending eastward to connect with the Pacific–Farallon and
Platform. The accretion led to a jump in the locus of subduction west- Farallon–Antarctic spreading ridges. At 67 Ma, a change in spreading
ward along the newly developed Panama–Costa Rica to accommodate direction is recorded in the fracture zones of the South Pacific.
the continued eastward motion of the Farallon plate, trapping Farallon In the Indian Ocean, spreading was occurring along the Wharton
oceanic lithosphere onto the Caribbean plate. Ridge, Southeast Indian Ridge, Southwest Indian Ridge and in the Mas-
The Pacific was dominated by the break-up of the Farallon plate carene Basin. Spreading in the Mascarene Basin ceased at 64 Ma jump-
into the Kula plate at 79 Ma initiating spreading along the E–W trend- ing northward, isolating the Seychelles microcontinent and initiating
ing Kula–Pacific ridge and the NE–SW trending Kula–Farallon ridge. spreading between India and the Seychelles along the Carlsberg Ridge.
The Kula–Farallon Ridge follows the location of the Yellowstone hot- The Southwest Indian Ridge connected with spreading in the Malvinas
spot and intersects the North American margin in Washington/British plate until 66 Ma. After this, the Southwest Indian Ridge connected di-
Columbia before migrating northward along the margin. The break up rectly with the American–Antarctic and South Atlantic Ridge.
of the Farallon plate into the Kula plate coincides with a major change
in spreading direction observed in all northeast Pacific fracture zones. 4.8. 60–40 Ma (Figs. 25 and 26)
In our model spreading continued along the Pacific–Izanagi ridge
after the establishment of the Kula–Pacific ridge to the east connected Seafloor spreading propagated into the Eurasia–Greenland margin
via a large offset transform fault. The Pacific–Izanagi ridge was rapidly along the Reykjanes Ridge by 58 Ma, forming a triple junction be-
approaching the East Asian margin and was proximal by 60 Ma. In the tween North America, Greenland and Eurasia. The Jan Mayen micro-
southern Pacific, spreading was occurring along the Pacific–Antarctic continent rifted off the margin forming the fan-shaped Norway
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 259
a GRN
60˚N
KUL North Eurasia
America
VAN IBR
30˚N
CAR
FAR India
PS
0˚ PAC AFR PAC
South
America
SS
30˚S Australia
NL
60˚S
East Antarctica
40 Ma
0 20 40 60 80 100 120 140 160 180 200 220 240 260 280
Age of Oceanic Lithosphere [m.y.]
b North Pole South Pole
40 Ma
PAC AFR
South
PS America
KUL India
East
Antarctica
FAR FAR
North Eurasia WANT
America
GRN India Australia
CAR
SAM IBR SS
NL
PAC PS
AFR
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170 180 190 200 210 220 230 240 280
Age of Oceanic Lithosphere [m.y.]
Basin along the Aegir Ridge. The Aegir Ridge connected to the Mohns In the Pacific, the Pacific–Izanagi ridge started to subduct under the
Ridge to the north and Reykjanes Ridge to the south via a series of East Asian margin between 55 and 50 Ma, signaling the death of the Iza-
transform faults. Spreading in the Eurasian Basin to the north initiat- nagi plate coincident with a dramatic change in spreading direction
ed around 55 Ma along the Gakkel/Nansen Ridge. This ridge con- from N–S to NW–SE between Kula–Pacific spreading. The Kula–Pacific
nected to the Baffin Bay ridge axis through the Nares Strait and the Ridge connected with the Pacific–Farallon Ridge and Kula–Farallon
Mohns Ridge to the south via major strike-slip faults with minor com- Ridge from 60 to 55 Ma. After 55 Ma, the eastern Pacific was dominated
pression between Greenland and Svalbard. In our model the Lomono- by the rupture of the Farallon plate close to the Pioneer Fracture Zone,
sov Ridge is coupled to North America. The initiation of spreading in forming the Vancouver plate. The break-up resulted in minor relative
the Eurasian Basin also coincides with the initiation of independent motion along the Pioneer fracture zone. Further south, spreading was
motion of the Porcupine Plate, resulting in a small clockwise rotation continuing along the Pacific–Farallon, Pacific–Antarctic, Farallon–
of Eurasia and counter-clockwise rotation of Iberia relative to the Por- Antarctic and Pacific–Aluk Ridges. The fracture zones associated with
cupine Plate. A change in spreading direction is also observed in the the Pacific–Antarctic Ridge close to the Campbell Plateau record a
Labrador Sea. change in spreading direction at 55 Ma, coincident with other events
The Mid-Atlantic Ridge connects with the west-dipping subduc- that occurred in the Pacific at this time.
tion zone bordering the Caribbean via a transform fault. By the middle In the western Pacific, spreading in the proto-South China Sea ceased
Eocene, relative motion between North America and the Caribbean at 50 Ma coincident with the clockwise rotation of the neighboring
began to form the Cayman Trough along sinistral faults that later Philippine Sea plate. The dramatic change in motion of the Philippine
merge with the Lesser Antilles trench. East-dipping subduction was Sea plate reorganized the plate boundaries in the area leading to the
still occurring along the Middle America margin bordering the Pacific. establishment of a subduction zone between Palawan and the proto-
260 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
a
60˚N North
America Eurasia
J J
30˚N
CAR AFR PS
COC India PAC
0˚
PAC
CS
South CAP SS
SOM
America NB
NAZ
30˚S
Australia
60˚S SCO
Antarctica
20 Ma
0 20 40 60 80 100 120 140 160 180 200 220 240 260 280
Age of Oceanic Lithosphere [m.y.]
20 Ma
AFR
PAC South
PS
America SOM
J SCO
CAP
COC NAZ Antarctica
North Eurasia
America
India
CAR Australia
ARA
SAM
PAC NB
SS PS
AFR
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170 180 190 200 210 220 230 240 280
Age of Oceanic Lithosphere [m.y.]
South China Sea, which led to the subduction of the proto-South China triple junction to the Eurasian Basin in the north. The cessation of inde-
Sea after 50 Ma. Spreading was occurring in the West Philippine Basin pendent Porcupine motion occurred at 33 Ma coinciding with the
and Celebes Sea. Further south, spreading initiated in the North Loyalty cessation of seafloor spreading in the neighboring Labrador Sea and
Basin behind the proto-Tonga–Kermadec Trench. Baffin Bay and the establishment of a simple two-plate system to de-
The Indian Ocean was dominated by a series of mid ocean ridges scribe the plate motions in the North Atlantic. From 33 Ma onwards,
such as the Wharton Ridge, Southeast Indian Ridge, Southwest Indian Greenland and North America have been fused into one plate. At
Ridge and Carlsberg Ridge. Prior to 55 Ma, subduction was occurring about 30 Ma, spreading jumped from the Aegir Ridge in the Norway
along the Tethyan subduction zone, consuming crust that formed Basin to the Kolbeinsey Ridge connecting up with the Mohns Ridge
during meso and neo Tethys spreading. At 55 Ma, the northern tip via a series of transform faults. Further south, adjacent to the Iberian
of Greater India marks the start of collision between India and Eurasia margin, a southern jump of the plate boundary at 28 Ma from the
and the uplift of the Himalayas. Closure of the Tethys Ocean in this Kings Tough to the Azores transform fault and along the Straits of
area occurred by about 43 Ma. Full closure of the neo-Tethys between Gibraltar led to the capture of Iberia by the Eurasian plate.
India and Eurasia also corresponds to the cessation of spreading in the In the Pacific, spreading between the Kula–Pacific and Kula–Farallon
Wharton Basin, which describes Australia–India motion. ceased at 40 Ma, leading to the Pacific plate consisting of the Pacific,
Vancouver, Farallon, Aluk and Antarctic plates. The intersection of the
4.9. 40–20 Ma (Figs. 26 and 27) Murray transform fault with the North American subduction zone
around 30 Ma led to the establishment of the San Andreas Fault and
At 40 Ma, the Atlantic Ocean consisted of a continuous mid-ocean corresponds to the establishment of the Juan De Fuca plate at the ex-
ridge system that extended from the South America–Antarctica–Africa pense of the Vancouver plate. A further rupture of the Farallon plate
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 261
a
60˚N
North Eurasia
J America
J
30˚N
ARA
PS
R CAR India
COC AFR
PAC PAC
0˚
South CAP
America SOM NFB
NAZ LB
Australia
30˚S
0 20 40 60 80 100 120 140 160 180 200 220 240 260 280
Age of Oceanic Lithosphere [m.y.]
Australia
ARA
SAM
SOM PAC
LB
NFB
PS
AFR
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170 180 190 200 210 220 230 240 280
Age of Oceanic Lithosphere [m.y.]
occurred at 23 Ma leading to the establishment of the Cocos and Nazca 4.10. 20–0 Ma (Figs. 27 and 28)
plates and initiation of the East Pacific Rise, Galapagos Spreading Centre
and Chile Ridge. Spreading in the South, Central and North Atlantic continued un-
In the Western Pacific, spreading in the West Philippine Basin abated for the last 20 million years. In the Caribbean, the Cayman
ceased at 38 Ma whereas spreading continued in the Celebes Sea. Trough continued to expand and develop, and the Chortis Block
The formation of the Caroline Sea occurred behind a rapidly south- moved over the Yucatan promontory. Westward motion of the North
ward migrating subduction zone. By 30 Ma, spreading initiated in American plate relative to the slow moving Caribbean plate was accom-
the Shikoku and Parece Vela Basins behind the west-dipping Izu- modating the opening of the Cayman Trough. Active subduction of At-
Bonin–Mariana Arc. Spreading terminated in the Celebes Sea. In the lantic oceanic lithosphere has been occurring along the Lesser Antilles
SW Pacific, spreading initiated in the Solomon Sea at 40 Ma and in Trench, which connects to the Mid-Atlantic Ridge along the Researcher
the South Fiji Basin at 35 Ma. Cessation of spreading in the South Ridge and Royal Trough.
Fiji Basin occurred at 25 Ma. In the Pacific, spreading was occurring along the Pacific–Juan De
In the Indian Ocean, spreading continued along the Southwest Indian Fuca, Pacific–Nazca, Pacific–Cocos, Cocos–Nazca, Pacific–Antarctic
Ridge, Southeast Indian Ridge, Central Indian Ridge and Carlsberg Ridge. and Nazca–Antarctic ridges. The Bauer microplate formed along the
Extension along the East Africa rifts was established at 30 Ma leading to East Pacific Rise at 17 Ma and continued until 6 Ma. The locus of
the break-up of Africa into Somalia plate. Rifting along the Sheba Ridge, spreading then jumped back to the East Pacific Rise (between the
separating Arabia from Africa/Somalia initiated at 30 Ma. Pacific and Nazca plates). The East Pacific Rise is the fastest spreading
262 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
ridge system (excluding back-arc opening) and currently encompasses the plateau break-up has consequences for the evolution of the Phoenix
microplate formation at the Easter, Juan Fernandez and Galapagos plate and the eastern Gondwana margin. Most Mesozoic models for east-
plates. Currently, the Juan De Fuca plate is limited at its southern end ern Gondwana propose a long-lived convergent plate margin along the
by the Mendocino Fracture Zone and is subducting slowly along the eastern edge of Australia (Veevers, 2006; Cluzel et al., 2010; Matthews
Cascadia subduction zone. et al., 2011), expressed through andesitic volcanism that occurred
The western Pacific is dominated by the opening of a series of along the Queensland margin north to Papua New Guinea (Jones and
back-arc basins due to the roll-back of the subduction hinge of the Veevers, 1983) and Aptian–Albian andesitic volcanogenic detritus in
Tonga–Kermadec and Izu-Bonin–Mariana trenches. Spreading in the east Australian continental basins (e.g. Eromanga and Surat Basins)
Shikoku and Parece Vela Basins and South China Sea ceased at (Hawlader, 1990; Veevers, 2006). Plate velocity vectors using either
15 Ma. By 9 Ma, spreading initiated in the Mariana Trough. We Gurnis et al. (2012) or this study, predict a convergent margin between
model complete closure of the proto-South China Sea at around the Phoenix plate and eastern Gondwana during this time (Fig. 29).
10 Ma behind a subduction zone located along Palawan and the There is ambiguity as to whether the margin continued as a convergent
north Borneo/Kalimantan margin. In the SW Pacific, spreading in margin or whether there was a major tectonic regime change after
the Lau Basin initiated by 7 Ma with back-arc extension occurring in ~120 Ma, coincident with the eruption of the Ontong-Java, Manihiki
the Havre Trough. and Hikurangi plateaus and subsequent change in the mid ocean ridge
In the Indian Ocean, diffuse deformation occurring in the middle configuration in southern Panthalassa. Extensive magmatism recorded
of the Indo-Australian plate led to the development of the Capricorn in the Whitsunday Volcanic Province is attributed to continental margin
plate in the central–east Indian Ocean at 20 Ma. Further west, we ini- break-up rather than from a convergent margin setting (Bryan et al.,
tiate spreading along the Sheba Ridge at 20 Ma. The Sheba Ridge 1997) while others invoke a rift-related volcanics associated with
propagated into the Red Sea at 15 Ma. west-dipping subduction (Veevers, 2006). New Caledonia and parts of
New Zealand, which were located at the easternmost boundary of the
5. Discussion Australian continent record subduction related magmatism until at
least 99 Ma (Veevers, 2006) or 95 Ma (Cluzel et al. 2010) suggesting con-
5.1. Comparison with other models vergence was occurring along eastern Gondwana. Although the plate
motion model of Gurnis et al. (2012) does not include the rotations asso-
Our plate motion model offers an alternative approach to tradi- ciated with the plateau break-up, both models predict continuing con-
tional global plate reconstructions. Tectonic features that reside on vergence until 100 Ma (Fig. 29).
the surface of the Earth are not modeled as discrete features but rath- At 100–99 Ma, a major tectonic regime change is recorded in east-
er the plates themselves are modeled as dynamically evolving fea- ern Australia (Veevers, 2006). Sedimentation in the east Australian
tures. The nature of the plate boundaries that combine to form a basins changed from volcanogenic dominated to quartzose sandstone
plate will necessarily change based on the magnitude and direction (Veevers, 2006), the basins themselves changed from a prolonged
of motion of each plate. Therefore, one of the supplementary out- period of subsidence to uplift (Matthews et al., 2011) and volcanism
comes of this approach is the ability to directly compare competing became alkalitic (Veevers, 2006). In addition, the eastern margin
tectonic models, most easily expressed through plate velocity vectors changed to a period of extension and passive margin formation (e.g. ex-
for a common set of points on the surface of the Earth. We directly tension in the Lord Howe Rise and New Caledonia Basins), which are
compare the plate motion model presented in Gurnis et al. (2012) believed to have formed adjacent to a strike-slip margin defining the
to the model presented in this study (Fig. 29). boundary between Panthalassa and eastern Gondwana (Jones and
In this study we have adopted a new absolute plate motion model Veevers, 1983; Veevers, 2006). A hiatus in subduction-related volca-
for Africa for times prior to 100 Ma based on a true-polar wander cor- nism in Eastern Australia, New Caledonia and New Zealand is recorded
rected paleomagnetic reference frame (Steinberger and Torsvik, between 95 and 83 Ma (Cluzel et al., 2010). This major tectonic regime
2008). This new reference frame allows us to extend our plate recon- change is coincident with a change in spreading direction in the
structions back to 200 Ma, the time of Pangea break-up, with the po- plates associated with the plateau break-up and bordering the eastern
tential to model processes occurring during supercontinent break-up Gondwana margin at this time. The result is that the eastern Gondwana
and dispersal. The Gurnis et al. (2012) dataset was restricted to the margin changes from convergent to strike-slip, as predicted by geolog-
past 140 million years. Adjusting the absolute reference frame causes ical observations. This is in contrast to the model of Gurnis et al. (2012)
a global shift in the absolute positioning of the continents but in the- which suggests oblique convergence after 100–99 Ma (Fig. 29). In our
ory, should not affect the relative motion and therefore the nature of current plate motion model, a strike-slip dominated margin is predicted
the plate boundary between plates. However prior to 83.5 Ma, the Pa- from 100 to 86 Ma, which marks the timing of Hikurangi plateau colli-
cific plate can no longer link to the African plate circuit via seafloor sion with the Chatham Rise and the cessation of mid ocean ridge sub-
spreading (see Section 2: Methodology) requiring a distinct absolute duction related to the plateau break-up. The plate adjacent to eastern
reference frame for the Pacific realm. As a result, a change in the ab- Australia became the Pacific plate and all subsequent motions have
solute reference frame for either the African or Pacific realms will been between the Pacific and Australian or Lord Howe Rise plates.
change the nature of the plate boundaries that border the Pacific/ Additional differences between the relative plate motions pre-
Panthalassic Ocean (Fig. 29). sented in Gurnis et al. (2012) and this study include an updated
Relative motions between most of the plates in Panthalassa have northern Atlantic based on Gaina et al. (2009) and the Arctic based
been updated compared to the Gurnis et al. (2012) model. We reinter- on Alvey et al. (2008). The changes here are minor adjustments and
preted the M-series Japanese magnetic lineations leading to a dramatic do not substantially change plate motion directions or the nature of
change in spreading direction by about 24° and an updated orientation the plate boundaries in the area.
of the Izanagi–Farallon and Izanagi–Phoenix ridges. The change in the
Izanagi plate motion results in an increase in the convergence rate 5.2. Future directions
and more orthogonal convergence in northern Panthalassa bordering
eastern Laurasia but more oblique convergence in the area further Our global plate motion model presents the development of the
south adjacent to the Junction plate (Fig. 29). continents and oceans on a global scale within a rigid plate frame-
Another major addition to the model presented in this study is the work, underpinned by a combination of marine geophysical data, on-
implementation of the plateau break-up model of Taylor (2006) for the shore geological data and plate tectonic principles. Although we have
Ontong-Java, Manihiki and Hikurangi plateaus (Fig. 29). Incorporating presented our preferred interpretations for each region based on
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 263
120 Ma 90 Ma
Eurasia Eurasia
Izanagi
Izanagi
JUN JUN
India
Australia
Australia
HIK
HIK
Antarctica Antarctica
CAT
CAT
Fig. 29. Global comparison between the Gurnis et al. (2012) plate motion model and the one presented in this study, centered on Australia and the western Panthalassic margin.
Reconstructions are shown at 120 and 90 Ma with red plate velocity vectors denoting the Gurnis et al. (2012) model and blue plate velocity vectors from this study. Dark green line
indicates the east Australian margin.
available data, there are regions that could benefit from re-analysis of for example their Pacific hotpot models, combined with a relative
the seafloor spreading and break-up history, which will have a sig- plate motion model for motion between the Farallon and Pacific
nificant flow-on effect further down the global plate circuit. These plates, result in transform motion between the Farallon and North
include: American plates, while geological observations indicate subduction
being active (DeCelles, 2004). This model also leads to an anomalous
(1) The early break-up history between Africa and South America
amount of material entering the mantle in the southern hemisphere
to account for significant overlaps and gaps between the two
(Shephard et al., 2012). This inconsistency may result from the as-
margins. Refining the history between these two plates will
sumption of Pacific hotspot fixity and poor sampling of Pacific sea-
lead to a revision of the Mesozoic history of the Caribbean re-
mount chains due to a paucity of available data. A new approach
gion (i.e. the accommodation space created to form the
using a combination of methods, for example moving hotspot models,
proto-Caribbean Sea and the rift basins associated with
paleomagnetics and coupled geodynamic-plate motion models, may
hydrocarbon-bearing basins in the Gulf of Mexico), a more
result in a more robust model for the Pacific plate prior to ~ 83 Ma
tightly constrained equatorial Atlantic and also the plate
and may potentially extend the Pacific absolute reference frame to
boundaries surrounding the Weddell Sea, which are very ill-
the earliest Mesozoic.
constrained due to a paucity of data.
A further limitation of the present model is that the entire surface
(2) The early break-up history and Mesozoic spreading between
of the earth is represented as rigid blocks, which is clearly not true for
Africa and Antarctica. A further refinement of the opening his-
some plate interiors and plate boundaries (Gordon and Stein, 1992;
tory of this area will affect the motions of Antarctica, India and
Bird, 2003). Deforming regions within plate interiors or straddling
Australia and the interaction (plate boundary processes) along
plate boundaries will clearly be required for reconstructions beyond
the eastern Gondwana margin bordering Panthalassa.
those presented here. For future models, deforming regions can
(3) The break-up history of the Pacific–Marie Byrd Land margin
now be encompassed within the domain of an evolving, closed poly-
(~100–83 Ma), which has consequences for the motion of the
gon and consequently incorporated as an extension of the CCP algo-
Pacific plate and associated plates, such as the Izanagi, Phoenix,
rithm (see Gurnis et al., 2012). We expect that such deforming
Farallon, Hikurangi, Manihiki, Catequil and Chasca plates. The
regions will be represented as deforming meshes within continuously
Pacific plate can only be linked to the plate circuit, through
closing polygons as the lowest level of a global hierarchy. Such func-
Africa, when there is a mid-ocean ridge (or rift) between the
tionality has now been incorporated in experimental versions of
Pacific and Antarctica/Marie Byrd Land. Greater constraints
GPlates and will be a part of a new generation of global plate recon-
on the timing of break-up between the Campbell Plateau and
structions. The first region to be addressed within a deforming plate
Antarctica and a revised set of finite rotations to describe the
network is the opening of the rift basins within the interior of Africa
opening will potentially mean we can confidently extend the
as the accounting of this extension will have flow-on effects for all
Pacific plate's link to the plate circuit further back in time and
the plates that hang-off the African-centered plate circuit.
decrease the uncertainty in Pacific plate motion during this
time interval.
6. Conclusions
A major improvement that is essential for global plate motion
models that extend into the Mesozoic is a more robust Pacific abso- There are currently three main types of plate motion models that
lute plate motion model. The latest models available with associated enable us to place features on the surface of the earth into their
published rotation poles for Pacific hotspots (Wessel et al. 2006; spatio-temporal context. Geologically-current plate motion models
Wessel and Kroenke 2008) result in major shifts and rotations of are ideal because they provide a set of plate velocity vectors and de-
the Pacific plate, which are inconsistent with geological observations; lineate the boundaries between tectonic plates in a self-consistent
264 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
way (i.e. the combined area of the plates equals the area of the Earth). Atwater, T., Severinghaus, J., 1990. Tectonic maps of the northeast Pacific. In: Winterer,
E.L., Hussong, D.M., Decker, R.W. (Eds.), The Eastern Pacific Ocean and Hawaii,
However, they are restricted to the Pliocene, making analysis of su- Volume N: The Geology of North America. Geological Society of America, Boulder,
percontinent break-up and accretion, the linkages between the deep Colorado, pp. 15–20.
earth and surface processes and larger-scale tectonic cycles unrealis- Atwater, T., Sclater, J., Sandwell, D., 1993. Fracture zone traces across the North Pacific
Cretaceous Quiet Zone and their tectonic implications. In: Pringle, M.S., et al. (Ed.),
tic. Traditional plate motion models do not treat plates in a self- Mesozoic Pacific: Geology, Tectonics, and Volcanism. Geophysical Monograph
consistent way but rather reconstruct discrete features on the surface Series, vol. 77. AGU, Washington, D.C., pp. 137–154.
of the Earth without regard to the evolving nature of plate bound- Badarch, G., Dickson Cunningham, W., Windley, B.F., 2002. A new terrane subdivision
for Mongolia: implications for the Phanerozoic crustal growth of Central Asia. Journal
aries. Coupled geodynamic models are prone to large uncertainties
of Asian Earth Sciences 21, 87–110.
and have not been successful at replicating past plate motions consis- Baillie, P.W., Jacobson, E., 1995. Structural evolution of the Carnarvon Terrace, Western
tently in deep time. Australia. The APEA Journal 35, 321–332.
Banerjee, B., Sengupta, B., Banerjee, P., 1995. Signals of Barremian (116 Ma) or younger
In this paper, we have presented a new type of global plate motion
oceanic crust beneath the Bay of Bengal along 14 N latitude between 81 E and 93 E.
model, which extends into deep time and involves a continuously Marine Geology 128, 17–23.
evolving and self-consistent set of plate polygons and plate bound- Banks, N., Bardwell, K., Musiwa, S., 1995. Karoo rift basins of the Luangwa Valley, Zambia.
aries from the time of Pangea break-up. Our model is underpinned Geological Society London Special Publications 80, 285.
Barckhausen, U., Ranero, C.R., Cande, S.C., Engels, M., Weinrebe, W., 2008. Birth of an
by a detailed analysis of the seafloor spreading record for the major intraoceanic spreading center. Geology 36, 767.
tectonic plates. Our regional models are built within a hierarchical Beauchamp, W.H., 1998. Tectonic evolution of the Atlas Mountains, North Africa.
plate circuit framework linked to a hybrid absolute reference frame Belasky, P., Runnegar, B., 1994. Permian longitudes of Wrangellia, Stikinia, and Eastern
Klamath terranes based on coral biogeography. Geology 22, 1095–1098.
that includes moving Indian/Atlantic hotspots and a true polar wan- Bernard, A., Munschy, M., 2000. Were the Mascarene and Laxmi Basins (western Indian
der corrected paleomagnetic-based model. Ocean) formed at the same spreading centre? Comptes Rendus de l'Academie des
The plate motion model presented in this study will be of particu- Sciences Series IIA Earth and Planetary Science 330, 777–783.
Binks, R., Fairhead, J., 1992. A plate tectonic setting for Mesozoic rifts of West and Cen-
lar use to geodynamicists who require surface boundary conditions tral Africa. Tectonophysics 213, 141–151.
for the motions of the plates through time to link to models of the Billen, M.I., Stock, J., 2000. Morphology and origin of the Osbourn Trough. Journal of
convecting mantle. However, our hope is that it can also be used as Geophysical Research-Solid Earth 105, 13481–13489.
Bird, P., 2003. An updated digital model of plate boundaries. Geochemistry, Geophysics,
a framework for further detailed work so that we may converge to- Geosystems 4, 1027. doi:10.1029/2001GC000252.
wards an ever-improved set of global plate reconstructions. We pro- Bird, R.T., Naar, D.F., 1994. Intrafransform origins of mid-ocean ridge microplates. Geology
vide all data freely in digital form, welcome feedback to improve 22, 987–990.
Bird, R.T., Naar, D.F., Larson, R.L., Searle, R.C., Scotese, C.R., 1998. Plate tectonic recon-
our models and anticipate that refinements to the plate model will
structions of the Juan Fernandez microplate: transformation from internal shear
be published in the future. The plate polygon data files with associat- to rigid rotation. Journal of Geophysical Research-Solid Earth 103, 7049–7067.
ed rotation file and an accompanying coastline and continent–ocean Bird, D., Hall, S., Burke, K., Casey, J., Sawyer, D., 2007. Early Central Atlantic Ocean sea-
boundary file can be downloaded from the following location: ftp:// floor spreading history. Geosphere 3, 282.
Blakely, R., 2008. Gondwana paleogeography from assembly to breakup: A 500 my od-
ftp.earthbyte.org/papers/Seton_etal_Global_ESR/Seton_etal_Data.zip. yssey. Resolving the late Paleozoic ice age in time and space, p. 1.
Boillot, G., Girardeau, J., Kornprobst, J., 1988. Rifting of the Galicia margin: crustal thin-
ning and emplacement of mantle rocks on the seafloor. In: Goillot, G., Winterer, E.L.
Acknowledgments (Eds.), Proceedings of the Ocean Drilling Program, Scientific Results 103, 741–756.
Boillot, G., Winterer, E., 1988. Drilling on the Galicia margin: retrospect and prospect.
Proceedings of the Ocean Drilling Program, Scientific Results 103, 809–828.
We thank Roi Granot and an anonymous reviewer for agreeing to Boulin, J., 1988. Hercynian and Eocimmerian events in Afghanistan and adjoining re-
review such a lengthy manuscript and for their thoughtful and careful gions. Tectonophysics 148, 253–278.
review, which greatly improved the manuscript. We would also like Boyden, J.A., Müller, R.D., Gurnis, M., Torsvik, T.H., Clark, J.A., Turner, M., Ivey-Law, H.,
Watson, R.J., Cannon, J.S., 2011. Next-generation plate-tectonic reconstructions
to thank members of the EarthByte Group and the group at Caltech using GPlates. In: Keller, G.R., Baru, C. (Eds.), Geoinformatics: Cyberinfrastructure
led by Michael Gurnis who have contributed over many years towards for the Solid Earth Sciences. Cambridge University Press, pp. 95–114.
the continuous improvement of the global plate motion model and Bradley, D.C., Haeussler, P.J., Kusky, T.M., 1993. Timing of early Tertiary ridge subduc-
tion in southern Alaska. US Geological Survey Bulletin 2068, 163–177.
associated files. This project was funded through Australian Research
Bradley, D., 2008. Passive margins through Earth history. Earth-Science Reviews 91, 1–26.
Council grants FL0992245 and DP0987713. Bradshaw, M.T., Yeates, A.N., Beynon, R.M., Brakel, A.T., Langford, R.P., Totterdell, J.M.,
Yeung, M., 1988. Paleogeographic evolution of the North West Shelf region. In:
Purcell, P.G., Purcell, R.R. (Eds.), The North West Shelf of Australia: Petroleum Ex-
References ploration Society of Australia Symposium: Perth, pp. 29–53.
Breitsprecher, K., Thorkelson, D., Groome, W., Dostal, J., 2003. Geochemical confirma-
Agar, S., Cliff, R., Duddy, I., Rex, D., 1989. Short Paper: Accretion and uplift in the Shi- tion of the Kula-Farallon slab window beneath the Pacific Northwest in Eocene
manto Belt, SW Japan. Journal of the Geological Society 146, 893. time. Geology 31, 351.
Aitchison, J.C., Ali, J.R., Davis, A.M., 2007. When and where did India and Asia collide. Brekke, H., 2000. The tectonic evolution of the Norwegian Sea continental margin with
Journal of Geophysical Research 112, B05423. emphasis on the Voring and More basins. Geological Society London Special Publi-
Alvey, A., Gaina, C., Kusznir, N., Torsvik, T., 2008. Integrated crustal thickness mapping cations, v. 167, p. 327.
and plate reconstructions for the high Arctic. Earth and Planetary Science Letters Brozena, J., Childers, V., Lawver, L., Gahagan, L., Forsberg, R., Faleide, J., Eldholm, O.,
274, 310–321. 2003. New aerogeophysical study of the Eurasia Basin and Lomonosov Ridge: im-
Apel, E., B¸Rgmann, R., Steblov, G., Vasilenko, N., King, R., Prytkov, A., 2006. Independent plications for basin development. Geology, 31, p. 825.
active microplate tectonics of northeast Asia from GPS velocities and block modeling. Bryan, S.E., Constantine, A.E., Stephens, C.J., Ewart, A., Schon, R.W., Parianos, J., 1997.
Geophysical Research Letters 33, L11303. Early Cretaceous volcano-sedimentary successions along the Eastern Australian
Argus, D.F., Heflin, M.B., 1995. Plate motion and crustal deformation estimated with continental margin — implications for the break-up of Eastern Gondwana. Earth
geodetic data from the Global Positioning System. Geophysical Research Letters and Planetary Science Letters 153, 85–102.
22, 1973–1976. Buchs, D.M., Arculus, R.J., Baumgartner, P.O., Baumgartner-Mora, C., Ulianov, A., 2010.
Argus, D.F., Gordon, R.G., Heflin, M.B., Ma, C., Eanes, R.J., Willis, P., Peltier, W.R., Owen, Late Cretaceous arc development on the SW margin of the Caribbean Plate: in-
S.E., 2010. The angular velocities of the plates and the velocity of Earth's centre sights from the Golfito, Costa Rica, and Azuero, Panama, complexes. Geochemistry,
from space geodesy. Geophysical Journal International 180, 913–960. Geophysics, Geosystems 11, Q07S24.
Atwater, T., 1970. Implications of plate tectonics for the Cenozoic tectonic evolution of Buffler, R.T., Sawyer, D.S., 1983. Distribution of crust and early history, Gulf of Mexico
Western North America. Geological Society of America Bulletin 81, 3513–3536. basin. Gulf Coast Association Geological Society Transcripts 35, 334–344.
Atwater, T., 1990. Plate tectonic history of the northeast Pacific and western North Bunce, E.T., Molnar, P., 1977. Seismic reflection profiling and basement topography in
America. In: Winterer, E.L., Hussong, D.M., Decker, R.W. (Eds.), The Eastern Pacific the Somali Basin: possible fracture zones between Madagascar and Africa. Journal
Ocean and Hawaii, Volume N: The Geology of North America. Geological Society of of Geophysical Research 82, 5305–5311.
America, Boulder, Colorado, pp. 21–72. Bunge, H., Grand, S., 2000. Mesozoic plate-motion history below the northeast Pacific
Atwater, T., 1989. Magnetic Anomalies in the North Pacific. Geological Society of America: Ocean from seismic images of the subducted Farallon slab. Nature 405, 337–340.
Geology of North America. Burke, K., 1988. Tectonic evolution of the Caribbean. Annual Review of Earth and Plan-
Atwater, T., Severinghaus, J., 1989. Magnetic Anomalies in the North Pacific: Geological etary Sciences 16, 201–230.
Society of America, v. Geology of North America. Byrne, T., 1979. Late Paleocene demise of the Kula–Pacific spreading center. Geology 7, 341.
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 265
Cande, S.C., Haxby, W.F., 1991. Eocene propagating rifts in the Southwest Pacific and their Davy, B., Hoernle, K., Werner, R., 2008. Hikurangi Plateau: Crustal structure, rifted for-
conjugate features on the Nazca Plate. Journal of Geophysical Research 96, 19609. mation, and Gondwana subduction history. Geochemistry, Geophysics, Geosys-
Cande, S.C., Herron, E.M., Hall, B.R., 1982. The early Cenozoic tectonic history of the tems 9, Q07004.
southeast Pacific. Earth and Planetary Science Letters 57, 63–74. De Graciansky, P., Poag, C., Cunningham Jr., R., Loubere, P., Masson, D., Mazzullo, J.,
Cande, S.C., Kent, D.V., 1992. A new geomagnetic polarity time scale for Late Cretaceous Montadert, L., Muller, C., Otsuka, K., Reynolds, L., 1985. The Goban Spur transect:
and Cenozoic. Journal of Geophysical Research 97, 13917–13951. geologic evolution of a sediment-starved passive continental margin. Bulletin of
Cande, S.C., Kent, D.V., 1995. Revised calibration of the geomagnetic polarity timescale for the Geological Society of America 96, 58.
the Late Cretaceous and Cenozoic. Journal of Geophysical Research 100, 6093–6095. Decelles, P., 2004. Late Jurassic to Eocene evolution of the Cordilleran thrust belt and
Cande, S.C., Larson, R.L., LaBrecque, J.L., 1978. Magnetic lineations in the Pacific Jurassic foreland basin system, western USA. American Journal of Science 304, 105.
Quiet Zone. Earth and Planetary Science Letters 41, 434–440. Demets, C., Gordon, R.G., Argus, D.F., Stein, S., 1990. Current plate motions. Geophysical
Cande, S.C., Mutter, J.C., 1982. A revised identification of the oldest sea-floor spreading Journal International 101, 425–478.
anomalies between Australia and Antarctica. Earth and Planetary Science Letters Demets, C., Gordon, R.G., Argus, D.F., 2010. Geologically current plate motions. Geophysical
58, 151–160. Journal International 181, 1–80.
Cande, S., Labrecque, J., Haxby, W., 1988. Plate kinematics of the South Atlantic: Chron Demets, C., Gordon, R., Vogt, P., 1994. Location of the Africa-Australia-India triple junc-
C34 to present. Journal of Geophysical Research 93, 13479–13492. tion and motion between the Australian and Indian plates: Results from an aero-
Cande, S.C., Raymond, C.A., Stock, J., Haxby, W.F., 1995. Geophysics of the Pitman fracture magnetic investigation of the Central Indian and Carlsberg ridges. Geophysical
zone and Pacific–Antarctic plate motions during the Cenozoic. Science 270, 947–953. Journal International 119, 893–930.
Cande, S., Stock, J., Raymond, C., Müller, R.D., 1998. New constraints on an old plate tec- Desa, M., Ramana, M., Ramprasad, T., 2006. Seafloor spreading magnetic anomalies
tonic puzzle of the Southwest Pacific. Eos 79, 81–82. south off Sri Lanka. Marine Geology 229, 227–240.
Caress, D.W., Menard, H.W., Hey, R.N., 1988. Eocene reorganization of the Pacific-Faral- DiTullio, L., 1993. Geologic summary and conceptual framework for the study of ther-
lon Spreading Center north of the Mendocino Fracture Zone. Journal of Geophysical mal maturity within the Eocene-Miocene Shimanto Belt, Shikoku, Japan. Thermal
Research 93, 2813–2838. evolution of the Tertiary Shimanto Belt, southwest Japan: an example of ridge-
Carey, S.W., 1955. The orocline concept in geotectonics. Royal Society of Tasmania Pro- trench interaction, p. 1.
ceedings 89, 255–288. Dore, A., Lundin, E., Jensen, L., Birkeland, ÿ, Eliassen, P., Fichler, C., 1999. Principal tec-
Carminati, E., Doglioni, C., Scrocca, D., 2004. Alps vs Apennines. Special volume of the tonic events in the evolution of the northwest European Atlantic margin. Geologi-
Italian Geological Society for the IGC, vol. 32, pp. 141–151. cal Society of London 5, 41.
Catuneanu, O., Wopfner, H., Eriksson, P., Cairncross, B., Rubidge, B., Smith, R., Hancox, Downey, N.J., Stock, J.M., Clayton, R.W., Cande, S.C., 2007. History of the Cretaceous
P., 2005. The Karoo basins of south-central Africa. Journal of African Earth Sciences Osbourn spreading center. Journal of Geophysical Research 112, B04102.
43, 211–253. Dove, D., Coakley, B., Hopper, J., Kristoffersen, Y., 2010. Bathymetry, controlled source
Chalmers, J.A., 1991. New evidence on the structure of the Labrador Sea/Greenland seismic and gravity observations of the Mendeleev ridge; implications for ridge
continental margin. Journal of the Geological Society of London 148, 899–908. structure, origin, and regional tectonics. Geophysical Journal International 1, 348.
Chalmers, J.A., Laursen, K.H., 1995. Labrador Sea: the extent of continental and oceanic Dunbar, J., Sawyer, D., 1989. Patterns of continental extension along the conjugate mar-
crust and the timing of the onset of seafloor spreading. Marine and Petroleum gins of the central and North Atlantic Ocean and Labrador Sea. Tectonics 8.
Geology 12, 205–217. Duncan, R., Hargraves, R., 1984. Plate tectonic evolution of the Caribbean region in
Chalmers, J., Pulvertaft, T., 2001. Development of the continental margins of the Labrador the mantle reference frame. The Caribbean-South American Plate Boundary and
Sea: a review. Geological Society London Special Publications 187, 77. Regional Tectonics, p. 81ñ93.
Chandler, M., Wessel, P., Seton, M., Taylor, B., Hyeong, K., Kim, S., in press. Reconstruct- Eagles, G., 2007. New angles on South Atlantic opening. Geophysical Journal Interna-
ing Ontong Java Nui: Implications for Pacific absolute plate motion, hotspot drift tional 168, 353–361.
and true polar wander. Earth and Planetary Science Letters. Eagles, G., König, M., 2008. A model of plate kinematics in Gondwana breakup. Geophysical
Channell, J.E.T., 1995. Recalibration of the geomagnetic polarity timescale. Reviews of Journal International 173, 703–717.
Geophysics 33, 161–168. Eagles, G., Gohl, K., Larter, R.D., 2004a. High-resolution animated tectonic reconstruction
Chaubey, A.K., Bhattacharya, G.C., Murty, G.P.S., Srinivas, K., Ramprasad, T., Rao, D.G., 1998. of the South Pacific and West Antarctic margin — art. no. Q07002. Geochemistry,
Early Tertiary seafloor spreading magnetic anomalies and paleo-propagators in the Geophysics, Geosystems 5 7002-7002.
Northern Arabian Sea. Earth and Planetary Science Letters 154, 41–52. Eagles, G., Gohl, K., Larter, R.D., 2004b. Life of the Bellingshausen plate. Geophysical
Churkin, M., Trexler, J.H., 1980. Circum-Arctic plate accretion-isolating part of a Pacific Research Letters 31 7603-7603.
plate to form the nucleus of the Arctic Basin. Earth and Planetary Science Letters Eakins, B.W., Lonsdale, P.F., 2003. Structural patterns and tectonic history of the Bauer
49, 356–362. microplate, Eastern Tropical Pacific. Marine Geophysical Researches 24, 171–205.
Cluzel, D., Adams, C., Meffre, S., Campbell, H., Maurizot, P., 2010. Discovery of Early Cre- Engebretson, D.C., Cox, A., Gordon, R.G., 1985. Relative Motions between Oceanic and
taceous rocks in New Caledonia; new geochemical and U–Pb zircon age constraints Continental Plates in the Pacific Basin. Geol. Soc. of Am.
on the transition from subduction to marginal breakup in the Southwest Pacific. Faccenna, C., Becker, T.W., Lucente, F.P., Jolivet, L., Rossetti, F., 2001. History of subduc-
Cocks, L.R.M., Torsvik, T.H., 2007. Siberia, the wandering northern terrane, and its tion and back arc extension in the Central Mediterranean. Geophysical Journal In-
changing geography through the Palaeozoic. Earth-Science Reviews 82, 29–74. ternational 145, 809–820.
Cochran, J.R., 1988. The Somali Basin, Chain Ridge and the origin of the northern Somali Falvey, D.A., Mutter, J.C., 1981. Regional plate tectonics and the evolution of Australia's
Basin gravity and geoid low. Journal of Geophysical Research 93, 11985–12008. passive margins. Journal of Australian Geology and Geophysics 6, 1–29.
Coffin, M.F., Rabinowitz, P.D., 1987. Reconstruction of Madagascar and Africa: evidence Forster, R., 1975. The geological history of the sedimentary basins of southern Mozam-
from the Davie Fracture Zone and western Somali Basin. Journal of Geophysical Re- bique, and some aspects of the origin of the Mozambique Channel. Palaeogeogra-
search 92, 9385–9406. phy, Palaeoclimatology, Palaeoecology 17, 267–287.
Coffin, M.F., Rabinowitz, P.D., 1988. Evolution of the East African continental margin Forsyth, D.A., Asudeh, I., Green, A.G., Jackson, H.R., 1986. Crustal structure of the north-
and the Western Somali Basin. 78 pp. ern Alpha Ridge beneath the Arctic Ocean. Nature 322, 349–352.
Cole, J., Peachey, J., 1999. Evidence for Pre-Cretaceous Rifting in the Rockall Trough: An Fullerton, L.G., Sager, W.W., Handschumacher, D.W., 1989. Late Jurassic–Early Cretaceous
Analysis using Quantitative Plate Tectonic Modelling. Geological Society Pub evolution of the eastern Indian Ocean adjacent to northwest Australia. Journal of
House, p. 359. Geophysical Research 94, 2937–2953.
Collier, J., Sansom, V., Ishizuka, O., Taylor, R., Minshull, T., Whitmarsh, R., 2008. Gaina, C., Müller, R.D., 2007. Cenozoic tectonic and depth/age evolution of the Indone-
Age of Seychelles–India break-up. Earth and Planetary Science Letters 272, sian gateway and associated back-arc basins. Earth-Science Reviews 83, 177–203.
264–277. Gaina, C., Müller, D.R., Royer, J.Y., Stock, J., Hardebeck, J., Symonds, P., 1998. The tectonic
Colpron, M., Logan, J.M., Mortensen, J.K., 2002. U–Pb zircon age constraint for late history of the Tasman Sea: a puzzle with 13 pieces. Journal of Geophysical Research
Neoproterozoic rifting and initiation of the lower Paleozoic passive margin of 103, 12413–12433.
western Laurentia. Canadian Journal of Earth Sciences 39, 133–143. Gaina, C., Müller, R.D., Royer, J.-Y., Symonds, P., 1999. The tectonic evolution of the
Colpron, M., Nelson, J.A.L., Murphy, D.C., 2007. Northern Cordilleran terranes and their Louisiade Triple Junction. Journal of Geophysical Research 104, 12927–12939.
interactions through time. GSA Today 17, 4–10. Gaina, C., Roest, W.R., Müller, R.D., 2002. Late Cretaceous–Cenozoic deformation of
Conrad, C., Lithgow-Bertelloni, C., 2002. How mantle slabs drive plate tectoncs. Science northeast Asia. Earth and Planetary Science Letters 197, 273–286.
298, 207–209. Gaina, C., Müller, R.D., Brown, B., Ishihara, T., 2003. Micro-continent formation
Cox, K.G., 1992. Karoo igneous activity, and the early stages of the break-up of Gondwa- around Australia. In: Hillis, R., Müller, R.D. (Eds.), The Evolution and Dynamics
naland. In: Storey, B.C., Alabaster, T., Pankhurst, R.J. (Eds.), Magmatism and the of the Australian Plate, vol. 22. Geol. Soc. Aust. Spec. Publ., Geological Society
causes of continental break-up. : Geological Society Special Publication, 68. Geo- of Australia.
logical Society, London, pp. 137–148. Gaina, C., Müller, R.D., Brown, B., Ishihara, T., Ivanov, S., 2007. Breakup and early sea-
Crawford, T., Meffre, S., Symonds, P.A., 2002. Tectonic Evolution of the SW Pacific: lessons floor spreading between India and Antarctica. Geophysical Journal International
for the geological evolution of the Tasman fold belt system in eastern Australia 170, 151–169.
from 600 to 220 Ma. In: Hillis, R., Müller, R.D. (Eds.), Evolution and Dynamics of the Gaina, C., Gernigon, L., Ball, P., 2009. Palaeocene–Recent plate boundaries in the NE
Australian Plate. Atlantic and the formation of the Jan Mayen microcontinent. Journal of the Geological
Croon, M.B., Cande, S.C., Stock, J.M., 2008. Revised Pacific–Antarctic plate motions Society 166, 601.
and geophysics of the Menard Fracture Zone. Geochemistry, Geophysics, Geo- Gaina, C., Labails, C., Reeves, C., 2010. The Early Opening of the Indian Ocean: An African
systems 9, 7. Perspective, 2010 Fall Meeting, Volume Abstract T13C-2222. AGU, San Francisco,
Daly, M., Chorowicz, J., Fairhead, J., 1989. Rift basin evolution in Africa: the influence California. 13–17 Dec.
of reactivated steep basement shear zones. Geological Society London Special Gee, J., Kent, D., 2007. Source of Oceanic magnetic anomalies and the geomagnetic
Publications 44, 309. polarity timescale. Treatise on Geophysics 5, 455ñ507.
266 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
Gehrels, G.E., 2001. Geology of the Chatham Sound region, southeast Alaska and coastal Heine, C., Muller, R., 2005. Late Jurassic rifting along the Australian North West Shelf:
British Columbia. Canadian Journal of Earth Sciences 38, 1579–1599. margin geometry and spreading ridge configuration. Australian Journal of Earth
Gehrels, G.E., 2002. Detrital zircon geochronology of the Taku terrane, southeast Alaska. Sciences 52 (1), 27–39.
Canadian Journal of Earth Sciences 39, 921–931. Heine, C., Müller, R.D., Gaina, C., 2004. Reconstructing the lost Eastern Tethys Ocean
Genik, G., 1992. Regional framework, structural and petroleum aspects of rift basins in Basin: constraints for the convergence history of the SE Asian margin and marine
Niger, Chad and the Central African Republic (CAR). Tectonophysics 213, 169–185. gateways. In: Clift, P., Hayes, D., Kuhnt, W., Wang, P. (Eds.), Continent–Ocean Interac-
Gerstell, M., Stock, J., 1994. Testing the porcupine plate hypothesis. Marine Geophysical tions in Southeast Asia, Volume 1149: Geophys. Monogr. Ser. American Geophysical
Researches 16, 315–323. Union, Washington.
Ghidella, M.E., Yanez, G., LaBrecque, J.L., 2002. Revised tectonic implications for the Heirtzler, J., Burroughs, R., 1971. Madagascar's paleoposition: new data from the
magnetic anomalies of the western Weddell Sea. Tectonophysics 347, 65–86. Mozambique channel. Science (New York, N.Y.) 174, 488.
Ghosh, N., Hall, S., Casey, J., 1984. Seafloor spreading magnetic anomalies in the Heirtzler, J.R., Dickson, G.O., Herman, E.M., Pitman III, W.C., Lepichon, X., 1968. Marine
Venezuelan Basin. The Caribbean–South American Plate Boundary and Regional magnetic anomalies, geomagnetic field reversals, and motions of the ocean floor
Tectonics, pp. 65–80. and continents. Journal of Geophysical Research 73, 2119–2136.
Goff, J.A., Cochran, J.R., 1996. The Bauer scarp ridge jump: a complex tectonic sequence Herron, E.M., 1972. Sea Floor Spreading and the Cenozoic History of the East-Central
revealed in satellite altimetry. Earth and Planetary Science Letters 141, 21–33. Pacific. Geological Society of America Bulletin 83, 1671–1692.
Golonka, J., 2007. Late Triassic and Early Jurassic palaeogeography of the world. Hessami, K., Koyi, H.A., Talbot, C.J., Tabishi, H., Shabanian, E., 2001. Progressive uncon-
Palaeogeography, Palaeoclimatology, Palaeoecology 244, 297–307. formities within an evolving foreland fold–thrust belt, Zagros Mountains. Journal
Golonka, J., Ford, D., 2000. Pangean (Late Carboniferous–Middle Jurassic) paleoenvir- of the Geological Society 158, 969.
onment and lithofacies. Palaeogeography, Palaeoclimatology, Palaeoecology 161, Hey, R., 1977. Tectonic evolution of the Cocos–Nazca spreading center. Geological Society
1–34. of America Bulletin 88, 1404–1420.
Golonka, J., Gahagan, L., Krobicki, M., Marko, F., Oszczypko, N., Slaczka, A., 2006. Plate-tectonic Hillier, J., 2007. Pacific seamount volcanism in space and time. Geophysical Journal
evolution and paleogeography of the circum-Carpathian region: the Carpathians and International 168, 877–889.
their foreland: geology and hydrocarbon resources. AAPG Memoir 84, 11-46. Hoernle, K., Hauff, F., Van Den Bogaard, P., Werner, R., Mortimer, N., Geldmacher, J.,
Gordon, R.G., Stein, S., 1992. Global tectonics and space geodesy. Science 256, 333. Garbe-SchˆNberg, D., Davy, B., 2010. Age and geochemistry of volcanic rocks
Goslin, J., Schlich, R., 1976. Structural limits of the South Crozet Basin; relations to from the Hikurangi and Manihiki oceanic plateaus. Geochimica et Cosmochimica
Enderby Basin and the Kerguelen-Heard Plateau. International Geological Congress, Acta.
Abstracts, vol. 25 (3), pp. 884–885. Holloway, N., 1982. North Palawan block, Philippines—its relation to Asian mainland
Guiraud, R., Maurin, J.C., 1992. Early Cretaceous rifts of Western and Central Africa: an and role in evolution of South China Sea. American Association of Petroleum Geologists
overview. Tectonophysics 213, 153–168. Bulletin 66, 1355ñ1383.
Gradstein, F.M., Agterberg, F.P., Ogg, J.G., Hardenbol, S., Vanveen, P., Thierry, J., Huang, Hopper, J., Funck, T., Tucholke, B., Larsen, H., Holbrook, W., Louden, K., Shillington, D.,
Z.H., 1994. A Mesozoic time scale. Journal of Geophysical Research-Solid Earth Lau, H., 2004. Continental breakup and the onset of ultraslow seafloor spreading
99, 24051–24074. off Flemish Cap on the Newfoundland rifted margin. Geology 32, 93.
Gradstein, F., Ogg, J., Smith, A., Agterberg, F., Bleeker, W., Cooper, R., Davydov, V., Hutchison, C.S., 1989. Geological Evolution of South-east Asia, p. 368.
Gibbard, P., Hinnov, L., House, M., 2004. A Geologic Time Scale 2004. Cambridge Hutchison, C.S., 2005. Geology of North-west Borneo. Elsevier Science Ltd., Sarawak,
University Press. Brunei and Sabah.
Granot, R., Cande, S.C., Gee, J.S., 2009. The implications of long-lived asymmetry of rem- Jackson, H., Gunnarsson, K., 1990. Reconstructions of the Arctic: Mesozoic to present.
anent magnetization across the North Pacific fracture zones. Earth and Planetary Tectonophysics 172, 303–322.
Science Letters 288, 551–563. James, K., 2006. Arguments for and against the Pacific origin of the Caribbean Plate.
Grantz, A., Clark, D., Phillips, R., Srivastava, S., Blome, C., Gray, L., Haga, H., Mamet, B., Geologica Acta 4, 279.
McIntyre, D., Mcneil, D., 1998. Phanerozoic stratigraphy of Northwind Ridge, mag- Johnson, B.D., Powell, C.M., Veevers, J.J., 1976. Spreading history of the eastern Indian
netic anomalies in the Canada basin, and the geometry and timing of rifting in the Ocean and Greater India's northward flight from Antarctica and Australia. Geolog-
Amerasia basin, Arctic Ocean. Geological Society of America Bulletin 110, 801. ical Society of America Bulletin 87, 1560–1566.
Greene, A.R., Scoates, J.S., Weis, D., 2008. Wrangellia flood basalts in Alaska: a record of Johnson, B.D., Powell, C.M., Veevers, J.J., 1980. Early spreading history of the Indian
plume–lithosphere interaction in a Late Triassic accreted oceanic plateau. Geochemistry, Ocean between India and Australia. Earth and Planetary Science Letters 47,
Geophysics, Geosystems 9, 12004. 131–143.
Gurnis, M., Turner, M., Zahirovic, S., Dicaprio, L., Spasojevic, S., Müller, R.D., Boyden, J., Jokat, W., Uenzelmann-Neben, G., Kristoffersen, Y., Rasmussen, T., 1992. Lomonosov
Seton, M., Manea, V., Bower, D., 2012. Plate tectonic reconstructions with continu- Ridge—a double-sided continental margin. Geology 20, 887.
ously closing plates. Computers and Geosciences 38, 35–42. Jokat, W., Ritzmann, O., Schmidt-Aursch, M.C., Drachev, S.S., Gauger, S., Snow, J.H.,
Haeussler, P.J., Bradley, D., Goldfarb, R., Snee, L., Taylor, C., 1995. Link between ridge 2003. Geophysical evidence for reduced melt production on the Arctic ultraslow
subduction and gold mineralization in southern Alaska. Geology 23, 995. Gakkel mid-ocean ridge. Nature 423, 962–965.
Hafkenscheid, E., Buiter, S.J.H., Wortel, M.J.R., Spakman, W., Bijwaard, H., 2001. Modelling Jokat, W., Nogi, Y., Leinweber, V.T., 2010. New aeromagnetic data from the western
the seismic velocity structure beneath Indonesia: a comparison with tomography. Enderby Basin and consequences for Antarctic–India break-up. Geophysical
Tectonophysics 333, 35–46. Research Letters 37, L21311.
Hager, B.H., O'Connell, R.J., 1981. A simple global model of plate dynamics and mantle Jones, J.G., Veevers, J.J., 1983. Mesozoic origins and antecedents of Australia's Eastern
convection. Journal of Geophysical Research 86, 4843–4867. Highlands. Journal of the Geological Society of Australia 30, 305–322.
Halgedahl, S., Jarrard, R., 1987. Paleomagnetism of the Kuparuk River formation from Joseph, D., Shor, A., Hussong, D., Malikides, M., 1990. The origin of the Nova Canton
oriented drill core: evidence for rotation of the North Slope block. In: Tailleur, Trough reexamined. Transactions, American Geophysical Union 71, 1640.
I.L., Weimer, P. (Eds.), Alaskan North Slope Geology. Soc. Econ. Paleont. and Min., Joseph, D., Taylor, B., Shor, A.N., 1992. New sidescan sonar and gravity evidence that the
Pacific Section, Los Angeles, pp. 581–617. Nova-Canton Trough is a fracture zone. Geology 20, 435–438.
Hall, R., 2002. Cenozoic geological and plate tectonic evolution of Southeast Asia and Jourdan, F., FÈraud, G., Bertrand, H., Kampunzu, A.B., Tshoso, G., Watkeys, M.K., Le Gall,
the SW Pacific: computer-based reconstructions, models and animations. Journal B., 2005. Karoo large igneous province: Brevity, origin, and relation to mass extinc-
of Asian Earth Sciences 20, 353–431. tion questioned by new 40Ar/39Ar age data. Geology 33, 745.
Hall, R., Fuller, M., Ali, J.R., Anderson, C.D., 1995. The Philippine Sea plate: magnetism Karig, D.E., 1971. Origin and development of marginal basins in the Western Pacific.
and reconstructions. In: Taylor, B., Natland, J.H. (Eds.), A Synthesis of Western Journal of Geophysical Research 76, 2542.
Pacific Drilling Results, vol. 88. American Geophysical Union, pp. 371–404. Keppie, J.D., Dostal, J., 2001. Evaluation of the Baja controversy using paleomagnetic
Hamilton, W., 1979. Tectonics of the Indonesian Region. USGS Professional Paper 1078, and faunal data, plume magmatism, and piercing points. Tectonophysics 339,
345. 427–442.
Handschumacher, D.W., 1976. Post-Eocene Plate Tectonics of the Eastern Pacific. In: Kimbell, G., Ritchie, J., Johnson, H., Gatliff, R., 2005. Controls on the structure and evo-
G.H.S., et al. (Ed.), The Geophysics of the Pacific Ocean Basin and its Margin: A Vol- lution of the NE Atlantic margin revealed by regional potential field imaging and
ume in Honour of George P. Woollard. American Geophysical Union, Washington 3D modelling. Geological Society of London 6, 933.
D.C., pp. 177–202. Klitgord, K., Schouten, H., 1986. Plate kinematics of the central Atlantic. In: Vogt, P.R.,
Handschumacher, D., Sager, W., Hilde, T., Bracey, D., 1988. Pre-Cretaceous tectonic evo- Tucholke, B.E. (Eds.), The Western North Atlantic Region, DNAG, Volume M: The
lution of the Pacific plate and extension of the geomagnetic polarity reversal time Geology of North America. Geol. Soc. Am., Boulder, CO, United States, pp. 351–378.
scale with implications for the origin of the Jurassic 'Quiet Zone'. Tectonophysics Knott, S., Burchell, M., Jolley, E., Fraser, A., 1993. Mesozoic to Cenozoic Plate Recon-
155, 365–380. structions of the North Atlantic and Hydrocarbon Plays of the Atlantic margins,
Hankel, O., 1994. Early Permian to middle Jurassic rifting and sedimentation in East Volume 4, Geological Society of London, p. 953.
Africa and Madagascar. Geologische Rundschau 83, 703–710. König, M., Jokat, W., 2006. The Mesozoic breakup of the Weddell Sea. Journal of Geophysical
Hansen, V.L., 1990. Yukon-Tanana terrane: a partial acquittal. Geology 18, 365. Research 111, B12102.
Harland, W.B., Armstrong, R.L., Cox, A.V., Craig, I.E., Smith, A.G., Smith, D.G., 1990. A König, M., Jokat, W., 2010. Advanced insights into magmatism and volcanism of the
Geologic Time Scale. Camibridge University Press, Cambridge. Mozambique Ridge and Mozambique Basin in the view of new potential field
Hassanzadeh, J., Stockli, D.F., Horton, B.K., Axen, G.J., Stockli, L.D., Grove, M., Schmitt, data. Geophysical Journal International 180, 158–180.
A.K., Walker, J.D., 2008. U–Pb zircon geochronology of late Neoproterozoic–Early Kovacs, L.C., Morris, P., Brozena, J., Tikku, A., 2002. Seafloor spreading in the Weddell
Cambrian granitoids in Iran: implications for paleogeography, magmatism, and Sea from magnetic and gravity data. Tectonophysics 347, 43–64.
exhumation history of Iranian basement. Tectonophysics 451, 71–96. Kravchinsky, V.A., Cognè, J.P., Harbert, W.P., Kuzmin, M.I., 2002. Evolution of the Mongol–
Hawlader, H., 1990. Diagenesis and reservoir potential of volcanogenic sandstones— Okhotsk Ocean as constrained by new palaeomagnetic data from the Mongol-
Cretaceous of the Surat Basin, Australia. Sedimentary Geology 66, 181–195. Okhotsk suture zone, Siberia. Geophysical Journal International 148, 34–57.
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 267
Kuzmichev, A.B., 2009. Where does the South Anyui suture go in the New Siberian Mammerickx, J., Klitgord, K., 1982. Northern East Pacific rise: evolution from 25 my BP
islands and Laptev Sea?: implications for the Amerasia basin origin. Tectonophysics to the present. Journal of Geophysical Research 87, 6751–6759.
463, 86–108. Mammerickx, J., Sharman, G.F., 1988. Tectonic evolution of the North Pacific during the
Labails, C., Olivet, J., Aslanian, D., Roest, W., 2010. An alternative early opening scenario Cretaceous quiet period. Journal of Geophysical Research 93, 3009–3024.
for the Central Atlantic Ocean. Earth and Planetary Science Letters. Marks, K.M., Stock, J.M., 2001. Evolution of the Malvinas Plate south of Africa. Marine
Labrecque, J.L., Hayes, D.E., 1979. Seafloor spreading history of the Agulhas Basin. Earth Geophysical Researches 22, 289–302.
and Planetary Science Letters 45, 411–428. Marks, K.M., Tikku, A.A., 2001. Cretaceous reconstructions of East Antarctica, Africa and
LaBrecque, J.L., Barker, P., 1981. The age of the Weddell Basin. Nature 290, 489–492. Madagascar. Earth and Planetary Science Letters 186, 479–495.
LaBrecque, J.L., Rabinowitz, P.D., 1977. Magnetic anomalies bordering the continental Marton, G., Buffler, R., 1999. Jurassic–early cretaceous tectono-paleogeographic evolution
margin of Argentina: Tulsa. Amer. Assoc. Petr. Geol. map series. of the southeastern gulf of Mexico basin. Sedimentary Basins of the World 4, 63–91.
Lancelot, Y., Larson, R.L., et al., 1990. Proc. ODP Initial Reports 129. Ocean Drilling Masson, D.G., 1984. Evolution of the Mascarene basin, western Indian Ocean, and the
Program, College Station, TX. significance of the Amirante arc. Marine Geophysical Researches 6, 365–382.
Lane, L.S., 1997. Canada basin, Arctic Ocean — evidence against a rotational origin Matthews, K.J., et al., 2011. Dynamic subsidence of Eastern Australia during the Creta-
[Review]. Tectonics 16, 363–387. ceous. Gondwana Research 19, 372–383.
Larson, R.L., 1995. The mid-Cretaceous superplume episode. Scientific American 272, 66–70. Maus, S., Sazonova, T., Hemant, K., Fairhead, J.D., Ravat, D., 2007. National geophysical
Larson, R.L., 1997. Superplumes and ridge interactions between Ontong Java and Manihiki data center candidate for the world digital magnetic anomaly map. Geochemistry,
plateaus and the Nova-Canton trough. Geology 25, 779–782. Geophysics, Geosystems 8, Q06017. doi:10.1029/2007GC001643.
Larson, R.L., Chase, C.G., 1972. Late Mesozoic evolution of the Western Pacific. Geological Maus, S., Barckhausen, U., Berkenbosch, H., Bournas, N., Brozena, J., Childers, V.,
Society of America Bulletin 83, 3627–3644. Dostaler, F., Fairhead, J., Finn, C., Von Frese, R., 2009. EMAG2: a 2-arc min resolution
Larson, R.L., Pitman, W.C.I., 1972. World-wide correlation of Mesozoic magnetic anom- Earth Magnetic Anomaly Grid compiled from satellite, airborne, and marine mag-
alies, and its implications. Geological Society of America Bulletin 83, 3645–3662. netic measurements. Geochemistry, Geophysics, Geosystems 10, Q08005.
Larson, R., Smith, S., Chase, C., 1972. Magnetic lineations of Early Cretaceous age in the Mayes, C.L., Lawver, L.A., Sandwell, D.T., 1990. Tectonic history and new isochron chart
western equatorial Pacific Ocean. Earth and Planetary Science Letters 15, 315–319. of the South Pacific. Journal of Geophysical Research 95, 8543–8567.
Larson, R.L., Pockalny, R.A., Viso, R.F., Erba, E., Abrams, L.J., Luyendyk, B.P., Stock, J.M., Mcelhinny, M.W., Embleton, B.J.J., Daly, L., Pozzi, J.P., 1976. Paleomagnetic evidence for
Clayton, R.W., 2002. Mid-Cretaceous tectonic evolution of the Tongareva triple the location of Madagascar in Gondwanaland. Geology 4, 455–457.
junction in the southwestern Pacific Basin. Geology 30, 67–70. McKenzie, D.P., Morgan, J., 1969. The evolution of triple junctions. Nature 224, 125–133.
Larson, R.L., 1976. Late Jurassic and Early Cretaceous evolution of the Western Central Menard, H.W., 1978. Fragmentation of the Farallon plate by pivoting subduction. Journal
Pacific Ocean. Journal of Geomagnetism and Geoelectricity 28, 219–236. of Geology 86, 99–110.
Larson, R.L., 1977. Early Cretaceous breakup of Gondwanaland off western Australia. Meschede, M., Frisch, W., 1998. A plate-tectonic model for the Mesozoic and Early
Geology 5, 57–60. Cenozoic history of the Caribbean plate. Tectonophysics 296, 269–291.
Larter, R.D., Cunningham, A.P., Barker, P.F., Gohl, K., Nitsche, F.O., 2002. Tectonic evolu- Meschede, M., Barckhausen, U., Worm, H.U., 1998a. Extinct spreading on the Cocos
tion of the Pacific margin of Antarctica 1. Late Cretaceous tectonic reconstructions. Ridge. Terra Nova 10, 211–216.
Journal of Geophysical Research-Solid Earth 107, 2345. Meschede, M., Barckhausen, U., Worm, H.U., 1998b. Extinct spreading on the Cocos
Lawver, L.A., Müller, R.D., 1994. Iceland hotspot track. Geology 22, 311. Ridge: Terra Nova. The European Journal of Geosciences 10, 211–216.
Lawver, L.A., Müller, R.D., Srivastava, S.P., Roest, W., 1990. Opening of the Arctic Ocean. Metcalfe, I., 1996. Gondwanaland dispersion, Asian accretion and evolution of Eastern
In: Bleil, U., Thiede, J. (Eds.), Geological History of the Polar Oceans: Arctic Versus Tethys. Australian Journal of Earth Sciences 43, 605–623.
Antarctic, pp. 29–62. Metcalfe, I., 1999. Gondwana dispersion and Asian accretion: an overview. In: Metcalfe, I.
Lawver, L.A., Scotese, C.R., 1987. A revised reconstruction of Gondwanaland. AGU Geo- (Ed.), Gondwana Dispersion and Asian Accretion. A.A. Balkema, Rotterdam, pp. 9–29.
physical Monograph 40, 17–23. Mihalynuk, M.G., Nelson, J.A., Diakow, L.J., 1994. Cache Creek terrane entrapment:
Lawver, L., Grantz, A., Gahagan, L., 2002. Plate kinematic evolution of the present Arctic oroclinal paradox within the Canadian Cordillera. Tectonics 13, 575–595.
region since the Ordovician. Geological Society of America Special Papers 360, 333. Molnar, P., Atwater, T., Mammerickx, J., Smith, S.M., 1975. Magnetic anomalies, ba-
Lee, T.-Y., Lawver, L.A., 1994. Cenozoic plate reconstruction of the South China Sea thymetry and the tectonic evolution of the South Pacific since the Late Cretaceous.
region. Tectonophysics 235, 149–180. Geophysical Journal of the Royal Astronomical Society 40, 383–420.
Lee, T.-Y., Lawver, L.A., 1995. Cenozoic plate reconstruction of Southeast Asia. Tectonophysics Moran, K., Backman, J., Brinkhuis, H., Clemens, S.C., Cronin, T., Dickens, G.R., Eynaud, F.,
251, 85–138. Gattacceca, J., Jakobsson, M., Jordan, R.W., 2006. The Cenozoic palaeoenvironment
Lemaux, J., Gordon, R.G., Royer, J.Y., 2002. Location of the Nubia–Somalia boundary of the Arctic Ocean. Nature 441, 601–605.
along the Southwest Indian Ridge. Geology 30, 339–342. Morgan, W.J., 1972. Plate motions and deep mantle convection. GSA Memoir 132, 7–22.
Lemoine, M., 1983. Rifting and early drifting: Mesozoic central Atlantic and Ligurian Mortensen, J., 1992. Pre-mid-Mesozoic tectonic evolution of the Yukon-Tanana terrane,
Tethys. Initial Reports of the Deep Sea Drilling Project 76, 885ñ895. Yukon and Alaska. Tectonics 11, 836–853.
Lewis, J.C., Byrne, T.B., Pasteris, J.D., London, D., Morgan, G.B., 2000. Early Tertiary fluid Moulin, M., Aslanian, D., Unternehr, P., 2010. A new starting point for the South and
flow and pressure-temperature conditions in the Shimanto accretionary complex Equatorial Atlantic Ocean. Earth-Science Reviews 98, 1–37.
of south-west Japan: constraints from fluid inclusions. Journal of Metamorphic Ge- Müller, R.D., Roest, W.R., 1992. Fracture zones in the North Atlantic from combined
ology 18, 319–333. Geosat and Seasat data. Journal of Geophysical Research 97, 3337–3350.
Lithgow-Bertelloni, C., Richards, M., 1998. The Dynamics of Cenozoic and Mesozoic Müller, R.D., Roest, W.R., Royer, J.-Y., Gahagan, L.M., Sclater, J.G., 1997. Digital isochrons
Plate Motions. Reviews of Geophysics 36, 27–78. of the world's ocean floor. Journal of Geophysical Research 102, 3211–3214.
Liu, L., Gurnis, M., Seton, M., Saleeby, J., M¸Ller, R., Jackson, J., 2010. The role of oceanic Müller, R.D., Mihut, D., Baldwin, S., 1998a. A new kinematic model for the formation
plateau subduction in the Laramide orogeny. Nature Geoscience 3, 353–357. and evolution of the Northwest and West Australian margin. In: Purcell, P.G.,
Livermore, R., Hunter, R., 1996. Mesozoic seafloor spreading in the southern Weddell Purcell, R.R. (Eds.), The Sedimentary Basins of Western Australia, 2. Petroleum Explo-
Sea. Geological Society London Special Publications 108, 227. ration Society of Australia, Perth, pp. 55–72.
Lonsdale, P., 1988. Paleogene history of the Kula plate: offshore evidence and onshore Müller, R.D., Roest, W.R., Royer, J.-Y., 1998b. Asymmetric seafloor spreading expresses
implications. Geological Society of America Bulletin 100, 733. ridge–plume interactions. Nature 396, 455–459.
Lonsdale, P., 1997. An incomplete geologic history of the southwest Pacific basin. GSA Müller, R.D., Cande, S.C., Royer, J.-Y., Roest, W.R., Maschenkov, S., 1999. New constraints
Abstracts with Programs, vol. 29, p. 25. on the Late Cretaceous/Tertiary plate tectonic evolution of the Caribbean. In: Mann,
Lonsdale, P., 2005. Creation of the Cocos and Nazca plates by fission of the Farallon P. (Ed.), Caribbean Basins, Volume 4: Sedimentary Basins of the World. Elsevier,
plate. Tectonophysics 404, 237–264. Amsterdam, pp. 39–55.
Lonsdale, P., Klitgord, K.D., 1978. Structure and tectonic history of the eastern Panama Müller, R.D., Gaina, C., Tikku, A., Mihut, D., Cande, S., Stock, J.M., 2000. Mesozoic/Cenozoic
Basin. Geological Society of America Bulletin 89, 981–999. tectonic events around Australia. In: Richards, M., Gordon, R. (Eds.), The History
Lowrie, W., Kent, D., 2004. Geomagnetic polarity timescales and reversal frequency and Dynamics of Global Plate Motions, Volume 121: American Geophysical Union
regimes: timescales of the Paleomagnetic Field. Geophysical Monograph Series Monograph. American Geophysical Union, pp. 161–188.
145, 117ñ129. Müller, R., Goncharov, A., Kritski, A., 2005. Geophysical evaluation of the enigmatic
Ludwig, W.J., Kumar, N., Houtz, R.E., 1979. Profiler-Sonobuoy measurements in the Bedout basement high, offshore northwestern Australia. Earth and Planetary Science
South China Sea basin. Journal of Geophysical Research 84, 3505–3518. Letters 237, 264–284.
Madsen, J.K., Thorkelson, D.J., Friedman, R.M., Marshall, D.D., 2006. Cenozoic to Recent Müller, R.D., Sdrolias, M., Gaina, C., Roest, W.R., 2008a. Age, spreading rates and spreading
plate configurations in the Pacific Basin: Ridge subduction and slab window mag- asymmetry of the world's ocean crust. Geochemistry, Geophysics, Geosystems 9.
matism in western North America. Geosphere 2, 11–34. Müller, R.D., Sdrolias, M., Gaina, C., Steinberger, B., Heine, C., 2008b. Long-term sea-
Mahoney, J.J., Storey, M., Duncan, R.A., Spencer, K.J., Pringle, M., 1993. Geochemistry level fluctuations driven by ocean basin dynamics. Science 319, 1357–1362.
and geochronology of Leg 130 basement lavas: nature and origin of the Ontong Munschy, M., Antoine, C., Gachon, A., 1996. Tectonic evolution in the Tuamotu Islands
Java Plateau. In: Berger, W.H., Kroenke, L.W., Mayer, L.A. (Eds.), Proceedings of Region, Central Pacific Ocean. Comptes Rendus de l'Academie des Sciences Series
the Ocean Drilling Program, Scientific Results, vol. 130. Ocean Drilling Program, IIA Earth and Planetary Science 323, 941–948.
Texas A&M University, College Station, TX, pp. 3–22. Nakanishi, M., Tamaki, K., Kobayashi, K., 1992. A new Mesozoic isochron chart of the
Mahoney, J., Duncan, R., Tejada, M., Sager, W., Bralower, T., 2005. Jurassic–Cretaceous northwestern Pacific Ocean: paleomagnetic and tectonic implications. Geophysical
boundary age and mid-ocean-ridge-type mantle source for Shatsky Rise. Geology Research Letters 19, 693–696.
33, 185. Nakanishi, M., Winterer, E.L., 1998. Tectonic history of the Pacific–Farallon–Phoenix tri-
Malfait, B.T., Dinkelman, M.G., 1972. Circum-Caribbean Tectonic and Igneous Activity ple junction from late Jurassic to early Cretaceous: an abandoned Mesozoic spread-
and the Evolution of the Caribbean Plate. Geological Society of America Bulletin ing system in the central Pacific basin. Journal of Geophysical Research-Solid Earth
83, 251–272. 103, 12453–12468.
268 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
Nakanishi, M., Sager, W.W., Klaus, A., 1999. Magnetic lineations within Shatsky Rise, Roeser, H.A., Fritsch, J., Hinz, K., 1996. The development of the crust off Dronning Maud Land,
northwest Pacific Ocean: implications for hot spot-triple junction interaction and East Antarctica. In: Storey, B.C., King, E.C., Livermore, R.A. (Eds.), Weddell Sea Tectonics
oceanic plateau formation. Journal of Geophysical Research-Solid Earth 104, and Gondwana Break-upGeological Society, London, Special Publications 108, 243–264.
7539–7556. Roeser, H., Steiner, C., Schreckenberger, B., Block, M., 2002. Structural development of
Nelson, J.A., Mihalynuk, M., 1993. Cache Creek ocean: closure or enclosure? Geology 21, the Jurassic Magnetic Quiet Zone off Morocco and identification of Middle Jurassic
173. magnetic lineations. Journal of Geophysical Research 107, 2207.
Nelson, J.A.L., Colpron, M., Piercey, S.J., Dusel-Bacon, C., Murphy, D.C., Roots, C.F., 2006. Roest, W.R., Srivastava, S.P., 1989. Seafloor spreading in the Labrador Sea: a new recon-
Paleozoic tectonic and Metallogenetic Evolution of Pericratonic Terranes in Yukon, struction. Geology 17, 1000–1004.
Northern British Columbia and Eastern Alaska 1: Paleozoic Evolution and Metallogeny Roest, W., Srivastava, S., 1991. Kinematics of the plate boundaries between Eurasia, Iberia,
of Pericratonic Terranes at the Ancient Pacific Margin of North America, Canadian and and Africa in the North Atlantic from the Late Cretaceous to the present. Geology 19, 613.
Alaskan Cordillera, p. 323. Rosa, J.W.C., Molnar, P., 1988. Uncertainties in reconstructions of the Pacific, Farallon,
Nokleberg, W., Parfenov, L., Monger, J., Norton, I., Chan Uk, A., Stone, D., Scotese, C., Vancouver, and Kula plates and constraints on the rigidity of the Pacific and
Scholl, D., Fujita, K., 2001. Phanerozoic Tectonic Evolution of the Circum-North Farallon (and Vancouver) plates between 72 and 35 Ma. Journal of Geophysical
Pacific. US Dept. of the Interior, US Geological Survey. Research 93, 2997–3008.
Norton, I.O., Sclater, J.G., 1979. A model for the evolution of the Indian Ocean and the Rosenbaum, G., Lister, G.S., Duboz, C., 2002. Relative motions of Africa, Iberia and
breakup of Gondwanaland. Journal of Geophysical Research 84, 6803–6830. Europe during Alpine orogeny. Tectonophysics 359, 117–129.
Norton, I., 2007. Speculations on Cretaceous tectonic history of the northwest Pacific Rosenbaum, G., Giles, D., Saxon, M., Betts, P.G., Weinberg, R.F., Duboz, C., 2005. Subduc-
and a tectonic origin for the Hawaii hotspot. Geological Society of America Special tion of the Nazca Ridge and the Inca Plateau: insights into the formation of ore de-
Papers 430, 451. posits in Peru. Earth and Planetary Science Letters 239, 18–32.
Nunns, A.G., 1983. Plate tectonic evolution of the Greenland-Scotland Ridge and surrounding Rosencrantz, E., Ross, M.I., Sclater, J.G., 1988. Age and spreading history of the Cayman
regions. In: Bott, M.H.P., Saxov, S., Talwani, M., Thiede, J. (Eds.), Structure and Develop- Trough as determined from depth, heat flow and magnetic anomalies. Journal of
ment of the Greenland–Scotland Ridge. Plenum, New York, New York, pp. 11–30. Geophysical Research 93, 2141–2157.
Nürnberg, D., Müller, R.D., 1991. The tectonic evolution of the South Atlantic from Late Rosendahl, B.R., Moberly, R., Halunen, A.J., Rose, J.C., Kroenke, L.W., 1975. Geological
Jurassic to present. Tectonophysics 191, 27–53. and geophysical studies of the Canton Trough region. Journal of Geophysical
O'Neill, C., Müller, R.D., Steinberger, B., 2005. On the uncertainties in hotspot recon- Research 80, 2565–2574.
structions, and the significance of moving hotspot reference frames. Geochemistry, Ross, M., Scotese, C., 1988. A hierarchical tectonic model of the Gulf of Mexico and
Geophysics, Geosystems 6. doi:10.1029/2004GC000784. Caribbean region. Tectonophysics 155, 139–168.
Pardo-Casas, F., Molnar, P., 1987. Relative motion of the Nazca (Farallon) and South Rowley, D.B., Lottes, A.L., 1988. Plate-kinematic reconstructions of the North Atlantic
American plates since Late Cretaceous time. Tectonics 6, 215–232. and Arctic: Late Jurassic to present. Tectonophysics 155, 73–120.
Patriat, P., Ségoufin, J., 1988. Reconstruction of the central Indian Ocean. Tectonophysics Royer, J.-Y., Chang, T., 1991. Evidence for relative motions between the Indian and
155, 211–234. Australian plates during the last 20 m.y. from plate tectonic reconstructions: impli-
Peron-Pinvidic, G., Manatschal, G., Minshull, T., Sawyer, D., 2007. Tectonosedimentary cations for the deformation of the Indo-Australian plate. Journal of Geophysical
evolution of the deep Iberia-Newfoundland margins: evidence for a complex Research 96, 11779–11802.
breakup history. Tectonics 26. Royer, J.-Y., Coffin, M.F., 1992. Jurassic to Eocene plate tectonic reconstructions in the
Pindell, J.L., B., S.F., 1987. Geological evolution of the Caribbean Region: A plate tectonic Kerguelen Plateau region. In: S.W., Wise, J., Julson, A.P., Schlich, R., Thomas, E.
perspective. The Caribbean Region. : The Geology of North America, vol. 1. Geol. (Eds.), Proceedings of the Ocean Drilling Program, Scientific Results, vol. 120.
Soc. America, Boulder, Colorado. Ocean Drilling Program, Texas A&M University, College Station, TX, pp. 917–930.
Pindell, J., Kennan, L., 2009. Tectonic evolution of the Gulf of Mexico, Caribbean and Royer, J.-Y., Patriat, P., Bergh, H., Scotese, C.R., 1988. Evolution of the Southwest Indian
northern South America in the mantle reference frame: an update. Geological Society Ridge from the Late Cretaceous (anomaly 34) to the middle Eocene (anomaly 20).
London Special Publications 328, 1. Tectonophysics 155, 235–260.
Pindell, J.L., Cande, S., Pitman III, W.C., Rowley, D.B., Dewey, J.F., Labrecque, J., Haxby, W., Royer, J.Y., Gordon, R.G., 1997. The motion and boundary between the Capricorn and
1988. A plate-kinematic framework for models of Caribbean evolution. Tectonophysics Australian plates. Science 277, 1268–1274.
155, 121–138. Ru, K., Pigott, J.D., 1986. Episodic rifting and subsidence in the South China Sea. American
Pindell, J., Kennan, L.S., K-P, M., 2006. Foundations of Gulf of Mexico and Caribbean Association of Petroleum Geologists Bulletin 70, 1136–1155.
evolution: eight controversies resolved. Geologica Acta 4, 303ñ341. Russell, S.M., Whitmarsh, R.B., 2003. Magmatism at the west Iberia non-volcanic rifted
Pockalny, R.A., Larson, R.L., Viso, R.F., Abrams, L.J., 2002. Bathymetry and gravity data continental margin: evidence from analyses of magnetic anomalies. Geophysical
across a mid-Cretaceous triple junction trace in the southwest Pacific basin. Geophysical Journal International 154, 706–730.
Research Letters 29, 1007. Sager, W.S., Pringle, M.S., 1987. Tectonic evolution of the Central Pacific during the Cre-
Pogue, K.R., Dipietro, J.A., Khan, S.R., Hughes, S.S., Dilles, J.H., Lawrence, R.D., 1992. Late taceous Quiet Period. EOS Transactions American Geophysical Union 68, 44–45.
Paleozoic rifting in northern Pakistan. Tectonics 11, 871–883. Sager, W.W., Fullerton, L.G., Buffler, R.T., Handschumacher, D.W., 1992. Argo Abyssal
Powell, C.M., Roots, S.R., Veevers, J.J., 1988. Pre-breakup continental extension in East Plain magnetic lineations revisited: implications for the onset of seafloor spreading
Gondwanaland and the early opening of the eastern Indian Ocean. Tectonophysics and tectonic evolution of the eastern Indian Ocean. Proceedings of the Ocean Dril-
155, 261–283. ling Program, Scientific Results 123 (College Station, Texas, ODP).
Rabinowitz, P.D., 1971. Gravity anomalies across the East African continental margin. Sager, W., 2005. What built Shatsky Rise, a mantle plume or ridge tectonics? Geological
Journal of Geophysical Research 76, 7107–7117. Society of America Special Papers 388, 721.
Rabinowitz, P.D., Coffin, M.F., Falvey, D., 1983. The separation of Madagascar and Africa. Sager, W., Handschumacher, D., Hilde, T., Bracey, D., 1988. Tectonic evolution of the
Science 220, 67–69. northern Pacific plate and Pacific–Farallon Izanagi triple junction in the Late Jurassic
Ramana, M.V., Nair, R.R., Sarma, K.V.L.N.S., Ramprasad, T., Krishna, K.S., Subrahmanyam, and Early Cretaceous (M21–M10). Tectonophysics 155, 345–364.
V., D'Cruz, M., Subrahmanyam, C., Paul, J., Subrahmanyam, A.S., Sekhar, D.V.C., Sager, W.W., Weiss, C.J., Tivey, M.A., Johnson, H.P., 1998. Geomagnetic polarity reversal
1994. Mesozoic anomalies in the Bay of Bengal. Earth and Planetary Science Letters model of deep-tow profiles from the Pacific Jurassic Quiet Zone. Journal of Geophysical
121, 469–475. Research-Solid Earth 103, 5269–5286.
Ramana, M.V., Krishna, K.S., Ramprasad, T., Desa, M., Subrahmanyam, V., Sarma, K., 2001. Sager, W.W., Kim, J., Klaus, A., Nakanishi, M., Khankishieva, L.M., 1999. Bathymetry of
Structure and tectonic evolution of the northeastern Indian Ocean. Indian Ocean - a Shatsky Rise, northwest Pacific Ocean: implications for ocean plateau development
Perspective, vols. 1 and 2. A A Balkema Publishers, Rotterdam, pp. 731–816. at a triple junction. Journal of Geophysical Research-Solid Earth 104, 7557–7576.
Rangin, C., Jolivet, L., Pubellier, M., 1990. A simple model for the tectonic evolution of Sagong, H., Kwon, S., Ree, J., 2005. Mesozoic episodic magmatism in South Korea and its
southeast Asia and Indonesia region for the past 43 my. Bulletin de la Société tectonic implication. Tectonics 24.
Géologique de France 8, 889–905. Sahabi, M., Aslanian, D., Olivet, J., 2004. A new starting point for the history of the cen-
Rea, D.K., Dixon, J.M., 1983. Late Cretaceous and Paleogene tectonic evolution of the tral Atlantic. Comptes Rendus Geoscience 336, 1041–1052.
North Pacific Ocean. Earth and Planetary Science Letters 65, 145–166. Sakaguchi, A., 1996. High paleogeothermal gradient with ridge subduction beneath the
Reeves, C., 2000. The geophysical mapping of Mesozoic dyke swarms in southern Africa Cretaceous Shimanto accretionary prism, southwest Japan. Geology 24, 795.
and their origin in the disruption of Gondwana. Journal of African Earth Sciences Sandwell, D.T., Smith, W.H.F., 1997. Marine gravity anomaly from Geosat and ERS-1
30, 499–513. satellite altimetry. Journal of Geophysical Research-Solid Earth 102,
Reeves, C., De Wit, M., 2000. Making ends meet in Gondwana: retracing the transforms of 10039–10054.
the Indian Ocean and reconnecting continental shear zones. Terra Nova 12, 272–280. Sandwell, D.T., Smith, W.H.F., 2005. Retracking ERS-1 altimeter waveforms for optimal
Richards, M.A., Jones, D.L., Duncan, R.A., Depaolo, D.J., 1991. A mantle plume initiation gravity field recovery. Geophysical Journal International 163, 79–89.
model for the formation of Wrangellia and other oceanic flood basalt plateaus. Sandwell, D., Smith, W., 2009. Global marine gravity from retracked Geosat and ERS-1
Science 254, 263–267. altimetry: ridge segmentation versus spreading rate. Journal of Geophysical Re-
Riisager, P., Hall, S., Antretter, M., Zhao, X., 2003. Paleomagnetic paleolatitude of Early search 114.
Cretaceous Ontong Java Plateau basalts: implications for Pacific apparent and Sashida, K., Igo, H., Hisafa, K.I., Nakornsri, N., Ampornmaha, A., 1993. Occurrence of Pa-
true polar wander. Earth and Planetary Science Letters 208, 235–252. leozoic and Early Mesozoic radiolaria in Thailand (preliminary report). Journal of
Robb, M.S., Taylor, B., Goodliffe, A.M., 2005. Re examination of the magnetic lineations Southeast Asian Earth Sciences 8, 97–108.
of the Gascoyne and Cuvier Abyssal Plains, off NW Australia. Geophysical Journal Sayers, J., Symonds, P.A., Direen, N.G., Bernadel, G., 2001. Nature of the continent–ocean
International 163 (1), 42–55. transition on the non-volcanic rifted margin in the central Great Australian Bight.
Robertson, A.H.F., 2000. Mesozoic–Tertiary tectonic–sedimentary evolution of a south In: Wilson, R.C.L., Whitmarsh, R.B., Taylor, B., Froitzheim, N. (Eds.), Non-volcanic
Tethyan oceanic basin and its margins in southern Turkey. Geological Society, rifting of continental margins: a comparison of evidence from land and sea, Vol-
London, Special Publications 173, 97–138. ume Special Publication, Geological Society of London, pp. 51–76.
M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270 269
Schaefer, B.F., Turner, S., Parkinson, I., Rogers, N., Hawkesworth, C., 2002. Evidence for Speranza, F., Villa, I., Sagnotti, L., Florindo, F., Cosentino, D., Cipollari, P., Mattei, M.,
recycled Archaean oceanic mantle lithosphere in the Azores plume. Nature 420, 2002. Age of the Corsica–Sardinia rotation and Liguro–Provençal basin spreading:
304–307. new paleomagnetic and Ar/Ar evidence. Tectonophysics 347, 231–251.
Schettino, A., Scotese, C.R., 2005. Apparent polar wander paths for the major continents Srivastava, S.P., 1985. Evolution of the Eurasian Basin and its implication to the motion
(200 Ma to the present day): a palaeomagnetic reference frame for global plate of Greenland along Nares Strait. Tectonophysics 114, 29–53.
tectonic reconstructions. Geophysical Journal International 163, 727–759. Srivastava, S.P., Roest, W.R., 1989, Seafloor spreading history II–IV, in East Coast Basin Atlas
Schettino, A., Turco, E., 2009. Breakup of Pangaea and plate kinematics of the central Series: Labrador Sea, J.S. Bell (co-ordinator). Atlantic Geoscience Centre, Geologic
Atlantic and Atlas regions. Geophysical Journal International 178, 1078–1097. Survey of Canada, Map sheets L17-2–L17-6.
Scotese, C.R., 1991. Jurassic and cretaceous plate tectonic reconstructions. Palaeogeography, Srivastava, S.P., Roest, W.R., 1996. Porcupine plate hypothesis — comment. Marine
Palaeoclimatology, Palaeoecology 87, 493–501. Geophysical Researches 18, 589–593.
Scotese, C.R., Gahagan, L.M., Larson, R.L., 1988. Plate tectonic reconstructions of the Srivastava, S.P., Tapscott, C.R., 1986. Plate kinematics of the North Atlantic. In: Vogt,
Cretaceous and Cenozoic ocean basins. Tectonophysics 155, 27–48. R.R., Tucholke, B.E. (Eds.), The Western North Atlantic Region, Volume M: The Ge-
Scrutton, R.A., Heptonstall, W.B., Peacock, J.H., 1981. Constraints on the motion of Mad- ology of North America. Geol. Soc. Am., pp. 379–405.
agascar with respect to Africa. Marine Geology 43, 1–20. Srivastava, S., Schouten, H., Roest, W., Klitgord, K., Kovacs, L., Verhoef, J., Macnab, R., 1990.
Sdrolias, M., Müller, R.D., 2006. The controls on back-arc basin formation. Geochemistry, Iberian plate kinematics: a jumping plate boundary between Eurasia and Africa.
Geophysics, Geosystems 7, Q04016. doi:10.1029/2005GC001090. Srivastava, S.P., Sibuet, J.C., Cande, S., Roest, W.R., Reid, I.D., 2000. Magnetic evidence for
Sdrolias, M., Müller, R.D., Gaina, C., 2003a. Tectonic evolution of the SW Pacific using slow seafloor spreading during the formation of the Newfoundland and Iberian
constraints from back-arc basins. In: Hillis, R., Müller, R.D. (Eds.), Evolution and margins. Earth and Planetary Science Letters 182, 61–76.
Dynamics of the Australian Plate, Geological Society of Australia Special Publication Stadler, G., Gurnis, M., Burstedde, C., Wilcox, L., Alisic, L., Ghattas, O., 2010. The dynamics of
22 and Geological Society of America Special Paper. plate tectonics and mantle flow: From local to global scales. Science 329, 1033.
Sdrolias, M., Müller, R.D., Gaina, C., 2003b. Tectonic evolution of the Southwest Pacific Stagg, H.M.J., Willcox, J.B., 1992. A case for Australia–Antarctica separation in the
using constraints from backarc basins. In: Hills, R.R., Müller, R.D. (Eds.), Evolution Neocomian (c.a. 125 Ma). Tectonophysics 210, 21–32.
and Dynamics of the Australian Plate. Geological Society of Australia Special Publi- Stamatakos, J.A., Trop, J.M., Ridgway, K.D., 2001. Late Cretaceous paleogeography of
cation, vol. 22, pp. 343–359. Wrangellia: paleomagnetism of the MacColl Ridge Formation, southern Alaska,
Sdrolias, M., Müller, R.D., Mauffret, A., Bernardel, G., 2004. Enigmatic formation of the revisited. Geology 29, 947–950.
Norfolk Basin, SW Pacific: a plume influence on back-arc extension. Geochemistry, Stampfli, G.M., Borel, G.D., 2002. A plate tectonic model for the Paleozoic and Mesozoic
Geophysics, Geosystems 5, Q06005. constrained by dynamic plate boundaries and restored synthetic oceanic iso-
Searle, R., Bird, R., Rusby, R., Naar, D., 1993a. The development of two oceanic micro- chrons. Earth and Planetary Science Letters 196, 17–33.
plates: Easter and Juan Fernandez microplates, East Pacific Rise. Journal of the Stampfli, G.M., 2000. Tethyan Oceans. In: Bozkurt, E., Winchester, J.A., Piper, J.D.A.
Geological Society 150, 965. (Eds.), Tectonics and Magmatism in Turkey and the Surroundings AreaGeological
Searle, R., Holcomb, R., Wilson, J., Holmes, M., Whittington, R., Kappel, E., Mcgregor, B., Society, London, Special Publications 173, 1–23.
Shor, A., 1993b. The Molokai fracture zone near Hawaii, and the Late Cretaceous Steinberger, B., Torsvik, T., 2008. Absolute plate motions and true polar wander in the
change in Pacific/Farallon spreading direction. The Mesozoic Pacific: Geology, Tectonics, absence of hotspot tracks. Nature 452, 620–623.
and Volcanism: A Volume in Memory of Sy Schlanger, p. 155. Stock, J., Molnar, P., 1987. Revised history of early Tertiary plate motion in the south-
Segoufin, J., Patriat, P., 1980. Existence d'anomalies mesozoîques dans le bassin de Mo- west Pacific. Nature 325, 495–499.
zambique. Comptes Rendus de l'Academie des Sciences 287, 109–112. Stock, J., Molnar, P., 1988. Uncertainties and implications of the Late Cretaceous and
Sengor, A., 1987. Tectonics of the Tethysides: orogenic collage development in a colli- Tertiary position of North America relative to the Farallon, Kula, and Pacific plates.
sional setting. Annual Review of Earth and Planetary Sciences 15, 213–244. Tectonics 7, 1339–1384.
Sengör, A.M.C., Cin, A., Rowley, D.B., S-Y, N., 1993. Space-time patterns of magmatism Stock, J., Luyendyk, B., Clayton, R., Party, S.S., 1998. Tectonics and structure of the Manihiki
along the Tethysides: A Preliminary Study. Journal of Geology 101, 51–84. Plateau, western Pacific Ocean. EOS Transactions of the American Geophysical Union 79.
Seton, M., Müller, R., 2008. Reconstructing the junction between Panthalassa and Tethys Stocklin, J., 1974. Possible ancient continental margins in Iran. The Geology of Conti-
since the Early Cretaceous. Eastern Australasian Basins III, p. 263ñ266. nental Margins, pp. 873–887.
Seton, M., Gaina, C., Müller, R.D., Heine, C., 2009. Mid-Cretaceous seafloor spreading Storey, M., Mahoney, J.J., Saunders, A.D., Duncan, R.A., Kelley, S.P., Coffin, M.F., 1995.
pulse: fact or fiction? Geology 37, 687. Timing of hot spot-related volcanism and the breakup of Madagascar and India.
Seton, M., Whittaker, J., Flament, N., Müller, R., Gurnis, M., in preparation. A case for a Science 267, 852–855.
top-down plate reorganization at 50 Ma. Nature Geosciences, v. Storey, B.C., Leat, P.T., Ferris, J.K., 2001. The location of mantleplume centers during the
Shaw, P., Cande, S.C., 1990. High-resolution inversion for South Atlantic plate kinematics initial stages of Gondwana breakup. Mantle Plumes: Their Identification Through
using joint altimeter and magnetic anomaly data. Journal of Geophysical Research Time, pp. 71–80.
95, 2625–2644. Straub, S., Goldstein, S., Class, C., Schmidt, A., 2009. Mid-ocean-ridge basalt of Indian
Shephard, G.E., Bunge, H.P., Schuberth, B.S.A., Müller, R.D., Talsma, A.S., Moder, C., type in the northwest Pacific Ocean basin. Nature Geoscience 2, 286–289.
Landgrebe, T.C.W., 2012. Testing absolute plate reference frames and the implica- Sun, W., Ding, X., Hu, Y.-H., Li, X.-H., 2007. The golden transformation of the Cretcaous plate
tions for the generation of geodynamic mantle heterogeneity structure. Earth subduction in the west Pacific. Earth and Planetary Science Letters 262, 533–542.
and Planetary Science Letters 317–318, 204–217. Sutherland, R., Hollis, C., 2001. Cretaceous demise of the Moa Plate and strike-slip mo-
Shi, G., Archbold, N., 1998. Permian marine biogeography of SE Asia. Biogeography and tion at the Gondwana margin. Geology 29, 279–282.
Geological Evolution of SE Asia. Backhuys, Leiden, p. 57ñ72. Symonds, P.A., Colwell, J.B., Struckmeyer, H.I.M., Willcox, J.B., Hill, P.J., 1996. Mesozoic
Sibuet, J., Srivastava, S., Spakman, W., 2004. Pyrenean orogeny and plate kinematics. rift basin development off eastern Australia. Mesozoic Geology of the Eastern
Journal of Geophysical Research 109, B08104. Australia Plate, vol. 43. Geological Society of Australia, Brisbane, pp. 528–542.
Sibuet, J., Srivastava, S., Manatschal, G., 2007. Exhumed mantle-forming transitional Talwani, M., Eldholm, O., 1977. Evolution of the Norwegian–Greenland Sea. Geological
crust in the Newfoundland–Iberia rift and associated magnetic anomalies. Journal Society of America Bulletin 88, 969–999.
of Geophysical Research 112, B06105. Tamaki, K., Larson, R.L., 1988. The Mesozoic tectonic history of the Magellan microplate
Silva, E.A., Miranda, J., Luis, J., Galdeano, A., 2000. Correlation between the Palaeozoic in the western Central Pacific. Journal of Geophysical Research 93, 2857–2874.
structures from West Iberian and Grand Banks margins using inversion of magnet- Tankard, A.J., Welsink, H.J., 1987. Extensional tectonics and stratigraphy of Hibernia Oil
ic anomalies. Tectonophysics 321, 57–71. Field, Grand Banks, Newfoundland. American Association of Petroleum Geologists
Simpson, E.S.W., Sclater, J.G., Parsons, B., Norton, I.O., Meinke, L., 1979. Mesozoic magnetic Bulletin 71, 1210–1232.
lineations in the Mozambique Basin. Earth and Planetary Science Letters 43, 260–264. Taylor, B., 2006. The single largest oceanic plateau: Ontong Java-Manihiki-Hikurangi.
Singh, S.C., Carton, H., Chauhan, A.S., Androvandi, S., Davaille, A., Dyment, J., Cannat, M., Earth and Planetary Science Letters 241, 372–380.
Hananto, N.D., 2010. Extremely thin crust in the Indian Ocean possibly resulting Taylor, B., Karner, G., 1983. On the evolution of marginal basins. Reviews of Geophysics
from Plume–Ridge Interaction. Geophysical Journal International. 21, 1727–1741.
Sinton, C., Duncan, R., Storey, M., Lewis, J., Estrada, J., 1998. An oceanic flood basalt Taylor, P.T., Kovacs, L.C., Vogt, P.R., Johnson, G.L., 1981. Detailed aeromagnetic investi-
province within the Caribbean plate. Earth and Planetary Science Letters 155, gation of the Arctic Basin. II. Journal of Geophysical Research 86, 6323–6333.
221–235. Tebbens, S.F., Cande, S.C., 1997. Southeast Pacific tectonic evolution from early Oligocene
Sisson, V.B., Pavlis, T.L., 1993. Geologic consequences of plate reorganization: An exam- to Present. Journal of Geophysical Research-Solid Earth 102, 12061–12084.
ple from the Eocene southern Alaska fore arc. Geology 21, 913. Thorkelson, D.J., 1996. Subduction of diverging plates and the principles of slab win-
Skogseid, J., Planke, S., Faleide, J., Pedersen, T., Eldholm, O., Neverdal, F., 2000. NE Atlantic dow formation. Tectonophysics 255, 47–63.
continental rifting and volcanic margin formation. Dynamics of the Norwegian Tikku, A.A., Cande, S.C., 1999. The oldest magnetic anomalies in the Australian–Antarctic
Margin, vol. 167, p. 295ñ326. Basin: are they isochrons? Journal of Geophysical Research 104, 661–677.
Sleep, N.H., Toksoz, M.N., 1971. Evolution of marginal seas. Nature 233, 548–550. Tikku, A.A., Cande, S.C., 2000. On the fit of Broken Ridge and Kerguelen plateau. Earth
Small, C., Abbott, D., 1998. Subduction obstruction and the crack-up of the Pacific plate. and Planetary Science Letters 180, 117–132.
Geology 26, 795–798. Tivey, M., Sager, W., Lee, S., Tominaga, M., 2006. Origin of the Pacific Jurassic quiet zone.
Smith, A.G., Hallam, A., 1970. The fit of the southern continents. Nature 225, 139–144. Geology 34, 789.
Sokolov, S., Ye, B., Morozov, O., Shekhovtsov, V., Glotov, S., Ganelin, A., Todd, B., Reid, I., Keen, C., 1988. Crustal structure across the southwest Newfoundland
Kravchenko-Berezhnoy, I., 2002. South Anjui suture, northeast Arctic Russia. transform margin. Canadian Journal of Earth Sciences 25, 744–759.
Tectonic Evolution of the Bering Shelf-Chukchi Sea-Arctic Margin and Adjacent Tominaga, M., Sager, W.W., 2010. Revised Pacific M‐anomaly geomagnetic polarity
Landmasses, pp. 209–223. timescale. Geophysical Journal International 182, 203–232.
Speed, R.C., 1985. Cenozoic collision of the Lesser Antilles arc and continental South Tominaga, M., Sager, W.W., Tivey, M.A., Lee, S.M., 2008. Deep-tow magnetic anomaly
America and the origin of the El Pilar fault. Tectonics 4, 41–69. study of the Pacific Jurassic Quiet Zone and implications for the geomagnetic
270 M. Seton et al. / Earth-Science Reviews 113 (2012) 212–270
polarity reversal timescale and geomagnetic field behavior. Journal of Geophysical Viso, R.F., Larson, R.L., Pockalny, R.A., 2005. Tectonic evolution of the Pacific-Phoenix-
Research 113, B07110. Farallon triple junction in the South Pacific ocean. Earth and Planetary Science Let-
Tomurtogoo, O., Windley, B., Kroner, A., Badarch, G., Liu, D., 2005. Zircon age and occur- ters 233, 179–194.
rence of the Adaatsag ophiolite and Muron shear zone, central Mongolia: constraints Vry, J., Baker, J., Maas, R., Little, T., Grapes, R., Dixon, M., 2004. Zoned (Cretaceous and
on the evolution of the Mongol–Okhotsk ocean, suture and orogen. Journal of the Cenozoic) garnet and the timing of high grade metamorphism, Southern Alps:
Geological Society 162, 125. New Zealand. Journal of Metamorphic Geology 22, 137ñ157.
Torsvik, T., Steinberger, B., Cocks, L., Burke, K., 2008. Longitude: linking Earth's ancient Weissel, J.K., Anderson, R.N., Geller, C.A., 1980. Deformation of the Indo-Australian
surface to its deep interior. Earth and Planetary Science Letters 276, 273–282. Plate. Nature 287, 284–291.
Torsvik, T., Rousse, S., Labails, C., Smethurst, M., 2009. A new scheme for the opening of Wessel, P., Kroenke, L., 2008. Pacific absolute plate motion since 145 Ma: an assessment
the South Atlantic Ocean and the dissection of an Aptian salt basin. Geophysical of the fixed hot spot hypothesis. Journal of Geophysical Research 113, B06101.
Journal International 177, 1315–1333. Wessel, P., Harada, Y., Kroenke, L.W., 2006. Toward a self-consistent, high-resolution absolute
Totterdell, J., Blevin, J., Bradshaw, B., Colwell, J., Kennard, J., 2000. A new sequence frame- plate motion model for the Pacific. Geochemistry, Geophysics, Geosystems 7, Q03L12.
work for the Great Australian Bight: starting with a clean slate. The APPEA Journal 95. Whitechurch, H., Juteau, T., Montigny, R., 1984. Role of the Eastern Mediterranean
Trop, J.M., Ridgway, K.D., Manuszak, J.D., Layer, P., 2002. Mesozoic sedimentary-basin ophiolites (Turkey, Syria, Cyprus) in the history of the Neo-Tethys. Geological
development on the allochthonous Wrangellia composite terrane, Wrangell Society, London, Special Publications 17, 301–317.
Mountains basin, Alaska: a long-term record of terrane migration and arc construc- Whitmarsh, R.B., Miles, P.R., 1995. Models of the development of the West Iberia rifted
tion. Geological Society of America Bulletin 114, 693. continental margin at 40-degrees-30′N deduced from surface and deep-tow mag-
Tucholke, B., Whitmarsh, R., 2006. The Newfoundland–Iberia conjugate rifted margins. netic anomalies. Journal of Geophysical Research-Solid Earth 100, 3789–3806.
Principles of Phanerozoic Regional Geology. Whittaker, J., Müller, R.D., Leitchenkov, G., Stagg, H., Sdrolias, M., Gaina, C., Goncharov,
Tucholke, B., Sawyer, D., Sibuet, J., 2007. Breakup of the Newfoundland Iberia rift. A., 2007. Major Australian–Antarctica plate reorganization at Hawaiian-Emperor
Geological Society London Special Publications 282, 9. bend time. Science 318, 83–86.
Umpleby, D., 1979. Geology of the Labrador shelf. Geological Survey of Canada Williams, P., Johnston, C., Almond, R., Simamora, W., 1988. Late Cretaceous to early
Commission Gèologique du Canada. Tertiary structural elements of West Kalimantan. Tectonophysics 148, 279–297.
Unternehr, P., Curie, D., Olivet, J.L., Goslin, J., Benzarty, P., 1988. South Atlantic fits and Wilson, D.S., 1988. Tectonic history of the Juan de Fuca Ridge over the last 40 million
intraplate boundaries in Africa and South America. Tectonophysics 155, 169–179. years. Journal of Geophysical Research 33, 11,863–11,876.
Uyeda, S., Kanamori, H., 1979. Back-arc opening and the mode of subduction. Journal of Wilson, D.S., 1996. Fastest known spreading on the Miocene Cocos‐Pacific plate boundary.
Geophysical Research 84, 1049–1061. Geophysical Research Letters 23, 3003–3006.
Van Der Lelij, R., Spikings, R., Kerr, A., Kounov, A., Cosca, M., Chew, D., Villagomez, D., Wilson, D.S., Hey, R.N., Nishimura, C., 1984. Propagation as a mechanism of reorienta-
2010. Thermochronology and Tectonics of the Leeward Antilles: evolution of the tion of the Juan de Fuca Ridge. Journal of Geophysical Research 89, 9215–9225.
Southern Caribbean Plate Boundary Zone and accretion of the Bonaire Block. Winterer, E., 1976. Anomalies in the tectonic evolution of the Pacific. In: Sutton, G.,
Van Der Voo, R., Spakman, W., Bijwaard, H., 1999a. Mesozoic subducted slabs under Manghnani, M., Moberly, R. (Eds.), The Geophysics of the Pacific Ocean Basin and
Siberia. Nature 397, 246–249. its Margins. AGU, Washington D.C., pp. 269–278.
Van Der Voo, R., Spakman, W., Bijwaard, H., 1999b. Tethyan subducted slabs under Winterer, E., Lonsdale, P., Matthews, J., Rosendahl, B., 1974. Structure and acoustic stra-
India. Earth and Planetary Science Letters 171, 7–20. tigraphy of the Manihiki Plateau. Deep Sea Research 21, 793–814.
Veevers, J.J., 2000. Change of tectono-stratigraphic regime in the Australian plate Withjack, M.O., Schlische, R.W., Olsen, P.E., 1998. Diachronous rifting, drifting, and inversion
during the 99 Ma (mid-Cretaceous) and 43 Ma (mid-Eocene) swerves of the Pacific. on the passive margin of central eastern North America — an analog for other passive
Geology 28, 47–50. margins. AAPG Bulletin-American Association of Petroleum Geologists 82, 817–835.
Veevers, J., 2006. Updated Gondwana (Permian–Cretaceous) earth history of Australia. Woods, M.T., Davies, G.F., 1982. Late Cretaceous genesis of the Kula plate. Earth and
Gondwana Research 9, 231–260. Planetary Science Letters 58, 161–166.
Veevers, J.J., 1988. Morphotectonics of Australia's north-western margin - a review. In: Worthington, T.J., Hekinian, R., Stoffers, P., Kuhn, T., Hauff, F., 2006. Osbourn Trough:
Purcell, P.G., Purcell, R.R. (Eds.), The North West Shelf of Australia: Petroleum Ex- structure, geochemistry and implications of a mid-Cretaceous paleospreading
ploration Society of Australia Symposium: Perth, pp. 19–27. ridge in the South Pacific. Earth and Planetary Science Letters 245, 685–701.
Veevers, J.J., Tayton, J.W., Johnson, B.D., 1985. Prominent magnetic anomaly along Zorin, Y.A., 1999. Geodynamics of the western part of the Mongolia-Okhotsk collisional
the continent–ocean boundary between the northwestern margin of Australia belt, Trans-Baikal region (Russia) and Mongolia. Tectonophysics 306, 33–56.
(Exmouth and Scott plateaus) and the Argo abyssal plain. Earth and Planetary
Science Letters 72, 415–426.
Veevers, J.J., Stagg, H.M.J., Willcox, J.B., Davies, H.L., 1990. Pattern of slow seafloor spread-
ing (b 4 mm/year) from breakup (96 Ma) to A20 (44.5 Ma) off the southern margin of
Australia. BMR Journal of Australian Geology and Geophysics 11, 499–507.