Vous êtes sur la page 1sur 9

PUBLICATIONS

Geophysical Research Letters


RESEARCH LETTER Evidence for long-lived subduction of an ancient tectonic
10.1002/2015GL066237
plate beneath the southern Indian Ocean
Key Points: N. A. Simmons1, S. C. Myers1, G. Johannesson2, E. Matzel1, and S. P. Grand3
• Seismic tomography of the mantle has
uncovered a subducted slab beneath 1
Geophysical Monitoring Programs, Lawrence Livermore National Laboratory, Livermore, California, USA, 2Systems and
the southern Indian Ocean
• The subduction event is significant to Decision Sciences, Lawrence Livermore National Laboratory, Livermore, California, USA, 3Jackson School of Geosciences,
past plate tectonic history University of Texas at Austin, Austin, Texas, USA
• Subducted slabs can subsist in the
shallow mantle much longer than
previously realized Abstract Ancient subducted tectonic plates have been observed in past seismic images of the mantle
beneath North America and Eurasia, and it is likely that other ancient slab structures have remained largely
Supporting Information: hidden, particularly in the seismic-data-limited regions beneath the vast oceans in the Southern Hemisphere.
• Texts S1–S6, Figures S1–S17, and Here we present a new global tomographic image, which shows a slab-like structure beneath the southern
Table S1
Indian Ocean with coherency from the upper mantle to the core-mantle boundary region—a feature that has
Correspondence to:
never been identified. We postulate that the structure is an ancient tectonic plate that sank into the mantle
N. A. Simmons, along an extensive intraoceanic subduction zone that migrated southwestward across the ancient Tethys
simmons27@LLNL.gov Ocean in the Mesozoic Era. Slab material still trapped in the transition zone is positioned near the edge of East
Gondwana at 140 Ma suggesting that subduction terminated near the margin of the ancient continent prior to
Citation: breakup and subsequent dispersal of its subcontinents.
Simmons, N. A., S. C. Myers,
G. Johannesson, E. Matzel, and
S. P. Grand (2015), Evidence for long-lived
subduction of an ancient tectonic plate 1. Introduction
beneath the southern Indian Ocean,
Geophys. Res. Lett., 42, doi:10.1002/ Subduction of cold lithosphere into the mantle produces seismically fast anomalies that have been seen in
2015GL066237.
tomographic models for decades, providing us with essential knowledge about tectonic processes and
evolution of Earth. Resolution of tomographic images depends on both seismic ray geometry and data
Received 17 SEP 2015
Accepted 27 OCT 2015 quality. Subducted lithosphere along active convergent margins is commonly visible in seismic images
Accepted article online 4 NOV 2015 because seismic rays emanating from related earthquakes provide the ray geometry needed to resolve the
structures. Resolving ancient, aseismic slabs relies upon data generated by earthquakes occurring in remote
regions, often far from the actual slabs. There are a number of notable examples of imaged ancient slabs
beneath North America and Eurasia owing to the ample coverage provided by relatively dense instrumentation
[e.g., Grand et al., 1997; van der Hilst et al., 1997; van der Voo et al., 1999a, 1999b; Li and van der Hilst, 2010]. In
other parts of the world, however, ancient slab anomalies are much less recoverable, particularly in the
Southern Hemisphere where seismic station density is generally low.
For nuclear explosion monitoring research purposes, we have developed a number of seismic data proces-
sing and tomographic imaging procedures with the ultimate goal of generating high-resolution images of
global Earth structure that allow for accurate seismic event location in case of an underground nuclear test
occurring anywhere on the planet [Myers et al., 2007, 2009, 2011; Simmons et al., 2010, 2011, 2012]. The
techniques include a Bayesian global multiple-event relocation process (Bayesloc) that has been proven to
generate accurate seismic event locations, leading to internally consistent travel time data and more focused
images of Earth structure as a result [Myers et al., 2011; Simmons et al., 2012]. Three-dimensional (3-D) ray
tracing algorithms have also been developed to predict seismic wave travel times and model sensitivity
while also considering multipathing for crustal, regional, and triplicated seismic phases [Simmons et al.,
2011, 2012]. In order to account for highly variable data densities common to many seismic tomographic
problems, we developed a multiresolution imaging technique called Progressive Multi-level Tessellation
Inversion (PMTI) that exploits hierarchical spherical tessellation recursions [Simmons et al., 2011].
Combining these techniques, we have constructed a new global tomographic image of crust and mantle
P and S wave velocity structure via joint inversion of ~3 million high-quality data including crustal, regional,
and teleseismic travel times for a large suite of P and S wave body wave phases (P, Pg, Pn, Pb, pP, pwP, PcP, S,
Sn, SS, sS, ScS, SKS, and SKKS). The seismic phases were generated by 12,607 well-recorded events occurring
©2015. American Geophysical Union. since the 1960s and recorded at 7783 seismic stations around the world. See the supporting information for
All Rights Reserved. more details regarding the model development.

SIMMONS ET AL. ANCIENT SLAB BENEATH THE INDIAN OCEAN 1


Geophysical Research Letters 10.1002/2015GL066237

Figure 1. Cross sections of the mantle cutting through the fast (blue) slab structures beneath the southern Indian Ocean. (a) A cross section from the Kerguelen
Plateau to Indonesia (section 1) shows a large flat-lying fast anomaly in the transition zone (TZ) that is connected to fast material extending into the deepest
mantle. The slab tends to steepen from west to east (from section 1 to section 3) and is near vertical in the mantle southwest of Tasmania. (b) A cross section through
the well-known Farallon slab anomaly from the same global image is shown for comparison. (c) Comparison to a selection of recent global tomography models along
cross section 1 shown in Figure 1a. The green outline encompasses the fast anomaly in our new model for reference. See the supporting information for more
comparisons. S40RTS [Ritsema et al., 2011], Savani [Auer et al., 2014], SEMum [Lekić and Romanowicz, 2011], and Zhao2013 [Zhao et al., 2013].

2. The LLNL-G3D-JPS Global Seismic Tomography Model


Our new global Earth model (LLNL-G3D-JPS) reveals slab-like, seismically fast anomalies situated beneath the
southern Indian Ocean that were previously unrecognized (Figure 1a). The fast anomalies form coherent
dipping structures in the mantle that are clearest in a cross section of the mantle from the Kerguelen
Plateau in the southern Indian Ocean toward the island of Java, Indonesia (Figure 1a, cross section 1).

SIMMONS ET AL. ANCIENT SLAB BENEATH THE INDIAN OCEAN 2


Geophysical Research Letters 10.1002/2015GL066237

The structure along this section is visible at depths ranging from the upper mantle transition zone to the dee-
pest mantle, and it has a remarkably similar morphology and intensity when compared to the well-known
Farallon slab beneath North America [Grand et al., 1997; van der Hilst et al., 1997] (Figure 1b). Analogous to
the Farallon slab, we forward the hypothesis that the anomaly seen beneath the southern Indian Ocean is
also cold subducted lithosphere. The slab traverses the full vertical extent of the lower mantle along multiple
cross sections, sometimes extending upward into the transition zone (Figure 1a, sections 2 and 3). The dip of
the slab increases from west to east, becoming nearly vertical beneath the region southwest of Tasmania.
We mapped out the areal extent of the fast anomalies to obtain a sense of coherence and scale of the imaged
structures beneath the Indian Ocean (Figures 2a–2d). A pattern of fast material emerges within the transition
zone along a 7000 km track roughly parallel to the Southeast Indian Ridge (SEIR) (Figure 2a). The line of fast
structures is bookended by broadened anomalies near the Kerguelen Plateau and one south of Tasmania,
the largest of which (near Kerguelen) spans an area of ~2 million km2. The other large transition zone anomaly
south of Tasmania is offset to the north of the mid-ocean ridge but with a distinctive east-west orientation. We
see a set of three large areas with anomalously fast material in the shallow lower mantle (660–1171 km) that fall
along a similar trend as the transition zone anomalies (Figure 2b). The fast material becomes more laterally
coherent with depth, developing into a massive linear structure from the central Indian Ocean to southern
Australia in the deep mantle (Figures 2c and 2d). The deep trend correlates well with the long-wavelength
geoid low in the southern Indian Ocean, which has been attributed to Mesozoic subduction [Richards and
Engebretson, 1992; Spasojevic et al., 2010]. Although there are clear lateral gaps in the slab material in the
shallow lower mantle, a systematic analysis of the trends of fast material at all depths suggests that they are
related to the same tectonic process. There are several possible reasons for the gaps including inundation of
hot material from below in the distant past, and/or localized viscosity decreases that allowed the material to
sink into the deep mantle at different rates. High variability of slab sinking rates is supported by recent
observations of Farallon slab remnants at various depths [Schmandt and Lin, 2014]. Based on the proximity
of the newly identified shallowest slab material to the SEIR, we further refer to the new features collectively
as the Southeast Indian Slab (SEIS) anomaly.
Numerous resolution tests were performed to assess the robustness of significant features observed in the
new global model (Figures S3 and S4 in the supporting information). A subset of these resolution analyses
demonstrate that the SEIS anomaly seen from Kerguelen to Indonesia (Figure 1a, section 1) is resolvable
by the data, and the lower mantle portion of the slab is likely continuous (Figure S3). Significant vertical
smearing does occur in the upper mantle near Kerguelen, however. Specifically, the pervasive low-velocity
anomalies above 410 km depth could be due to image smearing from low velocities in the upper 100 km
of the mantle. Also, resolution is not sufficient to determine whether the fast anomaly in the upper mantle
transition zone spans from ~410 km to ~660 km depth, or whether the fast anomaly is limited to a smaller
depth range within the transition zone. Nonetheless, a significant volume of fast material must exist within
the transition zone in this region to recreate the image shown in Figure 1a. The same conclusion is drawn
for the remainder of fast material seen in the transition zone, including the region southwest of Tasmania.

3. Comparison to Past Results


There are significant differences between tomographic models beneath the southern Indian Ocean; however,
many of the fast features we observe are partially visible in other global models (Figures 1c and S5–S8). For
instance, the broad high-velocity anomalies seen in the deep mantle (>1900 km depth) from the central
Indian Ocean to southern Australia are commonly visible in global models. With careful examination, the fast
transition zone anomalies we observe from Kerguelen to Tasmania are also partially visible in a variety of
other global models, and even the full 7000 km high-velocity trend can be seen in some cases (Figure S5).
Although there have been hints of the SEIS anomaly in prior images, none of the previous models show
focused slab-like features with the same level of clarity. To understand why we see these anomalies more
clearly than previous studies, we performed an extensive set of tests to determine the impact of each of
our procedures on the resulting image (Figures S9–S12). These tests demonstrate that the single most impor-
tant factor in revealing these anomalies is the combination of high-quality P wave data and S wave data in a
joint (simultaneous) inversion (Figure S9). The first-arriving P wave data from events along the SEIR and other
Indian ridges recorded at numerous stations in East and Southeast Asia provide the heaviest sampling of the

SIMMONS ET AL. ANCIENT SLAB BENEATH THE INDIAN OCEAN 3


Geophysical Research Letters 10.1002/2015GL066237

Figure 2. The lateral position of the SEIS anomaly and our inferred subduction scenario. (a) The extent of slab material in the transition zone was determined
by mapping all shear wave perturbations of >0.25% for each model layer between 410 and 660 km depth. The color saturation indicates the amount of material
at each point (0% means there is no fast material at that location; 100% means the entire depth range is fast). SEIR = Southeast Indian Ridge; SEIS = Southeast
Indian Slab; AAD = Australian-Antarctic Discordance. (b) The same calculation as in Figure 2a but within the depth range of 660 to 1171 km. (c) 1171–1871 km.
(d) 1871–2371 km. Material related to the Southeast Indian Slab (SEIS) anomaly is outlined and labeled in each panel. We identify additional fast material with a
distinctly different trend that we interpret to be the Phoenix (PHX) and/or Pacific (PAC) plates that subducted westward along the margins of Australia and Antarctica
in the Mesozoic. The Kerguelen Plateau is shown in yellow for reference. (e) The position of fast materials with depth indicates a southwestward retreat of a subduction
zone that initiated beneath the present-day eastern Indian Ocean and southern Australia and terminated mostly south of the Southeast Indian Ridge (SEIR). (f) The
proposed subduction positions compared to the position of East Gondwana at 140 Ma according to Wu and Kravchinsky [2014]. With a 5° Euler rotation, the position of
East Gondwana conforms very well to the position of the inferred subduction zone at its later stages, demonstrating that subduction could have terminated along the
northern edge of the ancient continent. We postulate that the Jurassic-Cretaceous age Woyla arc terrane found along the outer edges of Sumatra could have been
constructed along this subduction zone but never collided with East Gondwana due to a kinematic reorganization of the Tethyan region at ~140 Ma that created
spreading near the continental margin [e.g., Seton et al., 2012].

SIMMONS ET AL. ANCIENT SLAB BENEATH THE INDIAN OCEAN 4


Geophysical Research Letters 10.1002/2015GL066237

SEIS anomaly (Figures S13–15). Direct S wave data from the ridge events are also important, but the S wave
multiples (SS) are equally important because of their unique sampling characteristics, particularly in this
remote region (Figure S16). Other phases are of secondary importance; however, they each provide unique
sampling geometries that contribute to the overall image. In addition, multiple-event location process
(Bayesloc), 3-D ray tracing, and multiresolution inversion process (PMTI) each contribute by improving the
clarity and focus of the slab-like anomalies.
Past global studies have jointly considered P wave and S wave travel time data [e.g., Masters et al., 2000;
Antolik et al., 2003; Kennett and Gorbatov, 2004; Houser et al., 2008]; however, there are a number of technical
differences from the current study including parameterizations (resolution, splines versus cells versus nodes)
and unique data sets. Our waveform correlation-based S wave data set (from Grand [2002] and additional
measurements) is unique and not used in most studies. In addition, while most of our unprocessed P wave
data set is available in public databases, the amount of data we use tends to be greater than many studies.
In addition, the set of P wave events chosen and processing steps (outlier removal etc.) make this a unique
data set as well [see Simmons et al., 2012]. Although the data differences are very important, one of the most
significant differences is the inversion itself where we attempt to find a result that is most consistent with
expected thermal behaviors of Vp and Vs by using scaling relationships in the inversion (RP/S) that are within
the range of mineral physics estimates for thermally varying mantle material. The scaling constraint couples
the data and increases the resolving power of the complimentary data sets relative to, for example, uncon-
strained inversions of Vp and Vs individually. Most joint studies do not attempt to enforce such a relationship
and thus do not drive the solution toward the basic thermal expectation. By design, our final models of Vp and
Vs do not perfectly adhere to the scaling since the scaling constraint is effectively relaxed in the last iteration of
the tomographic process to accommodate potential compositional and other effects (see supporting informa-
tion). Perhaps not surprisingly, we do not need to invoke substantial heterogeneity that cannot be directly attrib-
uted to variations in temperature similar to our previous findings [Simmons et al., 2009, 2010]. Additional tests
show that any scaling factors between Vp and Vs heterogeneity that are generally consistent with mineral physics
yields similar images of the slab-like anomalies observed in our preferred model (Figure S17). While a “thermal”
model fits the data, this certainly does not rule out the existence of compositional variations that might
contribute to the velocity heterogeneity in the midmantle and elsewhere.

4. Subduction Timing and Additional Clues


It is evident that the areal patterns of fast anomalies in Figures 2a–2d systematically shift southwestward with
decreasing depth (Figure 2e), indicative of a subduction zone that retreated southwestward beneath the cur-
rent site of the southern Indian Ocean and southern Australia. Such a subduction episode is not present or
predicted in recent global plate reconstructions [e.g., Torsvik et al., 2010; Seton et al., 2012] likely because
of the lack of prior tomographic or accompanying geologic evidence to support its existence. The timing
of this event is therefore in question and open for debate.
The general tectonic setting of the Indian Ocean region is generally well understood during the Cenozoic Era
(0–65 Ma), indicating that the SEIS subduction episode occurred earlier. Prior to the Cenozoic, India separated
from East Gondwana (which consisted of India, Australia, Antarctica, and Madagascar) along a spreading center
that began developing at 130–140 Ma [Gibbons et al., 2013]. After separation, India traveled northward across
the ancient Tethys Ocean, eventually colliding with Eurasia and closing the Tethys Ocean. The long-term
spreading and the overall north-northeast motion of India seem to preclude the simultaneous existence of a
long-lived subduction zone sweeping southwestward across the Tethys/Indian Oceans. Based on this general
tectonic history and the position of the SEIS anomalies, we deduce that the subduction episode must have
ended before (or contemporaneous with) the early stages of the breakup of East Gondwana (130–140 Ma).
Although many plate reconstructions (with various frames of reference) place Australia farther north than
the transition zone SEIS anomaly at the time of East Gondwana breakup, a recent reconstruction places East
Gondwana immediately south of the final position of the inferred subduction zone at 140 Ma [Wu and
Kravchinsky, 2014] (Figure 2f). A minor caveat is that the northern extent of the Australian part of East
Gondwana (where New Guinea resides) slightly overlaps the SEIS anomalies in the transition zone. A 5°
Euler rotation of the reconstructed continent (a reasonable adjustment according to the authors L. Wu
and V. Kravchinsky (personal communication, 2015)) positions the subduction zone on the northern edge

SIMMONS ET AL. ANCIENT SLAB BENEATH THE INDIAN OCEAN 5


Geophysical Research Letters 10.1002/2015GL066237

of East Gondwana including the northward excursion of the Australian part of the continent. From this
spatial relationship, we propose that the subduction zone terminated near the edge of East Gondwana
around 130–140 Ma (Early Cretaceous). Prior to this time, the position of deep-mantle SEIS anomalies
marking the early stages of subduction would have been located beneath the ancient Tethys Ocean, signifying
that this was an intraoceanic subduction zone, possibly beginning in the Early Jurassic or Late Triassic.
Inferring ancient intraoceanic subduction events based on kinematic plate reconstruction is difficult since the
plates may have been completely consumed and geologic evidence becomes more obscured with time, with
perhaps only a few scattered relics near the surface to document the event. Remnant seamounts and oceanic
plateaus could have been subducted or otherwise destroyed during subsequent tectonic events, but it is
expected that intraoceanic subduction would produce a scattered set of volcanic island arcs that are poten-
tially observable in geologic structures today. One such volcanic arc terrane structure is found on the outer
edges of the island of Sumatra in Indonesia which may be a remnant of the SEIS subduction event
(Figure 2f). Specifically, the Woyla Group is believed to represent a Jurassic-Early Cretaceous oceanic volcanic
arc system that formed somewhere within the Tethys Ocean [Barber, 2000; Barber and Crow, 2003] and
accreted onto Sumatra at ~90 Ma [Hall, 2012]. The Woyla Group consists of arc volcanics and limestones
believed to be reefs that existed on the edge of former volcanic islands [Barber and Crow, 2003]. The actual
subduction episode responsible for the Woyla terrane construction is unknown, and the age fits with the time
frame that we infer for the SEIS subduction event.
One concern is how the Woyla terrane became part of Sumatra and not part of India. We believe the island arc
did not actually collide with East Gondwana due to a relatively abrupt shift in Tethyan tectonics around
140 Ma, which caused the SEIS subduction to cease and a spreading regime to begin shortly thereafter. It
is explicitly shown and discussed in the recent plate reconstruction in Seton et al. [2012] that a mid-ocean
ridge converged with Eurasia along the northern Tethyan subduction zone at around 140 Ma. In this recon-
struction, the ridge convergence caused a “southern ridge jump” and a new spreading center quickly formed
along the margin of East Gondwana on the opposite side of the Tethys Ocean. Although the introduction of
the hypothesized SEIS subduction event alters the overall kinematic history of the Tethys region depicted in
Seton et al. [2012] prior to 140 Ma, the same (or similar) ridge jump event could have still occurred with some
alteration to ancient spreading and convergence rates. Such a spreading event would have separated the
volcanic island arcs from continental East Gondwana, preventing them from ever colliding and accreting.
In the specific case of the Woyla terrane, it would have continued to separate from India as new oceanic crust
was formed along the margin of East Gondwana [see Seton et al., 2012]. The terrane would have then collided
with Sumatra at ~90 Ma as India began its northern journey across the Tethys Ocean [Hall, 2012]. It is
intriguing to consider that the position of the new spreading center that is believed to have developed along
the northern margin of East Gondwana at ~140 Ma may be directly related to the subduction, perhaps
forming along a zone of weakness between the two (formerly) convergent plates.
Another geologic feature potentially related to this subduction event involves an enigmatic part of the SEIR
known as the Australian-Antarctic Discordance (AAD) near 125°E longitude (Figure 2a). The eastern edge of
the AAD marks a sharp isotopic boundary between Indian- and Pacific-type mid-ocean ridge basalts
(MORB) and the excessively high bathymetry of the ridge segments suggests the presence of a cold down-
welling [Klein et al., 1988]. Gurnis et al. [1988] attributes the AAD anomalies to Pacific/Phoenix slab material
subducted in the Mesozoic, still residing in the upper mantle because of mid-ocean ridge-related convection
processes. We indeed see subducted material in the upper mantle in the region, but our image suggests the
AAD is near the intersection of two slabs in the transition zone with different orientations, the SEIS and
Pacific/Phoenix slabs. Rather than a single subduction episode, the unusually sharp isotopic boundary
observed in the MORB on the eastern edge of the AAD could be related to these intermingled slabs of
different origins.
The Indian MORB isotopic composition is known to be distinct from other mid-ocean ridge basalts, which
could be additional evidence for the existence of a large-scale subduction episode. Kempton et al. [2002] con-
cludes that the isotopic signatures are consistent with subduction-modified mantle wedge material injected
into the MORB source region (the upper mantle). Previously, this mechanism was most viable for the AAD
region since subduction of the PAC/PHX plate was expected to have occurred nearby. In light of the new
evidence for subduction occurring beneath the entire length of the present-day SEIR, this mechanism could

SIMMONS ET AL. ANCIENT SLAB BENEATH THE INDIAN OCEAN 6


Geophysical Research Letters 10.1002/2015GL066237

be reasonably expanded to a broader Indian MORB source region as previously conjectured [Kempton et al.,
2002], explaining some of the distinct ocean-scale MORB characteristics.

5. Long-Term Slab Stagnation


Slabs that have been stalled or stagnated in the transition zone have been seismically imaged in multiple
regions [e.g., van der Hilst and Seno, 1993; Fukao et al., 2009]. There is also evidence that slabs may stall
and sink very slowly in some cases, leaving behind upper mantle remnants that subducted during the late
Mesozoic [Li and van der Hilst, 2010; Schmandt and Lin, 2014]. If timing of the subduction event we infer is
correct, portions of the SEIS slab have remained within the transition zone longer than any known slab, at
least any known slab that eventually does penetrate into the lower mantle. Recent estimates of subducted
slab sinking rates are 1.3 ± 0.3 cm/yr through the whole mantle on average [Butterworth et al., 2014]. If
subduction of the SEIS began at ~200–250 Ma, then the sinking rate of the deep-mantle part of the slab is
consistent with the low-end value (1 cm/yr). However, the material we observe in the transition zone near
the SEIR (inferred to have subducted >130 Myr ago) is averaging a much lower sinking rate (~0.4 cm/yr).
This low sinking rate likely reflects slab stagnation at the upper-lower mantle boundary for upward of
100 Myr if our interpretation is indeed correct. Interestingly, the global estimates of sinking rates in the lower
mantle (excluding the upper mantle) are similar to the whole mantle rate or potentially much greater [van der
Meer et al., 2010; Steinberger et al., 2012]. Given that slabs are expected to sink perhaps 4–5 times faster in the
upper mantle owing to the lower viscosity [e.g., Ricard et al., 1993], these whole and lower mantle sinking
rates combined imply substantial delays in the transition zone on average [Butterworth et al., 2014]. The sink-
ing rate estimates and disparities might indicate that many slabs seen in the lower mantle today stalled in the
transition zone for tens of millions of years, and our interpretation implies that the stall time could exceed
100 Myr in some cases.
The fundamental causes for stagnation are the phase transitions and viscosity contrast near ~660 km depth.
The metastability of olivine [e.g., Tetzlaff and Schmeling, 2000] and pyroxene [e.g., King et al., 2015] in cold sink-
ing slabs inhibits penetration into the lower mantle and the viscosity increase presents a rheological resistance
to flow. There are many additional factors that control the delay time in the transition zone, including the age of
the downgoing slab and the subduction zone trench migration history [e.g., van der Hilst and Seno, 1993;
Christensen, 1996; Torii and Yoshioka, 2007; Goes et al., 2008; Čízková and Bina, 2013; Sharples et al., 2014].
^

Trench retreat (i.e., rollback) promotes stagnation since the slab is flattened and subsequent subduction does
not pile on to force the material downward. Older (colder) subducting slabs are more prone to stagnation
due to mineralogical effects (i.e., metastability) and because they tend to increase trench rollback rates
and lie down flat [Goes et al., 2008; Čízková and Bina, 2013]. In addition, the properties of the overriding plate
^

strongly control trench retreat/advance and whether stagnation mode is entered [Sharples et al., 2014; Holt
et al., 2015]. A thin overriding plate (such as young oceanic lithosphere) leads to fast trench retreat and thus
stagnation. At the same time, slab rollback results in significant back-arc spreading and the thinning or
generation of young oceanic lithosphere; therefore, a positive feedback loop-promoting continuous slab rollback
is conceivable. Numerical simulations have demonstrated stagnation lasting beyond 80–140 Myr using a variety
of rheological properties, subduction behaviors, and slab geometries [Christensen, 1996; Čízková and Bina, 2013;
^

Sharples et al., 2014; Jarvis and Lowman, 2007]. Our new observations may support these classes of numerical
simulations with long-term slab stagnation in the upper mantle transition zone, largely driven by significant
subduction zone retreat.

6. Discussion and Future Research


Our tectonic interpretation implies that the SEIS slab stalled in the transition zone longer than any known slab
before eventually penetrating into the lower mantle. How can a slab be delayed so long but eventually
penetrate into the mantle? We speculate that the SEIS subduction featured an increasingly older/thicker slab
sinking beneath a young/thin oceanic plate and the subduction zone rollback rate increased with time further
promoting transition zone stagnation as the trench approached East Gondwana. Subduction would have been
more “normal” in the early stages, and the slab could have penetrated more readily into the lower mantle. As
the trench retreated more rapidly, longer stagnation times would have occurred in the latter stages. The process
we envision is similar to a recent 3-D spherical convection simulation [Yanagisawa et al., 2010] with penetrating

SIMMONS ET AL. ANCIENT SLAB BENEATH THE INDIAN OCEAN 7


Geophysical Research Letters 10.1002/2015GL066237

slabs that flatten and stall in the upper mantle once trench retreat occurs. Therefore, there can be the simulta-
neous existence of penetration and extended stagnation observed within a single subduction episode.
Another mechanism may be at work to prevent the SEIS slab from sinking into the lower mantle as fast as
other slabs. Specifically, Gurnis et al. [1988] concluded that the slab beneath the eastern part of the SEIR is
currently being drawn upward by the convection currents that result from the mid-ocean ridge above the
slab. They numerically demonstrate that a slab beneath a ridge can be trapped in the transition zone beyond
110 Myrs after subduction while also being attached to a portion in the lower mantle. Given that the SEIR
passed over the SEIS structure in the transition zone and now resides approximately above much of the
structure today (Figure 2a), this could be a significant contributor to delayed slab penetration into the lower
mantle, setting it apart from other subduction/sinking scenarios that have been observed thus far.
Aside from the Woyla volcanic arc terrane now in Sumatra, we have not yet identified other volcanic arc terrane
candidates generated along other parts of the SEIS subduction zone. We note that there is ample opportunity for
other related arcs to have been altered, destroyed, or obfuscated by subsequent tectonic events and magmatic
overprints. Nonetheless, further geologic investigation and possible reinterpretations of existing geologic data
may uncover other terranes that were separated from continental East Gondwana during the same spreading
event that separated the Woyla arc (as well as continental fragments) from the former supercontinent.
The absolute positioning of the continents at and before 140 Ma is open to debate, and there is a great deal of
research dedicated to establishing an appropriate frame of reference for absolute plate reconstructions [e.g.,
Müller et al., 1993; Steinberger and Torsvik, 2008; Torsvik et al., 2008; van der Meer et al., 2010]. Shephard et al.
[2012] show that there is as much as 30° variation in paleo-plate boundary position (latitude and longitude) at
140 Ma. While the position of the continents in Wu and Kravchinsky [2014] is compatible with our postulated
subduction scenario, many reconstructions place the northernmost edge of Australia ~10° (or more) farther
north which overlaps the fast material we observe in the transition zone southwest of Tasmania. Resolving
these discrepancies in the context of our observations should be a future topic of research.

Acknowledgments References
We thank the Office of Nuclear Detonation
Detection within the National Nuclear Antolik, M., Y. J. Gu, G. Ekström, and A. M. Dziewonski (2003), J362D28: A new joint model of compressional and shear velocity in the Earth’s
Security Administration for their mantle, Geophys. J. Int., 153, 443–466.
support. Jeroen Ritsema and an Auer, L., L. Boschi, T. W. Becker, T. Nissen-Meyer, and D. Giardini (2014), Savani: A variable resolution whole-mantle model of anisotropic shear
anonymous reviewer provided velocity variations based on multiple data sets, J. Geophys. Res. Solid Earth, 119, 3006–3034, doi:10.1002/2013JB010773.
constructive reviews that helped Barber, A. J. (2000), The origin of the Woyla terranes in Sumatra and the Late Mesozoic evolution of the Sundaland margin, J. Asian Earth Sci.,
improve the manuscript. We are 18, 713–738.
grateful for the insight provided by Barber, A. J., and M. J. Crow (2003), An evaluation of plate tectonic models for the development of Sumatra, Gondwana Res., 6, 1–28.
David Rowley, Chris Scotese, Lei Wu, Butterworth, N. P., A. S. Talsma, R. D. Müller, M. Seton, H. P. Bunge, B. S. A. Scuberth, G. E. Shepard, and C. Heine (2014), Geological, tomographic,
and Vadim Kravchinsky. We also kinematic and geodynamic constraints on the dynamics of sinking slabs, J. Geodynamics, 73, 1–13.
thank the several global tomography Christensen, U. R. (1996), The influence of trench migration on slab penetration into the lower mantle, Earth Planet. Sci. Lett., 140, 27–39.
experts that made their models avail- Číz ková, H., and C. Bina (2013), Effects of mantle and subduction-interface rheologies on slab stagnation and trench rollback, Earth Planet.
^

able for comparison. This work per- Sci. Lett., 379, 95–103.
formed under the auspices of the U.S. Fukao, Y., M. Obayashi, T. Nakakuki, and Deep Slab Project Group (2009), Stagnant slab: A review, Annu. Rev. Earth Planet. Sci., 37, 19–46.
Department of Energy by Lawrence Gibbons, A., J. Whittaker, and R. Müller (2013), The breakup of East Gondwana: Assimilating constraints from Cretaceous ocean basins around
Livermore National Laboratory India into a best-fit tectonic model, J. Geophys. Res. Solid Earth, 118, 808–822, doi:10.1002/jgrb.50079.
under contract DE-AC52-07NA27344. Goes, S., F. A. Capitanio, and G. Morra (2008), Evidence of lower-mantle slab penetration phases in plate motions, Nature, 451, 981–984.
LLNL-JRNL-668546. Grand, S. P. (2002), Mantle shear-wave tomography and the fate of subducted slabs, Phil. Trans. R. Soc. London A, 360(1800), 2475–2491.
Grand, S. P., R. D. Van der Hilst, and S. Widiyantoro (1997), Global seismic tomography: A snapshot of convection in the Earth, GSA Today, 7(1–7).
Gurnis, M., R. D. Müller, and L. Moresi (1988), Cretaceous vertical motion of Australia and the Australian-Antarctic discordance, Science, 279,
1499–1504.
Hall, R. (2012), Late Jurassic-Cenozoic reconstructions of the Indonesian region and the Indian Ocean, Tectonophysics, 570–571, 1–41.
Holt, A. F., T. W. Becker, and B. A. Buffett (2015), Trench migration and overriding plate stress in dynamic subduction models, Geophys. J. Int.,
201, 172–192.
Houser, C., G. Masters, P. Shearer, and G. Laske (2008), Shear and compressional velocity models of the mantle from cluster analysis of long-
period waveforms, Geophys. J. Int., 174, 195–212.
Jarvis, G. T., and J. P. Lowman (2007), Survival times of subducted slab remnants in numerical models of mantle flow, Earth Planet. Sci. Lett.,
260, 23–36.
Kempton, P. D., J. A. Pearce, T. L. Barry, J. G. Fitton, C. Langmuir, and D. M. Christie (2002), Sr-Nd-Pb-Hf isotope results from ODP Leg 187:
Evidence for mantle dynamics of the Australia-Antarctic discordance and origin of the Indian MORB source, Geochem. Geophys. Geosyst.,
3(12), 1074, doi:10.1029/2002GC000320.
Kennett, B. L. N., and A. Gorbatov (2004), Seismic heterogeneity in the mantle-strong shear wave signature of slabs from joint tomography,
Phys. Earth Planet. Int., 146, 87–100.
King, S. D., D. J. Frost, and D. C. Rubie (2015), Why cold slabs stagnate in the transition zone, Geology, 43, 231–234.
Klein, E. M., C. H. Langmuir, A. Zindler, H. Staudigel, and B. Hamelin (1988), Isotope evidence of a mantle convection boundary at the
Australian-Antarctic discordance, Nature, 333, 623–629.

SIMMONS ET AL. ANCIENT SLAB BENEATH THE INDIAN OCEAN 8


Geophysical Research Letters 10.1002/2015GL066237

Lekić, V., and B. Romanowicz (2011), Inferring upper-mantle structure by full waveform tomography with the spectral element method,
Geophys. J. Int., 185, 799–831.
Li, C., and R. D. van der Hilst (2010), Structure of the upper mantle and transition zone beneath Southeast Asia from traveltime tomography,
J. Geophys. Res., 115, B07308, doi:10.1029/2009JB006882.
Masters, G., G. Laske, H. Bolton, and A. M. Dziewonski (2000), The relative behavior of shear velocity, bulk sound speed, and compressional
velocity in the mantle: Implications for chemical and thermal structure, in Earth’s Deep Interior: Mineral Physics and Tomography From the
Atomic to the Global Scale, edited by S.-I. Karato et al., pp. 63–87, AGU, Washington, D. C.
Müller, R. D., J.-Y. Royer, and L. A. Lawver (1993), Revised plate motions relative to the hot- spots from combined Atlantic and Indian Ocean
hotspot tracks, Geology, 21, 275–278.
Myers, S. C., G. Johannesson, and W. Hanley (2007), A Bayesian hierarchical method for multiple-event seismic location, Geophys. J. Int., 171,
1049–1063.
Myers, S. C., G. Johannesson, and W. Hanley (2009), Incorporation of probabilistic seismic phase labels into a Bayesian multiple-event seismic
locator, Geophys. J. Int., 177, 193–204.
Myers, S. C., G. Johannesson, and N. A. Simmons (2011), Global-scale P wave tomography optimized for prediction of teleseismic and
regional travel times for Middle East events: 1. Data set development, J. Geophys. Res., 116, B04304, doi:10.1029/2010JB007967.
Ricard, Y., M. Richards, C. Lithgow-Bertelloni, and Y. Le Stunff (1993), A geodynamic model of mantle density heterogeneity, J. Geophys. Res.,
98, 21,895–21,909, doi:10.1029/93JB02216.
Richards, M. A., and D. C. Engebretson (1992), Large-scale mantle convection and the history of subduction, Nature, 355, 437–440.
Ritsema, J., A. Deuss, H. J. van Heijst, and J. H. Woodhouse (2011), S40RTS: A degree-40 shear-velocity model for the mantle from new
Rayleigh wave dispersion, teleseismic traveltime and normal-mode splitting function measurements, Geophys. J. Int., 184, 1223–1236.
Schmandt, B., and F.-C. Lin (2014), P and S wave tomography of the mantle beneath the United States, Geophys. Res. Lett., 41, 6342–6349,
doi:10.1002/2014GL061231.
Seton, M., et al. (2012), Global continental and ocean basin reconstructions since 200 Ma, Earth Sci. Rev., 113, 212–270.
Sharples, W., M. A. Jadamec, L. N. Moresi, and F. A. Capitanio (2014), Overriding plate controls on subduction evolution, J. Geophys. Res. Solid
Earth, 119, 6684–6704, doi:10.1002/2014JB011163.
Shephard, G. E., H. P. Bunge, B. S. A. Schuberth, R. D. Müller, A. S. Talsma, C. Moder, and T. C. W. Landgrebe (2012), Testing absolute plate
reference frames and the implications for the generation of geodynamic mantle heterogeneity structure, Earth Planet. Sci. Lett., 317–318,
204–217.
Simmons, N. A., A. M. Forte, and S. P. Grand (2009), Joint seismic, geodynamic and mineral physical constraints on three-dimensional mantle
heterogeneity: Implications for the relative importance of thermal versus compositional heterogeneity, Geophys. J. Int., 177(5), 1284–1304.
Simmons, N. A., A. Forte, L. Boschi, and S. P. Grand (2010), GyPSuM: A joint tomographic model of mantle density and seismic wave speeds,
J. Geophys. Res., 115, B12310, doi:10.1029/2010JB007631.
Simmons, N. A., S. C. Myers, and G. Johannesson (2011), Global-scale P wave tomography optimized for prediction of teleseismic and
regional travel times for Middle East events: 2. Tomographic inversion, J. Geophys. Res., 116, B04305, doi:10.1029/2010JB007969.
Simmons, N. A., S. C. Myers, G. Johannesson, and E. Matzel (2012), LLNL-G3Dv3: Global P wave tomography model for improved regional and
teleseismic travel time prediction, J. Geophys. Res., 117, B10302, doi:10.1029/2012JB009525.
Spasojevic, S., M. Gurnis, and R. Sutherland (2010), Mantle upwellings above slab graveyards linked to the global geoid lows, Nat. Geosci., 3,
435–438.
Steinberger, B., and T. H. Torsvik (2008), Absolute plate motions and true polar wander in the absence of hotspot tracks, Nature, 452, 620–623.
Steinberger, B., T. Torsvik, and T. Becker (2012), Subduction to the lower mantle—A comparison between geodynamic and tomographic
models, Solid Earth, 3, 415–432.
Tetzlaff, M., and H. Schmeling (2000), The influence of olivine metastability on deep subduction of oceanic lithosphere, Phys. Earth Planet.
Inter., 120, 29–38.
Torii, Y., and S. Yoshioka (2007), Physical conditions producing slab stagnation: Constraints of the Clapeyron slope, mantle viscosity, trench
retreat, and dip angles, Tectonophysics, 445, 200–209.
Torsvik, T., B. Steinberger, L. Cocks, and K. Burke (2008), Longitude: Linking Earth’s ancient surface to its deep interior, Earth Planet. Sci. Lett.,
276, 273–282.
Torsvik, T., B. Steinberger, M. Gurnis, and C. Gaina (2010), Plate tectonics and net lithosphere rotation over the past 150 My, Earth Planet. Sci.
Lett., 291, 106–112.
van der Hilst, R. D., and T. Seno (1993), Effects of relative plate motion on the deep-structure and penetration depth of slabs below the
Izu-Bonin and Mariana Island arcs, Earth Planet. Sci. Lett., 120(3–4), 395–407.
van der Hilst, R. D., S. Widiyantoro, and E. R. Engdahl (1997), Evidence for deep mantle circulation from global tomography, Nature, 386,
578–584.
van der Meer, D., W. Spakman, D. van Hinsbergen, M. Amaru, and T. Torsvik (2010), Towards absolute plate motions constrained by
lower-mantle slab remnants, Nat. Geosci., 3, 36–40.
van der Voo, R., W. Spakman, and H. Bijwaard (1999a), Tethyan subducted slabs under India, Earth Planet. Sci. Lett., 171, 7–20.
van der Voo, R., W. Spakman, and H. Bijwaard (1999b), A Mesozoic subducted slab under Siberia, Nature, 397, 246–249.
Wu, L., and V. Kravchinsky (2014), Derivation of paleolongitude from the geometric parametrization of apparent polar wander path:
Implication for absolute plate motion reconstruction, Geophys. Res. Lett., 41, 4503–4511, doi:10.1002/2013GL059051.
Yanagisawa, T., Y. Yamagishi, Y. Hamano, and D. Stegman (2010), Mechanism for generating stagnant slabs in 3-D spherical mantle
convection models at Earth-like conditions, Phys. Earth Planet. Int., 183, 341–352.
Zhao, D., Y. Yamamoto, and T. Yanada (2013), Global mantle heterogeneity and its influence on teleseismic regional tomography, Gondwana
Res., 23, 595–616.

SIMMONS ET AL. ANCIENT SLAB BENEATH THE INDIAN OCEAN 9

Vous aimerez peut-être aussi