Control of seafloor aging on the migration of the Izu–Bonin–Mariana trench

14
This article appeared in a journal published by Elsevier. The attached copy is furnished to the author for internal non-commercial research and education use, including for instruction at the authors institution and sharing with colleagues. Other uses, including reproduction and distribution, or selling or licensing copies, or posting to personal, institutional or third party websites are prohibited. In most cases authors are permitted to post their version of the article (e.g. in Word or Tex form) to their personal website or institutional repository. Authors requiring further information regarding Elsevier’s archiving and manuscript policies are encouraged to visit: http://www.elsevier.com/copyright

Transcript of Control of seafloor aging on the migration of the Izu–Bonin–Mariana trench

This article appeared in a journal published by Elsevier. The attachedcopy is furnished to the author for internal non-commercial researchand education use, including for instruction at the authors institution

and sharing with colleagues.

Other uses, including reproduction and distribution, or selling orlicensing copies, or posting to personal, institutional or third party

websites are prohibited.

In most cases authors are permitted to post their version of thearticle (e.g. in Word or Tex form) to their personal website orinstitutional repository. Authors requiring further information

regarding Elsevier’s archiving and manuscript policies areencouraged to visit:

http://www.elsevier.com/copyright

Author's personal copy

Control of seafloor aging on the migration of the Izu–Bonin–Mariana trench

Claudio Faccenna a,⁎, Erika Di Giuseppe a, Francesca Funiciello a, Serge Lallemand b, Jeroen van Hunen c

a Università “Roma TRE”, Dipartimento Scienze Geologiche, Roma, Italyb UUMR 5573, CNRS, Laboratoire Geosciences Montpellier, Montpellier II Université Montpellier, Francec Department of Earth Sciences, Durham University, Durham DH1 3LE, UK

a b s t r a c ta r t i c l e i n f o

Article history:Received 17 January 2009Received in revised form 23 September 2009Accepted 26 September 2009Available online 22 October 2009

Editor: R.D. van der Hilst

Keywords:subduction zonetrenchesWestern Pacificnumerical modelling

Recent global kinematic studies reveal that most of the trenches roll back but a significant number of themadvance toward the upper plate. Those advancing trenches are mostly located in the Western Pacific andcorrespond to the subduction of very old, Mesozoic oceanic lithosphere. While retreating trenches arecommonly explained by the slab pull action of the descending lithosphere, the origin of advancing trenches isstill debated. Since this relationship is dependent upon the adopted reference frame, we select region wheregeological studies show the variability of trench migration style with time. The Izu–Bonin–Mariana (IBM)region represents a key example. The detailed reconstruction of the trench migration of the IBM subductingsystem reveals that after a long episode of asymmetric rollback, the IBM trench recently started advancing.We propose that this change from retreating to advancing trench mode results from the subduction ofprogressively older and stiffer lithospheric material. We test this hypothesis by means of two-dimensional(2-D) numerical models, reproducing the effects of the lithospheric aging during subduction. The result ofour numerical tests shows that the entrance of old and stiff lithosphere forces the trench to advance becausethe increasing stiffness of the slab prevents the slab to unbend once it has subducted. We adapt this physicalresult to the IBM evolution, showing that the switch from trench retreat to trench advance can reconcile theshape of IBM slab as previously suggested. Finally we discuss the possibility that the subduction of theCretaceous lithosphere might have triggered the change of the motion of the Pacific plate around 5 Ma.

© 2009 Elsevier B.V. All rights reserved.

1. Introduction

Themigration of trenches represents a first order ingredient on thesubduction zone kinematics, as it determines the subduction rateand the back-arc deformation regime (Molnar and Atwater, 1978;Dewey, 1980). Molnar and Atwater (1978) first related the style ofback-arc deformation along the Pacific with the age of subductingseafloor to infer subduction zone dynamics. The basic idea is that anold and cold subducted lithosphere, such as the one subducting alongthe Western Pacific trenches, is expected to produce trench retreatbecause of its negative buoyancy. This style of migration, commonlyreferred as trench “rollback” (Dewey, 1980), was considered as afavorable condition for the opening of back-arc basins. Molnar andAtwater's (1978) model was designed to explain the Pacific dichot-omy where back-arc extension and compression occurred incorrespondence with old and young subducting lithosphere, respec-tively. This model inspired a large number of subsequent studies thatuse trench motion as the result of the lithosphere–mantle interactionduring subduction (e.g., Spence, 1987; Zhong and Gurnis, 1995;Faccenna et al., 2001; Funiciello et al., 2004; Bellahsen et al., 2005;

Enns et al., 2005; Stegman et al., 2006; Faccenna et al., 2007; Schellartet al., 2007; Billen and Hirth, 2007; Capitanio et al., 2007; Doglioni etal., 2007; Funiciello et al., 2008; Goes et al., 2008). If most of thesemodels agree to identify the thermal aging of the oceanic lithosphereas the engine for subduction, they diverge to define the way thepotential energy stored inside the oceanic lithosphere is dissipated.Part of this controversy derived by the different modelling techniquebut also by the loose kinematics constraints on subduction system.The rate of trench migration, for example, is poorly defined as itcannot be directly measured. A possible way is to subtract back-arcdeformation from upper plate motion (Heuret and Lallemand, 2005),but incertitude derives from both back-arc deformation estimate(geodetic or geologic) and from possible erosion or accretion attrench. On this concern, the IBM subduction system, which exem-plifies the back-arc extensional style (Uyeda and Kanamori, 1979),represents a test case to illustrate the relationships between trenchmotion and back-arc stress. The funding that IBM trench is, in fact,presently advancing towards the upper plate (Fig. 1; Carlson andMelia, 1984; Seno andMaruyama, 1984; van der Hilst and Seno, 1993)demonstrates that back-arc extension is related to the north-westward motion of the Philippine Sea plate (PSP), which evidentlymoves at a higher rate than the IBM trench does (Uyeda andKanamori, 1979, Carlson et al, 1983). The counter-intuitive deductionis that, where subducting lithosphere is the oldest on Earth, the trench

Earth and Planetary Science Letters 288 (2009) 386–398

⁎ Corresponding author.E-mail address: [email protected] (C. Faccenna).

0012-821X/$ – see front matter © 2009 Elsevier B.V. All rights reserved.doi:10.1016/j.epsl.2009.09.042

Contents lists available at ScienceDirect

Earth and Planetary Science Letters

j ourna l homepage: www.e lsev ie r.com/ locate /eps l

Author's personal copy

is advancing (Fig. 1) whereas below the Andes, where oceaniclithosphere is young, the trench is retreating (Jarrard, 1986; Sdroliasand Müller, 2006; Schellart et al., 2007). Global compilations onmodern subduction zones confirm these observations: the age ofsubducting lithosphere at the trench inversely scales with trenchmotion, so that old subducting lithospheres are more likely toadvance, whereas the young ones retreat more (Heuret andLallemand, 2005; Faccenna et al., 2007; Lallemand et al., 2008; DiGiuseppe et al., 2008, 2009; Funiciello et al., 2008). Because the trenchvelocity is usually small, however, the relationship between trenchmigration and subducting seafloor age is not robust depending uponthe adopted net-rotationmodel (Fig. 2). As a matter of fact, the decentVt–age relationships observed in both HS3 (Gripp and Gordon, 2002)and SB04 (Steinberger et al., 2004) reference frame model flats downusing GJ86 (Gordon and Jurdy, 1986) or a no-net-rotation (NNR)model (Fig. 2). As the exact amount of net rotation is unknown, and isprobably variable in both space and time, the trench motion attitudeand its bearing with physical subduction parameters are then leftpoorly constrained. A possible solution is to analyze specific regionswhere geological studies ensure that trench migration changedduring time. The IBM subduction zone represents an excellentexample, because – during its history – it showed a remarkablechange in migration style as after a long episode of trench rollback itstarted advancing towards the upper plate. We propose that thischangemay be caused by the subduction of progressively older, Lower

Cretaceous, oceanic lithosphere. We test this idea by means of 2-Dnumerical, fully dynamic models, showing how the style of migrationis related to the age and, in turn, to the stiffness of the oceaniclithosphere. We propose this model as the most appropriate to mimicnot only the history of trench migration of the IBM but also the slabmorphology as revealed by seismic tomography. We finally speculateon the possible influence of the IBM trench on the kinematics changesobserved on the Pacific plate.

2. Migration history and slab morphology below the IBM trench

The PSP is entirely surrounded by subduction zones. Consequently,the reconstruction of its kinematics and the rotation poles can be doneonly indirectly, using information from paleomagnetic data, spreadinghistory and positions of volcanic arcs (see also Seno et al., 1993).Different paleo-tectonic reconstructions have been proposed so far(Rangin et al., 1990; van der Hilst and Seno, 1993; Okino et al., 1994;Hall et al., 1995a,b; Okino et al., 1998; Hall, 2002; Deschamps andLallemand, 2002; Sdrolias et al., 2004; Sdrolias and Müller, 2006;Miller et al., 2006). All of them indicate that back-arc extensionoccurred during an overall northward drifting and clockwise rotationof the PSP (van der Hilst and Seno, 1993; Hall et al., 1995a). Fig. 3shows the paleo-tectonic evolution of the PSP and the position of theIBM trench during the last 50 Ma using the Sdrolias andMüller (2006)plate reconstruction based on O'Neill et al. (2005) moving Indo-Atlantic hot spot track rotation poles. The position of the trench hasbeen corrected for spreading history using the reconstruction of Hallet al. (1995a,b) and Deschamps and Lallemand's (2002) (for time laps>30 Ma). Seafloor age maps are based on the Sdrolias and Müller(2006) reconstruction.

The formation of the IBM trench started around or just before55 Ma as attested by the oldest volcanic rocks drilled onto the activeand remnant arcs (Deschamps and Lallemand, 2002 and referencestherein). The onset of rifting started at that time to the back of thePhilippine arc (Fig. 3a). The spreading and the progressive formationof the West Philippine back-arc basin (WPB) have been related tothe northward rollback and clockwise rotation of the IBM trench(Hall et al., 1995b), consuming 30 to 70 Ma old oceanic lithosphereat an average rate of 5 cm/yr (Figs. 3a, b and 4). Seafloor magneticanomalies show that the spreading ended in early Oligocene around33–30 Ma (Hilde and Lee, 1984; Deschamps and Lallemand, 2002;Fig. 3b).

The second spreading episode produced the Shikoku–Parece-Velabasins (Fig. 3a–c). The rifting initiated at about 30 Ma in the north(Shikoku basin) and 29–26 Ma in the south (Parece-Vela basin; Okinoet al., 1994, 1998; Sdrolias et al., 2004) and ended around 15 Ma. TheShikoku Basin in the north underwent three spreading stages: NNW–

SSE, N–S and NW–SE opening, whereas the Parece-Vela changed itsspreading direction from E–W to NE–SW at about 20 Ma (Okino et al.,1994, 1998; Sdrolias et al., 2004). The changes of the spreadingdirection are interpreted as related to the continuous rotation of thePSP plate (Sdrolias et al., 2004). The half spreading rate ranged from 2to 3 cm/yr increasing afterward to 5–6 cm/yr. The net amount oftrench migration, measured perpendicularly to the trench, during thisepisode is about 1000 km in the north with the consumption of 70to 95 Ma old seafloor lithosphere (Sdrolias andMüller, 2006; Figs. 3b–c and 5) decreasing towards the south to less than 400 km, consuming90 to 115 Ma old oceanic lithosphere. The third episode of back-arcextension is confined to the Mariana trough, which starts riftingaround 7 Ma (Martinez et al., 1995). But at about 5–10 Ma, the rapidthought pulsating course of the IBM trench rollback ceases and trenchbegan to advance towards the west (Carlson and Melia, 1984). Senoand Maruyama (1984) calibrated the advancing motion of the IBMtrench of about 70 km from 10 to 4 Ma onward using the paleo-position of the volcanic arc in Japan. Paleomagnetic data providedconstraints on the timing of clockwise rotation and on the northward

Fig. 1. Trench kinematics in south-east Asia. Grey andwhite arrows show the absoluterate and direction of trenchmigration in theWestern Pacificwith respect to Pacific hotspots (Gripp and Gordon, 2002) and Indo-Atlantic hot spot (Steinberger et al., 2004),respectively. Ocean floor grid is from Müller et al. (1993). Trench migration rate isestimated by subtracting the rate of deformation from the motion of the main upperplates as given by GPS studies, assuming erosion and accretion negligible (see Heuretand Lallemand, 2005 for detailed references). Abbreviations: PSP is Philippine SeaPlate.

387C. Faccenna et al. / Earth and Planetary Science Letters 288 (2009) 386–398

Author's personal copy

drift of the PSP (Hall et al., 1995b) including trench advance duringthe last 5 Ma (Hall, 2002). Based on the migration of the triplejunction between the IBM and Japan trenches, Le Pichon and Huchon(1987) estimated that, at 3 Ma, the trench was about 30 km eastwardof the present-day position. Fig. 1 shows that the Izu–Bonin trenchsegment presently advances at an average rate ranging from 65 mm/yr to 20 mm/yr, whereas the Mariana trench advances at a rate from68 to 40 mm/yr. Fig. 1 also show that higher values are obtained usinga Pacific hotspot reference frame (Gripp and Gordon, 2002), while thelower one is based on the Indo-Atlantic hotspots (O'Neill et al., 2005).The morphology of the Western Pacific subducting plate has beenilluminated by several studies of seismic tomography and seismicity(van der Hilst et al., 1991; van der Hilst and Seno, 1993; Bijwaardet al., 1998; Fukao et al., 2001; Miller et al., 2006). The Izu–Bonin slabis characterized by a 50° to 70° dipping Wadati–Benioff zone, whichflattens out in the transition zone into a long high velocity anomaly,and stalls on the 660-km discontinuity. Northward, below Japan, theslab shallows to a dip of about 30° whereas, southward, below theMariana trench, the slab becomes nearly vertical, and even overturnslocally. The different migration rates between parts of the IBM slab isthought to be accommodated by a slab tear, separating the Marianafrom the Izu–Bonin slab (van der Hilst and Seno, 1993; Miller et al.,2006). Miller et al. (2006) interpreted the seismic gap at mid-uppermantle depth as the slab window related to the subduction of theMarcus-Necker ridge. van der Hilst and Seno (1993) first related thehistory of trench migration with slab morphology, proposing thatmaterial accumulated at the 660-km discontinuity results from therapid trench migration episode of the Izu–Bonin trench. The sub-vertical attitude of the Mariana slab, conversely, could be related tothe fact that the Mariana trench remained fixed during much of itshistory favoring the penetration of the slab in the lower mantle, andits anchoring at depth.

Fig. 4 shows the track of the IBM trench during the last 40 Ma. Itsretrograde motion is clearly asymmetric, as trench retreat wasmostly attained in the northern portion of the trench with respect tothe southern one. The asymmetric trench migration occurred duringthe northward drifting and the clockwise rotation of the PSP (Hallet al., 1995a,b; Hall, 2002). Fig. 5 illustrates the net migrationperpendicular to the trench, along the southern section (across theMariana islands) and across the northern section (across the Boninislands), along with the age of the subducting seafloor as recon-structed by Sdrolias and Müller (2006). This illustrates that thenorthern branch of the IBM trench backwardmigration is three timeslarger than the southern one, but always consuming a 20 Ma youngerseafloor. In other words, at any specific time lap, a younger seafloorretreats faster than an old one. The switch in themigration style, fromretreating to advancing marks an important step in the history of theIBM trench (Seno and Maruyama, 1984; van der Hilst and Seno;1993). The switch in trench motion is asymmetric, migratesnorthward, occurring 15 to 20 Ma ago in the southern section andaround 5 Ma ago in the northern section (Figs. 4 and 5). This changeoccurred during the consumption of 110±10 Ma seafloor in thesouth and 120±10 Ma in the north. However, these age of theseafloor are not representative of the thermal state of the lithosphere.Seismic wave disperison analysis shows that old Pacific floor is sub-stantially thinner than expected, probably due to a heating eventbetween 70 and 100 Ma followed by subsequent cooling (Ritzwolleret al., 2004). As a consequence, the Pacific seafloor that has an agecomprises between 70 and 130 Ma behaves thermally as lithosphereyounger by about 30 Ma.

While the mechanism of trench rollback is relatively well under-stood (e.g., King, 2001), and is in agreement with the slabmorphologyof Izu–Bonin (van der Hilst and Seno, 1993; Guillou-Frottier et al.,1995; Griffiths et al., 1995; Christensen, 1996), the reason why the

Fig. 2. Trench velocity Vt of the natural trenches versus the age of the subducting lithosphere. Vt for the present-day in the HS3, SB04, GJ86 and no-net-rotation reference frames.Seafloor age from Müller et al., (1997). Note the dependence of the scaling relationships with the adopted reference frame and net-rotation.

388 C. Faccenna et al. / Earth and Planetary Science Letters 288 (2009) 386–398

Author's personal copy

IBM trench started to advance towards the upper plate during theNeogene is less intuitive. Three models have been proposed so far.Carlson and Melia (1984) first proposed that the advancing motion ofthe trench could be related to the suction exerted by the fastretreating upper plate. Later, Pacanovsky et al. (1999) estimated,using a spherical shell elastic finite element analysis to characterizeintraplate stresses and plate driving forces in the PSP, that trenchsuction forces can reach no more than 10% of slab pull forces. Thesecondmodel, proposed van der Hilst and Seno (1993), integrated thedeep structure of the slab with trench kinematics. This modelsuggested that the overall clockwise rotation of the IBM trench isrelated to the ongoing penetration of the Mariana slab into the lowermantle and that the recent advancing motion of the IBM trench couldbe related to the northward propagation of this process, about toinitiate also beneath the Izu–Bonin. The third model, proposed byMiller et al. (2004, 2006), attributes the change of the motion of the

Izu–Bonin trench to the subduction of the Necker-Marcus (Ogasawaraplateau) aseismic ridge, at about 8 Ma.

Using numerical dynamical self-consistent models, we hereexplore an alternative possibility, based on the idea that the trenchadvancing could be forced by the subduction of increasingly older andstiffer Pacific seafloor.

3. Numerical models

Several numerical and analogical studies have been performed toanalyze the role of strength on subduction dynamics during the lastdecades. Dynamically self-consistent numerical (Gurnis and Hager,1988; Zhong and Gurnis, 1995; Enns et al., 2005; Stegman et al., 2006,in press; Capitanio et al., 2007; Billen, 2008) and laboratory models(Funiciello et al., 2003; Bellahsen et al., 2005; Faccenna et al., 2007)show that the slab strength plays a key role in controlling subduction

Fig. 3. Evolution of the Izu–Bonin–Mariana subduction system in four steps at 50 Ma (a), 30 Ma (b), 10 Ma (c), and present-day (d). Paleo-oceanic grid and Pacific plate motion fromSdrolias and Müller (2006). Numbers represent age of the oceanic crust (Myrs). Grey regions show all non-oceanic crust, i.e. continental crust, volcanic arcs, and oceanic plateaus.Dashed white lines denote the West Philippine basin, Shikoku and Parece-Vela and Caroline Sea ridges (number are the age of the back-arc oceanic crust). Dashed black lines showtrench position at intermediate step (white age labels). Long arrows represent the modern motion of the major plates. Abbreviations: PSP is Philippine Sea Plate.

389C. Faccenna et al. / Earth and Planetary Science Letters 288 (2009) 386–398

Author's personal copy

velocity and its partitioning between trench and plate motion. Tankexperiments of Bellahsen et al. (2005) first point out that a switchfrom a “retreating” to an “advancing” style can be easily obtained ifthe bending resistance is increased relatively to the slab pull force.Bellahsen et al. (2005) and Funiciello et al. (2008) show also that this

switch can be obtained only in a narrow field where slab/mantlelithosphere viscosity ratio ranges between 102 and 104, a range inwhich the tank depth (scaled to upper mantle) is about the double ofthe radius of curvature of the slab. Faccenna et al. (2007) applied theexperimental scaling laws to natural system to match both plate andtrench kinematics. But a complete understanding of the experimentalresults needs a more systematic parametric analysis better achievedusing numerical models. On the ground of previous studies (Enns etal., 2005; Stegman et al., 2006), Di Giuseppe et al. (2008) explored therole of subducting lithosphere strength on subduction kinematicsusing 2-D and 3-D numerical solutions. The latter study presented acomplete sensitivity analysis showing the conditions (plate thickness,viscosity, density, pseudoplasticity on the hinge zone, slab width) forthe switch from retreating to advancing style. Even if thesemodels arenot directly comparable to Bellahsen et al. (2005) result, because ofthe lack of free surface (see also Schmeling et al., 2008; Kaus andBecker, 2008), they indeed confirm the possibility to obtain these twosubducting styles without any external forcing (cfr. Schellart, 2005).Recently, Kaus and Becker (2008), Stegman et al. (in press) and Ribe(in press) confirm the Bellahsen et al. (2005) and Di Giuseppe et al.(2008) models showing how trench migration depend upon slabstrength. The application of these results to the Earth's subductingfield highlights the relationships between subducting seafloor andsubduction style (Di Giuseppe et al., 2009).

The numerical simulations presented here are based on the resultsby Di Giuseppe et al. (2008), implemented with new numericalsimulations to analyze the influence of progressive change in the ageof subducting oceanic lithosphere on the kinematics of trenches.

Models simulate subduction with a visco-plastic flow with com-positional buoyancy by using the parallel finite element code CITCOM(e.g. Moresi and Gurnis, 1996; Zhong et al., 2000). The code solves thecontinuity equation, the conservation of momentum equation and theequation for chemical advection. The mantle is treated as an incom-pressible viscous medium, with infinite Prandtl-number and underthe Boussinesq approximations. Our models are purely compositional,

Fig. 4. Migration of the IBM trench during the last 45 Ma (coloured dashed lines) alongwith the corresponding position of northern and southern remnant volcanic arcs(coloured patches, from Hall et al., 1995a,b). The position of the remnant arc is used toestimate the net trench migration, shown in Fig. 5.

Fig. 5. Amount of Izu–Bonin trench migration in time of the Northern section (a), and that of the Southern section (b) estimate from the migration of remnant volcanic arc(see Fig. 4). In the upper part of the panels, along x-axis, is the age of the subducting plate (from Sdrolias and Müller, 2006). The age of the switch from retreating to advancing styleis ~120±10 Ma and ~110±10 Ma for the northern and southern sections, respectively. This age should be reduced to about ~30 Ma considering the apparent thermal age derivedfrom the Ritzwoller et al. (2004) seismic study.

390 C. Faccenna et al. / Earth and Planetary Science Letters 288 (2009) 386–398

Author's personal copy

since temperature is not explicitly solved for. So thermal effects andphase change are neglected. The computational domain of the 2-Dcalculation is a Cartesian box with aspect ratio 7×1 that correspondsto a box of ~4600 km long and 660 kmdeep,with 384 and 64 elementsin horizontal and vertical directions, respectively. The 3-D computa-tional box has a ~4600 km y-axis directions equal to the x- one with aslab width of ~1350 km and a plate length of l~3300 km. 32 elementsare in the vertical direction and 192 in both horizontal directions.

Compositional information is carried by more than hundredtracers per finite element that are transported with the velocityfield (van Hunen et al., 2002), with a second-order Runge–Kuttamethod (Schmeling et al., 2008; Di Giuseppe et al., 2008).

The model set up is designed to let the slab/plate system to adaptin a “dynamic” self-consistent way that is plate is free at ridge (Kincaidand Olson, 1987) from any external force. Therefore, in the model,subduction is driven only by potential energy force and plate bearingslab is free to move (on its right hand side). In addition, to investigatethe trench migration tendency, we eliminate also the possible influ-ence of the upper plate, whose kinematic could be influenced by otherexternal factor (Lallemand et al., 2008).

The upper mantle viscosity is fixed at 1020Pa s, while, lithosphereviscosity is varied between at 5×1022 and 1023Pa s. A completeparametric study on the various physical properties and on its influ-ence on subduction style can be found in Di Giuseppe et al. (2008).Yielding of the slab locally occurs during bending at the trench if thesecond invariant of the stress exceeds the yield stress value, τyield,using Byerlee's empirical frictional law (Byerlee, 1968):

τyield = μf P ð1Þ

with µf the friction coefficient and P the lithostatic pressure. When thebending stresses are larger than τyield, the plate strength is stronglyreduced in the bending zone. Weakening in the bending zone can beexpressed by an effective viscosity, ηyield, calculated as τyield/2 ε

.II, with

ε.II the second invariant of the strain rate (Di Giuseppe et al., 2009).

Therefore, the weakening process is determined by the stress levelacting on the bending zone.

We set a coefficient of internal friction µf of 0.08. Previous tests (DiGiuseppe et al., 2008) show that this represents a most favorablevalue to obtain a realistic slab shape avoiding slab necking. For µf=1dripping instabilities are obtained, whereas for µf<0.01 the slab isdrastically weakened lacking its coherency. Value around 0.1, andmore precisely 0.08, gives a velocity field comparable to laboratory(Bellahsen et al., 2005) and previous numerical modelling results(Enns et al., 2005; Stegman et al., 2006).

The rheology of the slab–plate system does not include elasticity.Elasticity indeed is considered a fundamental ingredient to investi-gate, for example, the response to rapid unloading such as for the caseof slab break-off (e.g., Buiter et al., 2002). However, our simulationsaim to reproduce long-term tectonic processes, longer than theMaxwell relaxation time, where elasticity is considered to play aminor role (Kaus and Becker, 2006; Capitanio et al., 2007).Nonetheless we cannot rule out that elastic deformation would playa role if implemented. Since the calculations don't explicitly taketemperature advection and diffusion into account, no warming andthermal weakening of the subducting lithosphere will occur. Webelieve this to be a reasonable assumption for the purpose of thiswork, as illustrated in previous work (Karato et al., 2001; Cıžková etal., 2002, 2007; van Hunen and van den Berg, 2008). These studiesshow that for realistic subduction velocities of a few cm/yr, thestrength of slabs is largely controlled by a stress-limiting deformationmechanism such as Peierl's stress mechanism (which has only a smalltemperature dependence), and is therefore little influenced bytemperature. Warming of the slab also changes the buoyancydistribution of the downgoing slab: the core of the slabs heat up

and become less negatively buoyant, while the area immediatelysurrounding slabs cools down.

The bottom boundary is associated with the seismic discontinuityat 660-km depth and is modelled as an impermeable barrier. Thissetting is acceptable for an increase in viscosity in the lower mantlelarger than 50–100 Pa s (e.g. Mitrovica, 1996), if the time scale of theinvestigated process is in the order of few ten of million years (e.g.,Davies, 1995). Numerical and laboratory simulations, in fact, showthat direct penetration of slab into the lower mantle, well imaged byseismic tomography (e.g. van der Hilst et al., 1991, van der Hilst andSeno, 1993; van der Voo et al., 1999), can be temporarily inhibited bythe combined effect of the increased viscosity in the lower mantle andof the endothermic phase change and the lateral migration of trenches(e.g., Kincaid and Olson, 1987; Ringwood and Irfune, 1988; Tackley etal., 1993; Zhong and Gurnis, 1995; Christensen, 1995; Griffiths et al.,1995; Guillou-Frottier et al., 1995; Christensen, 1996; Ita and King,1998). In particular, our models are directly comparable with theStegman et al. (2006, 2009) and Schellart et al. (2007) simulationswhere slabs stagnating over the 660-km discontinuity for a viscosityjump in the lower mantle of 100 Pa s.

The boundary conditions are described by a free-slip top and byno-slip bottom and sidewalls. Comparison between 2-D and 3-Dmodelling results shows only slightly time scale differences whilemain results are respected (Di Giuseppe et al., 2008). Lithospherictemperature is converted in density, and taken constant, assumingthat convection is largely more efficient then conduction. Thisassumption is valid and realistic only if subduction rate are largebeing in the order of several cm/yr (Wortel, 1982; Bunge et al., 1997;Goes and van der Lee, 2002). In particular, the progressive agevariation of the oceanic lithosphere is simulated varying the platethickness along its length. Lithospheric thickness is converted in termsof lithospheric age, t. Assuming the half-space cooling model, we setthe thickness of a plate as proportional to the square-root of its age(expressed in seconds), and thermal diffusivity, κ (~10-6 m2s-1)(Turcotte and Schubert, 1982):

h = 2ðκtÞ1=2:

Using this model, the maximum thickness of the lithosphere(~100 km) is attained by a ~80 Ma old lithosphere. Thermal param-eterization, however, based on surface wave dispersion indicates thatover the Pacific lithosphere older than about 70 Ma is significantlyhotter than what predicted for this conductively cooling model andthat a renewed growth of the lithosphere is expected beyond 100 Ma(Ritzwoller et al., 2004). Therefore, although simplified, the half-spacecoolingmodel profile is here adopted as a first order proxy for the age/lithosphere thickness relationships in numerical simulation, beingaware of the possible regional departure from this general law.

Density contrast between lithosphere and mantle is set here as100 kg m−3. Previous tests indicate that changing the density of thesystem influence the time scale of the process but not its physicalbehavior (Di Giuseppe et al., 2008). The density contrast is takenconstant during progressive lithosphere aging. The effective thermalexpansivity regulating the density contrast is tough to be dependentupon the mantle rheology, which is nearly time-independent for theeffect of the viscoelastic response (Korenaga, 2007). Under this con-dition, the density contrast varies of only 2.3% from age ~15 Ma to~150 Ma (Korenaga, 2007). Finally, following previous model, phasechange is not included in the system.

We here show two sets of models to illustrate the influence ofsubducting seafloor age on the trench kinematics. The first set isthree-dimensional and characterized by a constant plate thickness,whereas the second one shows a progressive aging/younging of theplate during continuous subduction.

We select three representative models to illustrate the first set(Fig. 6). These models show the evolution of the subduction process

391C. Faccenna et al. / Earth and Planetary Science Letters 288 (2009) 386–398

Author's personal copy

during subduction of ~60 Ma (model 1), ~77 Ma (model 2), and anintermediate case of ~67 Ma (model 3), corresponding to a constantplate thickness of 84 km, 96 km and 90 km, respectively.

In the first model (model 1, Fig. 6a), the slab sinks sub-verticallyinto the upper mantle accelerating while trench rollback (Fig. 7a,phase I). At 30 Ma, the slab tip hits the 660-km discontinuity verticallyand, afterwards, slab folds back while system slightly decelerate(phase II). During the third phase the subduction system attains asteady state retreating configuration (phase III), with constant trench/plate velocity and invariable geometry with the slab flatting down onthe 660-km discontinuity (phase III, Fig. 7a). The three-phasebehavior described in this model is respected also in the next twomodels and is characteristic of a slab impinging on a high viscousbarrier (e.g., Funiciello et al., 2004; Bellahsen et al., 2005).

The second model has an older slab, being characterized by aconstant lithospheric thermal age of ~77 Ma, corresponding to a platethickness of h~96 km (Fig. 6c, phase I). As in the previous model, theslab begins to sink into the mantle while trench migrates backward(Figs. 6c and 7a), and the subduction velocity accelerates (Fig. 7a). Theoverall slab shape shows a backward reclined shape similar to butmore pronounced than previous model. During the second phase(phase II, Fig. 6c) the slab tip hits the 660-kmdiscontinuity. The trenchsuddenly inverts its direction and starts moving in advancing mode(Fig. 7a). Differently from thepreviousmodel here the slab does not re-fold backward, preserving the backward reclined, hook-shape atti-tude. This process is best illustrated in the third model (model 3,Fig. 6b) characterized by an intermediate thermal age,fixed to ~67 Ma,corresponding to a thickness of h~90 km. During its first phase, theslab tip hits the 660-km discontinuity with a backward reclinedconfiguration similar to model 2. However, after 50–60 Myrs of model

run, the slab folds again backward and trench keeps retreatingbackward reaching a steady state configuration with velocities similarto model 1.

These three models illustrate that the steady state configurationdoes not depend upon the way the tip of the slab hits the 660-kmdiscontinuity, but on the subsequent deep deformation of the slab.The final style of subduction depends upon the possibility of the slabto re-fold backward under the downward push exerted by negativebuoyancy of the subducting material. A weak, young, slab will bendbackward easier than a stiff old one, thereby favoring rollback. Astronger slab, conversely, has difficulties to unbend at depth and thenthe trench/slab system will accommodate its backward reclinedconfiguration by advancing towards the upper plate. This process isthen entirely controlled by the slab strength that is expected to scalewith the cube of the slab thickness (i.e. Turcotte and Schubert, 1982;Becker et al., 1999; Conrad and Hager, 1999; Buffet and Rowley, 2006;Faccenna et al., 2007).

We have performed a large number of simulations to investigate indetails the transition from the retreating (model 1) to advancing(model 2) style of subduction, progressively increasing the thicknessof the lithosphere under different viscosity contrasts (ηl/ηm) betweenlithosphere and mantle. Fig. 7b shows the variation of the trenchvelocity as a function of the subducting plate age. Plate aging producea rather sharp flip of the trench velocity from positive (retreating) tonegative (advancing), achieved once the plate age reaches a thresholdvalue. The switch from one style to the other is, in fact, obtainedincreasing the plate age of only 10–20 Myrs, and characterized by anintermediate step of re-folding subduction style (model 3). Thecritical age for the switch from one style to another scales directlywith the density contrast between the lithosphere and the mantle,

Fig. 6. Lateral views of three representative numerical models characterized by a constant lithospheric thermal age along the plate length of (a) ~60 Ma (model 1), (b) ~67 Ma(model 3) and (c) ~77 Ma (model 2), corresponding to a constant plate thickness of 84 km, 90 km and 96 km, respectively. Arrows describe the position of the trench and of the platein time. Black line shows trench position at the beginning of the subduction process; dashed line follows trench motion in time.

392 C. Faccenna et al. / Earth and Planetary Science Letters 288 (2009) 386–398

Author's personal copy

and inversely with the lithosphere–mantle viscosity contrast (DiGiuseppe et al., 2008, 2009), with a shift of about 50 Myrs halving theviscosity contrast ηl/ηm. Fig. 7c shows that same results hold for 2-Dsetting, although the switch is less abrupt and shifted of fewmillion ofyears.

These models illustrate the way the mode of subduction dependsupon the thermal thickness of the subducting lithosphere. To repro-

duce a more realistic plate configuration, we performed a second setof 2-D models with a progressive variation of the plate age. In model4 (Fig. 8a), the aging of the lithosphere has been simulated bythickening the plate from 30 to 80 Ma (h~67–100 km). In model 5(Fig. 8b), plate thickness is varied from old (~80 Ma) to young(~30 Ma) lithosphere. The starting configuration is characterized byan initial portion (~150 km long) of already subducted lithosphere

Fig. 7. (a) Amount of trench motion versus the time of the subduction process for the numerical models 1, 2, and 3 of Fig. 6. Trench velocity (measure at steady state) Vt vs.lithosphere age for (b) 3-D numerical models (modified from Di Giuseppe et al., 2009) and (c) 2-D numerical models. Symbols represent relative viscosity values: white and greycolours describe retreating and advancing trenches, respectively. Numerical trench velocities of plate with different ages and viscosity ratios fit a transitional trend (Fig. 3a) given byVt=−Vtotan

−1(t− to), where Vto describes the amplitude of trench motion, and to marks the threshold at which the transition from retreating (positive Vt) to advancing style(negative Vt) occurs.

393C. Faccenna et al. / Earth and Planetary Science Letters 288 (2009) 386–398

Author's personal copy

able to ensure the initiation of an efficient subduction. Afterwards, theevolution of the models shows the same characteristic sequences ofevents previously illustrated. During the first phase, the slabprogressively subducts into the upper mantle and its trench migratesbackwards. The rate of trench rollback depends upon the startingconfiguration (Fig. 9). For example, the initial slow rollback rate ofmodel 4 is probably due to the fact that the subducted portion of theslab is thinner than the one entering at trench (Fig. 8a). This producesa progressive incremental increase of the resisting force for bending attrench with respect to the smaller increase in active buoyancy force.

The second transient phase, as already observed, is characterized bythe interaction of the slab with the 660-km deep box bottom (Figs. 8aand 9). Slab folding at depth produces a slow down of subduction forfew Myrs (Figs. 8a and 9), where plate lithosphere progressivelygrows older, the trench continues its retreating trend for about 5 Myrsafter the slab interaction with the 660-km discontinuity at depth. Thishappens during subduction of 40 to 60 Ma seafloor. The arrival ofolder and stiffer lithosphere then causes a progressive decrease intrench migration rate. In a time scale of about 5–8 Myrs, the style oftrench migration progressively changes, while subducted material

Fig. 8. Evolution of the numerical model 4 in which plate lithosphere grows older by thickening from 30 Ma to 80 Ma (a) and model 5 in which plate thickness is varied from old(~80 Ma) to young (~30 Ma) lithosphere (b). Arrows describe the position of the trench and of the plate in time. Black line shows trench position at the beginning of the subductionprocess; dashed line follows trench motion in time.

Fig. 9. Plot showing the amount of trench motion versus the time of the subduction process for the numerical models 4 and 5 of Fig. 8. Symbols identify each run. Lithospheric age ofthe plate at the subduction time is described with different colours. Arrows indicate the moment when the slab tip interacts with the 660 km discontinuity.

394 C. Faccenna et al. / Earth and Planetary Science Letters 288 (2009) 386–398

Author's personal copy

becomes older. The arrival of ~70 Ma old (stiff) material at the trenchslows down the trench migration and then, force the plate to move inadvancing. The variation from the retreating to the advancing styleoccurs when the plate lithosphere is ~80 Ma old.

Model 5 (Fig. 8b), is instead characterized by the progressivethinning of plate thickness. In this case, the trenchmigrates forward inadvance just after the slab interactionwith the upper–lower boundarydiscontinuity. After that, a progressively thinner (and weaker)lithosphere material (<70 Ma) subducts within the mantle, favoringa switch to retreating mode.

4. Discussion

4.1. Mode of subduction: advancing vs retreating trench

Global compilation of kinematic data reveals that trenches whereold lithosphere subducts are more likely advancing (Heuret andLallemand, 2005; Funiciello et al., 2008; Di Giuseppe et al., 2009). Thecritical age at which the change in the subduction behavior occursfrom rollback to advance depends upon the adopted absolute velocitymodel and is estimated to span from 70 Ma for HS3 (Gripp andGordon, 2002) to 90 Ma for SB04 (Steinberger et al., 2004) referenceframes. The thermal aging of the lithosphere is then considered as apotential driving factor for trench migration. This global correlationcould indeed also be accidental, or regulated by the asymmetry of thePacific and by the different momentum generated by upper platekinematics which, on the eastern side, could be particularly important(Husson et al., 2008). In this respect, the case of the IBM trench, wherethe switch in the mode of subduction occurred during its evolution(i.e. independency from selected reference frame), is relevant.

Our numerical simulations show that the retreating mode isfavored for young lithosphere. In terms of subduction dynamics, theseafloor aging can be directly related to thermal thickness and canproduce two competing effects. On one side, it increases the pull forceof the slab. On the other side, it increases the bending resisting force.The latter contribution is probably more efficient as the resistance forbending at trench scales with the cubic power of age, whereas slabpull linearly increaseswith age (Conrad and Hager, 1999; Becker et al.,1999; Buffett and Rowley, 2006, Lallemand et al., 2008). Laboratory(Bellahsen et al., 2005; Funiciello et al., 2008) and numerical models(Di Giuseppe et al., 2008) recently supported this hypothesis andshow that rollover slab related to trench advance is a rather stablefeature. For a given level of slab pull level, the advancing subductionstyle is observed when the resisting force stored in the lithosphereand in the mantle is relatively important. Distribution of energydissipation inside the mantle/slab system shows that in this case, thepull of the slab is not efficient enough to unbend the slab after thebeginning of its travel inside the mantle (Di Giuseppe et al., 2008).Kaus and Becker (2008) recently show that a viscous subductionmodels with free surface give similar geometric and kinematic stylesobtained with Byerlee-type models (Enns et al., 2005; Di Giuseppeet al., 2008). Recent studies also confirms our finding showing thatstiff or thick slabs result in forward trench motion once the slab tiparrives at the 660-km transition (Kaus and Becker, 2008; Ribe, inpress; Stegman et al., in press).

Other numerical tests, conversely, shows that bending contribu-tion of the lithosphere at trench is so small that rollback velocitydirectly scales with thermal plate aging (Capitanio et al., 2007). Onthis concern, it would be probably necessary to compare and bench-mark subduction models, as both set up and boundary conditionsprobably influence the results (Schmeling et al., 2008).

Coupling numerical and laboratory approaches we have found thatadvancing trench is a possible option under reasonable parameters.However, the threshold age found in our models for the switch fromrollback to advance is dependent on several parameters, such as platewidth or lithospheric rheology. Using a robust correlation between

thermal aging and buoyancy, a critical age of 80–90 Ma can beobtained if the viscosity ratio between lithosphere and mantle ofabout 700 (Di Giuseppe et al., 2009). A decrease in this viscosity ratioproduces a linear increase in the critical age (Di Giuseppe et al., 2009).In addition, this critical age could vary by about 10–20 Ma, dependingon the previous history of subduction and deep folding, eventuallydelaying the shift from retreating to advancing mode.

4.2. Possible role of seafloor aging on the migration of theIzu–Bonin trench

The application of the numerical modelling results to the case ofthe IBM migration history is, of course, limited by the inherentsimplicity of the modelling set up. The first simplification is probablyrelated to the possible 3-D effects related in the along strike variationof seafloor age during subduction (Fig. 4). Other possible effect isrelated to the lack of the upper plate, although its role would beprobably minor considering the fact the PSP is moving westward,escaping away from trench faster than the rate of trench advance(Pacanovsky et al., 1999). The other possible influence could be playedby temperature, set constant here as in several previous regional-stylemodel (e.g. Funiciello et al., 2003; Enns et al., 2005; Stegman et al.,2006; Capitanio et al., 2007). Temperature could be influent as thelower portion of the slab is possibly heated up, and, in turn, weakenedduring the subduction process. However, tomography images andtheoretical models illustrates that in a time span of only few millionyears, being the subducting rate of IBM trench in the order of 10 cm/yr, the slab is able to preserve its initial temperature. The last problem,involving also previous models proposed for the IBM trench, is relatedto the anchoring effect possibly exerted by penetration of the slab intothe lower mantle. Indeed, slab penetration into lower mantle isexpected to produce an anchoring effect (e.g., Uyeda and Kanamori,1979), stabilizing the trench (Zhong and Gurnis, 1995; Ita and King,1998; Goes et al., 2008). However, previous model show that in thetime span of 20–40 Myrs, the 660-km discontinuity inhibited directpenetration by the concurrent action of the endothermic phasechange and of the viscosity increase (Hager, 1984; Ringwood andIrfune, 1988; Tackley et al., 1993; Zhong and Gurnis, 1995; Ita andKing, 1998; Stegman et al., 2006; Schellart et al., 2007). This effectcould be enhanced by the retreatingmotion of the trench, favoring thespreading of the subducted lithosphere over the 660-kmdiscontinuity(Guillou-Frottier et al., 1995; Griffiths et al., 1995; Christensen, 1996).Tomographic images show that for the case of the Izu–Bonin trench,where rollback was pronounced, the lithosphere is indeed stagnantover the 660-km discontinuity, whereas partial penetration has beenproposed for the case of the Mariana slab (van der Hilst and Seno,1993; Bijwaard et al., 1998; Fukao et al., 2001; Miller et al., 2006).

Being aware of the inherent assumptions adopted in our model,we suggest that the trench migration history of the IBM trench can beinterpreted using the physical relationships derived by the numericalsimulations. Two simple considerations can be drawn from themigration history of the IBM trench. First, the subducting lithospherein the northern part of the IBM trench was on average 20 Myrsyounger than in the south. From our numerical models, the differentrate of backward migration, three times higher in the north withrespect to the south, could be explained by this age difference. Inparticular, the stationary tendency of the Mariana trench could beinterpreted as a result of slab re-folding process, where material islikely to accumulate over the transition zone. As proposed by van derHilst and Seno (1993), this process could favor partial penetration ofslab tip into the lower mantle, producing an anchoring effect and,in turn, trench stabilization. This process could be responsible forthe clockwise rotation and northward drift of PSP plate. Second,the change from retreating to advancing mode occurred during theentrance at trench of the 110–120±10 Ma old Cretaceous seafloor.This process is then similar, although with a substantial delay, with

395C. Faccenna et al. / Earth and Planetary Science Letters 288 (2009) 386–398

Author's personal copy

the tendency observed in numerical models. This delay can beexplained not only by the possible uncertainties in the modellingparameters but also by the concurrent action of the velocity of theincoming plate and rotation of the Philippine Sea plate. However itcan be considerably reduced considering that seismic wave dispersionof the Central-Western Pacific plate found that 70–130 Ma oldseafloor is substantially thinner than expected from an half-spacecooling model (Ritzwoller et al., 2004). Adopting this recent seismicmodel, the switch from retreating to advancing occurred when aseafloor with an apparent thermal age of about 80–90±20 Ma wasentering at trench, in agreement with model expectation.

In addition, the rotation of the PSP could be driven also by other,external pull, such as the one exerted by the onset of subduction onthe Ryukyu trench or the interaction with the Japan slab. All thesemechanisms can represent external factors influencing the subduc-tion kinematics. In this sense, more sophisticated 3-D models shouldbe integrated to better reproduce the role of the lateral variations ofthe age of the subducting lithosphere.

Despite the simplifications of the model, we are confident that ourresults are robust in terms of subduction dynamics. The subduction ofthe oldest core of the Pacific could have induced the recent episode oftrench advancing, because lithosphere stiffness precludes the possi-bility to unbend the slab at intermediate depth. This model is thenable to reconcile slab shape and trench kinematics. As proposed byvan der Hilst and Seno (1993), the shape of the Izu–Bonin slab can bebest explained by the superimposition of a short pulse of advancingtrench with the long-term history of trench retreat during theprogressive subduction of older and stronger oceanic lithosphere(Fig. 10). Similarly, this model can explain the steep, backwardreclined profile of the Mariana slab (van der Hilst and Seno, 1993;Miller et al., 2006). In the Izu–Bonin the advancing migration of thetrench could have produce a steepening of the slab in the order of 10°.In addition, this model is able to reconcile the onset of compressionaldeformation observed north of IBM trench. Significant compressionaldeformation has been recognized along the eastern margin of theJapan Sea (Tamaki and Honza, 1984; Lallemand et al., 1985). Theinitiation of the compressive phase has been dated around 2–3 Ma(Lallemand and Jolivet, 1986) and its present-day magnitude is esti-mated to reach 1.5 cm/yr based on GPS data (Mazzotti et al., 2001).The southern Kuriles presently undergo compression in the back-arcarea as revealed by seismicity (Savostin et al., 1983). We thus suggest

that the trench migration mode change is not only limited to IBM butmigrated further to the north, starting at the latitude of the Marianaislands about 15–20 Ma ago, then reaching the Bonin islands latitudearound 10–5 Ma ago, then northern Japan–southern Kuriles latitudearound 5–0 Ma ago. This trend diachronously reflects the arrival attrench of progressively older lithosphere.

5. Implication for Pacific plate kinematics

The IBM trench represents an outstanding example of a well-documented switch from retreating to advancing trench, extendingprobably further north in Japan and Kuriles. Based on ocean floor agereconstructions, the Western Pacific subduction zone never involvedthe subduction of a seafloor as old as Late Jurassic–Early Cretaceousbefore present (e.g., Scotese et al., 1988; Richards and Engebretson,1992; Müller et al., 1993). In this view, the advancing stage of IBMcould represent an unusual episode in the evolution of the Pacificmargin, related to the arrival of this anomalous, stiff core of oceaniclithosphere.

Numerical and laboratory models also show that the trenchbehavior can modulate the plate velocity (Iaffaldano et al., 2006) andthat advancing trench is accompanied by a larger velocity of theincoming plate with respect to the retreating configuration (Faccennaet al., 2007). If this kinematic relationship is correct, then the northwardpropagation of the trench migration switch from about 15 Ma ago inthe Marianas until present in the Japan–Kurile area, could haveproduced an acceleration and perhaps the clockwise rotation on thePacific plate during the last Ma. Indeed, several authors described achange in the Pacific motion during the last few million of years. Coxand Engebretson (1985) and Pollitz (1986) describe a change in thePacific absolutemotion, i.e., clockwise rotation and acceleration, around5–3 Ma. Other kinematic models (Harbert and Cox, 1989; Cande et al.,1995; Wessel and Kroenke, 2000) confirm such change between 7 and3 Ma. Pollitz (1986) related this process to a change in the subductionvelocity along the western Pacific trench, whereasWessel and Kroenke(2000) proposed that it coincides with the collision of the Ontong-Javaplateau. Here we proposed an alternative mechanism, related to theentrance at trench of the Early Cretaceous lithosphere. The oldlithosphere likely induces a change from retreating to advancing trenchstyle propagating northward along the Central-Western Pacific and, inturn, producing a clockwise rotation of the subducting plate. Thissuggestion should be considered as only preliminary and more refined3-D models would be necessary to strengthen this hypothesis.

6. Conclusion

Global survey on trench kinematics reveals that in the WesternPacific, independently from the adopted reference frame, there aresome trenches that are presently advancing. We here reconstructedthe history of trench migration of the IBM trench finding that anadvancing episode followed the long history of retreating trench,leading to the opening of the WPB and then the Shikoku–Parece-VelaBasin. The episode of advancing trench occurred when old(>100 Myrs) seafloor entered at trench. We simulate this processusing 2-D fully dynamical numerical models. Those models show thatthe entrance of progressively old seafloor at the trench induced atrench advance because of the increased strength of the subductedlithosphere prevents the slab from unbending. This model can beapplied to the case of the IBM trench reconciling slab shape with therecent switch in trench kinematics from retreating to advancing style.Because the subducting Pacific seafloor has never been as old as atpresent time one may consider the advancing trenches as a newcondition in the history of plate tectonics. We hypothesize thatthe advancing episode of the IBM could have triggered the well-documented change in the absolute motion of the Pacific plate around5–3 Ma.

Fig. 10. Cartoon of the slab evolution below the Izu–Bonin trench, inspired by van derHilst and Seno (1993). Light greydark grey old lithosphere. From 30 Ma to 15 Ma slabrolls back and young lithosphere material lays down in the 660 km discontinuity (a).From 10 Ma to present-day, oldmaterial starts to sink and trench advances, producing asteep slab morphology (b).

396 C. Faccenna et al. / Earth and Planetary Science Letters 288 (2009) 386–398

Author's personal copy

Acknowledgements

Thispaperbenefits fromtheconstructivediscussionswithSaskiaGoes,Fabio A. Capitanio, Boris Kaus, Thorsten Becker and Gabriele Morra onsubduction dynamics andwith Bernhard Steinberger, ThorstenW. Beckerand Richard Gordon on absolute plate motion. We thank DomenicoGiardini and Renato Funiciello for continuous support, encouraging theRome TRE-ETHZ collaboration. Rob van der Hilst and Susanne Buiterhelped to improve this work considerably with their comments on themanuscript. A first draft of the manuscript was prepared during asabbatical leave of CF at the University of Montpellier II in 2007.

References

Becker, T.W., Faccenna, C., O'Connell, R.J., Giardini, D., 1999. The development of slabs inthe upper mantle: insights from numerical and laboratory experiments. J. Geophys.Res. 104 (B7), 15207–15226.

Bellahsen, N., Faccenna, C., Funiciello, F., 2005. Dynamics of subduction and platemotion in laboratory experiments: insights into the “plate tectonics” behavior ofthe Earth. J. Geophys. Res. 110, B01401. doi:10.1029/02044JB002999.

Bijwaard, H., Spakman, W., Engdahl, E.R., 1998. Closing the gap between regional andglobal travel time tomography. J. Geophys. Res. 103, 30055–30078.

Billen, M.I., Hirth, G., 2007. Rheologic controls on slab dynamics. Geochem. Geophys.Geosyst. 8, Q08012. doi:10.1029/2007GC001597.

Billen, M.I., 2008. Modeling the dynamics of subducting slabs. Annu. Rev. Earth Planet.Sci. 36, 325–356.

Buiter, S.J.H., Govers, R., Wortel, M.J.R., 2002. Two-dimensional simulations of surfacedeformation caused by slab detachment. Tectonophysics 354, 195–210.

Buffett, B.A., Rowley, D.B., 2006. Plate bending at subduction zones: consequences forthe direction of plate motions. Earth Planet. Sci. Lett. 245 (1–2), 359–364.

Bunge, H.P., Richards, M.A., Baumgardner, J.R., 1997. A sensitivity study of three-dimensional spherical mantle convection at 108 Rayleigh number: effects of depth-dependent viscosity, heating mode, and endothermic phase change. J. Geophys.Res. 102, 11991–12007.

Byerlee, J.D., 1968. Brittle–ductile transition in rocks. J. Geophys. Res. 73, 4741–4750.Cande, S.C., Raymond, C.A., Stock, J., Haxby, W.F., 1995. Geophysics of the Pitman

fracture zone and Pacific–Antarctic plate motions during the Cenozoic. Science 270,947–953.

Capitanio, F.A., Morra, G., Goes, S., 2007. Dynamic models of downgoing plate buoyancydriven subduction: subduction motions and energy dissipation. Earth Planet. Sci.Lett. 262, 284–297.

Carlson, R.L., Hilde, T.W.C., Uyeda, S., 1983. The driving mechanism of plate tectonics:relation to age of the lithosphere at trenches. Geophys. Res. Lett. 10, 297–300.

Carlson, R.L., Melia, P.J., 1984. Subduction hinge migration. Tectonophysics 102,399–411.

Christensen, U.R., 1995. Effects of phase transition on mantle convection. Annu. Rev.Earth Planet. Sci. 23, 65–87.

Christensen, U.R., 1996. The influence of trench migration on slab penetration into thelower mantle. Earth Planet. Sci. Lett. 140, 27–39.

Cıžková, H., van Hunen, J., van den Berg, A.P., Vlaar, N.J., 2002. The influence ofrheological weakening and yield stress on the interaction of slabs with the 670 kmdiscontinuity. Earth Planet. Sci. Lett. 199, 447–457.

Cıžková, H., van Hunen, J., van den Berg, A.P., 2007. Stress distribution withinsubducting slabs and their deformation in the transition zone. Phys. Earth Planet.Inter. 161 (3–4), 202–214.

Conrad, C.P., Hager, B.H., 1999. Effects of plate bending and fault strength at subductionzones on plate dynamics. J. Geophys. Res. 104, 17551–17571.

Cox, A., Engebretson, D., 1985. Change in motion of Pacific plate at 5 Myr BP. Nature 313(7), 472–474.

Davies, G.F., 1995. Penetration of plates and plumes through themantle transition zone.Earth Planet. Sci. Lett. 133, 507–516.

Deschamps, A., Lallemand, S., 2002. The West Philippine Basin: an Eocene to earlyOligocene back arc basin opened between two opposed subduction zones. J. Geophys.Res. 107 (B12), 2322. doi:10.1029/2001JB001706.

Dewey, J.F., 1980. Episodicity, sequence and style at convercent plate boundaries. TheContinental Crust and its Mineral Deposits: In: S.D.W. (Ed.), Geological Associationof Canada Special Paper, pp. 553–573.

Di Giuseppe, E., van Hunen, J., Funiciello, F., Faccenna, C., Giardini, D., 2008. Plate strengthcontrols trenchmotion: insights fromnumericalmodels. Geochem. Geophys. Geosyst.9, Q02014.

Di Giuseppe, E., van Hunen, J., Funiciello, F., Faccenna, C., Giardini, D., 2009. On therelation between trench migration, seafloor age and the strength of the subductinglithosphere. Lithosphere 1, 121–128.

Doglioni, C., Carminati, E., Cuffaro, M., Scrocca, D., 2007. Subduction kinematics anddynamic constraints. Earth Sci. Rev. 83, 125–175.

Enns, A., Becker, T.W., Schmeling, H., 2005. The dynamics of subduction and trenchmigration for viscosity stratification. Geophys. J. Int. 160, 761–769.

Faccenna, C., Funiciello, F., Giardini, D., Lucente, P., 2001. Episodic back-arc extensionduring restricted mantle convection in the central Mediterranean. Earth Planet. Sci.Lett. 187 (1–2), 105–116. doi:10.1016/S0012-821X(01)00280-1.

Faccenna, C., Heuret, A., Funiciello, F., Lallemand, S., Becker, W.T., 2007. Predicting trenchandplatemotion fromthedynamicsof a strong slab. Earth Planet. Sci. Lett. 257, 29–36.

Fukao, Y., Widiyantoro, S., Obayashi, M., 2001. Stagnant slabs in the upper and lowermantle transition region. Rev. Geophys. 39, 291–323.

Funiciello, F., Faccenna, C., Heuret, A., Lallemand, S., Di Giuseppe, E., Becker, T.W., 2008.Trench migration, net rotation and slab–mantle coupling. Earth Planet. Sci. Lett.271, 233–240.

Funiciello, F., Faccenna, C., Giardini, D., 2004. Role of lateral mantle flow in the evolutionof subduction system: insights from 3-D laboratory experiments. Geophys. J. Int.157, 1393–1406.

Funiciello, F., Faccenna, C., Giardini, D., Regenauer-Lieb, K., 2003. Dynamics ofretreating slabs (part 2): insights from 3D laboratory experiments. J. Geophys.Res. 108 (B4).

Goes, S., Capitanio, F.A., Morra, G., 2008. Evidence of lower-mantle slab penetrationphases in plate motions. Nature 451. doi:10.1038/nature06691.

Goes, S., van der Lee, S., 2002. Thermal structure of the North American uppermostmantleinferred from seismic tomography. J. Geophys. Res. 107 (B3), 2050. doi:10.1029/2000JB000049.

Gordon, R.G., Jurdy, D.M., 1986. Cenozoic global plate motion. J. Geophys. 91, 12,389–12,406.Griffiths, R.W., Hackney, R.I., Vanderhilst, R.D., 1995. A laboratory investigation of

effects of trench migration on the descent of subducted slabs. Earth Planet. Sci. Lett.133 (1–2), 1–17.

Gripp, A.E., Gordon, R.G., 2002. Young tracks of hotspots and current plate velocities.Geophys. J. Int. 150, 321–361.

Guillou-Frottier, L., Buttles, J., Olson, P., 1995. Laboratory experiments on structure ofsubducted lithosphere. Earth Planet. Sci. Lett. 133, 19–34.

Gurnis, M., Hager, B.H., 1988. Controls of the structure of subducted slabs. Nature 335,317–321.

Hager, B.H., 1984. Subducted slabs and the geoid: constraints on mantle rheology andflow. J. Geophys. Res. 89, 6003–6015.

Hall, R., 2002. Cenozoic geological and plate tectonic evolution of SE Asia and the SWPacific: computer-based reconstructions, model and animations. J. Asian Earth Sci.20, 353–431.

Hall, R., Ali, J.R., Anderson, C.D., Baker, S.J., 1995a. Origin and motion history of thePhilippine Sea Plate. Tectonophysics 251, 229–250.

Hall, R., Fuller, M., Ali, J.R., Anderson, C.D., 1995b. The Philippine Sea Plate: magnetismand reconstructions. In: Taylor, B., Netland, J. (Eds.), Active Margins and MarginalBasins of the Western Pacific. Geophys. Monogr. Ser., vol. 88. AGU,Washington, D. C.,pp. 371–404.

Harbert, W., Cox, A., 1989. Late Neogene motion of the Pacific plate. J. Geophys. Res. 94,3052–3064.

Heuret, A., Lallemand, S., 2005. Plate motions, slab dynamics and back-arc deformation.Phys. Earth Planet. Inter. 149, 31–51.

Hilde, T.W.C., Lee, C.S., 1984. Origin and evolution of the West Philippine basin: a newinterpretation. Tectonophysics 102, 85–104.

Husson, L., Conrad, C.P., Faccenna, C., 2008. Tethyan closure, Andean Orogeny andWestward drift of the Pacific Basin. Earth Planet. Sci. Lett. 271, 303–310.

Iaffaldano, G., Bunge, H.P., Dixon, T.H., 2006. Feedback between mountain belt growthand plate convergence. Geology 34 (10), 893–896.

Ita, J., King, S.D., 1998. The influence of thermodynamic formulation on simulations ofsubduction zone geometry and history. Geophys. Res. Lett. 25, 1463–1466.

Jarrard, R.D., 1986. Relations among subduction parameters. Rev. Geophys. 24, 217–284.Karato, S., Riedel, M.R., Yuen, D.A., 2001. Rheological structure and deformation of

subducted slabs in the mantle transition zone: implications for mantle circulationand deep earthquakes. Phys. Earth Planet. Int. 3994, 1–26.

Kaus, B., Becker, T.W., 2006. Effects of elasticity on the Rayleigh–Taylor instability:implications for large-scale geodynamics. Geophys. J. Int. doi:10.1111/j.1365-246X.2006.03201.x.

Kaus, B., Becker, T.W., 2008. A numerical study on the effects of surface boundarycondition and rheology on slab dynamics. Boll. Geofis. Teor. Appl. 49 (2), 177–181.

Kincaid, C., Olson, P., 1987. An experimental study of subduction and slab migration.J. Geophys. Res. 92, 13832–13840.

Korenaga, M., 2007. Effective thermal expansivity of Maxwellian oceanic lithosphere.Earth Planet. Sci. Lett. 257, 343–349.

King, S.D., 2001. Subduction zones: observations and geodynamic models. Phys. EarthPlanet. Inter. 127, 9–24.

Lallemand, S., Jolivet, L., 1986. Japan Sea: a pull-apart basin? Earth Planet. Sci. Lett. 76,375–389.

Lallemand, S., Okada, H., Otsuka, K., Labeyrie, L., 1985. Tectonique en compression sur lamarge est de la mer du Japon: mise en évidence de chevauchements à vergenceorientale. C. R. Acad. Sci. Paris 301 (3), 201–206 II.

Lallemand, S., Heuret, A., Faccenna, C., Funiciello, F., 2008. Subduction dynamics asrevealed by trench migration. Tectonics 27, Tc3014. doi:10.1029/2007tc002212.

Le Pichon, X., Huchon, P., 1987. Central Japan triple junction revisited. Tectonics 6, 35–45.Martinez, F., Fryer, P., Baker, N.A., Yamazaki, T., 1995. Evolution of back-arc rifting:

Mariana Trough, 20–24N. J. Geophys. Res. 100, 3807–3827.Mazzotti, S., Henry, P., Le Pichon, X., 2001. Transient and permanent deformation of

central Japan estimated by GPS—2. Strain partitioning and arc–arc collision. EarthPlanet. Sci. Lett. 184, 455–469.

Miller, M.S., Kennett, B.L.N., Toy, V.G., 2006. Spatial and temporal evolution of thesubducting Pacific plate structure along the western Pacific margin. J. Geophys. Res.111, B02401. doi:10.1029/2005JB003705.

Miller, M.S., Kennett, B.L.N., Lister, G.S, 2004. Imaging changes in morphology,geometry, and physical properties of the subducting Pacific plate along the Izu-Bonin-Mariana arc. Earth Planet. Sci. Lett. 224 (3-4), 363–370.

Mitrovica, J.X., 1996. Haskell [1935] revisited. J. Geophys. Res. 101, 555–569.Molnar, P., Atwater, T., 1978. Interarc spreading and Cordilleran tectonics as alternates

related to age of subducted oceanic lithosphere. Earth Planet. Sci. Lett. 41, 330–340.

397C. Faccenna et al. / Earth and Planetary Science Letters 288 (2009) 386–398

Author's personal copy

Moresi, L., Gurnis, M., 1996. Constrains on the lateral strength of slabs from three-dimensional dynamic flow models. Earth Planet. Sci. Lett. 138, 15–28.

Müller, R.D., Royer, J.-Y., Lawver, L.A., 1993. Revised plate motions relative to thehotspots from combined Atlantic and Indian Ocean hotspot tracks. Geology 16,275–278.

O'Neill, C., Müller, D., Steinberger, B., 2005. On the uncertainties in hot spotreconstructions and the significance of moving hot spot reference frames.Geochem. Geophys. Geosyst. 6, Q04003. doi:10.1029/2004GC000784.

Okino, K., Kasuga, S., Ohara, Y., 1998. A new scenario of the Parace-Vela Basin genesis.Mar. Geophys. Res. 20, 21–40.

Okino, K., Shimakawa, Y., Nagaoka, S., 1994. Evolution of the Shikoku basin. J. Geomagn.Geoelectr. 46, 463–479.

Pacanovsky, K.M., Davis, D.M., Richardson, R.M., Coblentz, D.D., 1999. Intraplate stressesand plate driving forces in the Philippine Sea plate. J. Geophys. Res. 104 (B1),1095–1110.

Pollitz, F.F., 1986. Pliocene change in Pacific-plate motion. Nature 320, 738–741.Rangin, C., Jolivet, L., Pubellier, M., 1990. A simple model for the tectonic evolution of

southeast Asia and Indonesia region for the past 43 m.y. Bull. Soc. Géol. Fr. 8 (6),889–905.

Ribe, Bending mechanics and mode selection in free subduction: a thin-sheet analysis.Geophysical Journal International, in press.

Richards, M.A., Engebretson, D.C., 1992. Large-scale mantle convection and the historyof subduction. Nature 355, 437–440.

Ringwood, A.E., Irfune, T., 1988. Nature of the 650-km seismic discontinuity. Nature331, 131–136.

Ritzwoller, M.H., Shapiro, N.M., Zhong, S., 2004. Cooling history of the Pacificlithosphere. Earth Planet. Sci. Lett. 226, 69–84.

Savostin, L., Zonenshain, L., Baranov, B., 1983. Geology and plate tectonics of the Sea ofOkhotsk. In: Hilde, T.W.C., Uyeda, S. (Eds.), Geodynamics of the Western Pacific-Indonesian Region. : Geodynamics series, vol. 11. AGU, pp. 189–221.

Scotese, C.R., Gahagan, L.M., Larson, R.L., 1988. Plate tectonic reconstructions of theCretaceous and Cenozoic ocean basins. Tectonophysics 155, 27–48.

Schellart, W.P., 2005. Influence of the subducting plate velocity on the geometry of theslab and migration of the subduction hinge. Earth Planet. Sci. Lett. 231, 197–219.

Schellart, W.P., Freeman, J., Stegman, D.R., Moresi, L., May, D., 2007. Evolution anddiversity of subduction zones controlled by slab width. Nature 446, 308–311.

Schmeling, H., Babeyko, A., Enns, A., Faccenna, C., Funiciello, F., Gerya, T., Golabek, G.,Grigull, S., Morra, G., van Hunen, J., 2008. A benchmark comparison of subductionmodels. Phys. Earth Planet. Inter. 171, 198–223.

Sdrolias, M., Müller, R.D., 2006. Controls on back-arc basin formation. Geochem.Geophys. Geosyst. 7, Q04016. doi:10.1029/2005GC001090.

Sdrolias, M., Roest, W.R., Müller, R.D., 2004. An expression of Philippine Sea platerotation: the Parece Vela and Shikoku Basins. Tectonophysics 394, 69–86.doi:10.1016/j.tecto.2004.07.061.

Seno, T., Maruyama, S., 1984. Paleogeographic reconstruction and origin of thePhilippine Sea. Tectonophysics 102, 53–84.

Seno, T., Stein, S., Gripp, A.E., 1993. A model for the motion of the Philippine Sea Plateconsistent with NUVEL-1 and geological data. J. Geophys. Res. 98, 17941–17948.

Spence, W., 1987. Slab pull and the seismotectonics of subducting lithosphere. Rev.Geophys. 25 (1), 55–69.

Steinberger, B., Sutherland, R., O'Connell, R.J., 2004. Prediction of Hawaiian-Emperorseamount locations from a revised model of global plate motion and mantle flow.Nature 430, 167–173.

Stegman, D.R., Freeman, J., Schellart, W.P., Moresi, L., May, D., 2006. Influence of trenchwidth on subduction hinge retreat rates in 3-D models of slab rollback. Geochem.Geophys. Geosyst. 7, Q03012. doi:10.1029/2005GC001056.

Stegman, D.R., Farrington, R., Capitanio, F.A. and Schellart, W.P., in press. A regimediagram for subduction styles from 3-D numerical models of free subduction.Tectonophysics.

Tackley, P.J., Stevenson, D.J., Glatzmaier, G.A., Schubert, G., 1993. Effects of anendothermic phase-transition at 670 km depth in a spherical model of convectionin the Earth's mantle. Nature 361, 699–704.

Tamaki, K., Honza, E., 1984. Incipient subduction and obduction along the easternmargin of the Japan Sea. Tectonophysics 119, 381–406.

Turcotte, D.L., Schubert, G., 1982. Geodynamics Application of Continuum Physics toGeological Problems. John Wiley and Sons, New York, p. 450.

Uyeda, S., Kanamori, H., 1979. Back-arc opening and the mode of subduction.J. Geophys. Res. 84, 1049–1061.

van der Hilst, R.D., Engdahl, E.R., Spakman, W., Nolet, G., 1991. Tomographic imaging ofsubducted lithosphere below northwest Pacific island arcs. Nature 353 (6339),37–43.

van der Hilst, R.D., Seno, T., 1993. Effects of relative plate motion on the deep structureand penetration depth of slabs below the Izu–Bonin and Mariana island arcs. EarthPlanet. Sci. Lett. 120, 375–407.

van der Voo, R., Spackman, W., Bijwaard, H., 1999. Mesozoic subducted slabs underSiberia. Nature 397, 246–249.

van Hunen, J., van den Berg, A.P., Vlaar, N.J., 2002. On the role of subducting oceanicplateaus in the development of shallow flat subduction. Tectonophysics 352,317–333. doi:10.1016/S0040-1951(02)00263-9.

van Hunen, J., van den Berg, A.P., 2008. Plate tectonics on the early Earth: limitationsimposed by strength and buoyancy of subducted lithosphere. Lithos 103, 217–235.

Wortel, M.J.R., 1982. Seismicity and rheology of subducted slabs. Nature 296, 553–556.Wessel, P., Kroenke, L., 2000. Ontong java Plateau and the late Neogene change in

Pacific motion. J. Geophys. Res. 105, 28255–28277.Zhong, S., Gurnis, M., 1995. Mantle convection with plates and mobile, faulted plate

margins. Science 267, 838–843.Zhong, S., Zuber, M.T., Moresi, L., Gurnis, M., 2000. Role of temperature-dependent

viscosity and surface plates in spherical shell models of mantle convection.J. Geophys. Res. 105, 11063–11082.

398 C. Faccenna et al. / Earth and Planetary Science Letters 288 (2009) 386–398