Why cold slabs stagnate in the transition zone

Similar documents
Why does the Nazca plate slow down since the Neogene? Supplemental Information

Durham Research Online

Interpreting Geophysical Data for Mantle Dynamics. Wendy Panero University of Michigan

The upper mantle is the source of almost all magmas. It contains major

Hydration of Olivine and. Earth s Deep Water Cycle

α phase In the lower mantle, dominant mineralogy is perovskite [(Mg,Fe)SiO 3 ] The pyrolite mantle consists of: 60% olivine and 40% pyroxene.

WATER IN THE EARTH S MANTLE:

Time-varying subduction and rollback velocities in slab stagnation and buckling

Absence of density crossover between basalt and peridotite in the cold slabs passing through 660 km discontinuity

Rheology of the Mantle and Plates (part 1): Deformation mechanisms and flow rules of mantle minerals

Topography of the 660-km discontinuity beneath northeast China: Implications for a retrograde motion of the subducting Pacific slab

Viscosity in transition zone and lower mantle: Implications for slab penetration

Effect of water on the spinel-postspinel transformation in Mg 2 SiO 4

Benchmarks for subduction zone models Subduction zone workshop, University of Michigan, July 2003

Laboratory Electrical Conductivity Measurement of Mantle Minerals

Stress field in the subducting lithosphere and comparison with deep earthquakes in Tonga

Numerical Simulation of the Thermal Convection and Subduction Process in the Mantle

The Effect of H 2 O on the 410-km Seismic Discontinuity

Defects, Diffusion, Deformation and Thermal Conductivity in the Lower Mantle and D

SUPPLEMENTARY INFORMATION

GSA DATA REPOSITORY

GEOL540: The Mantle System

Textures in experimentally deformed olivine aggregates: the effects of added water and melt.

Modeling the Dynamics of Subducting Slabs

Equation of state of (Mg 0.8,Fe 0.2 ) 2 SiO 4 ringwoodite from synchrotron X-ray diffraction up to 20 GPa and 1700 K

Weidner and Wang (1999) Phase transformations - implications for mantle structure

A BADER S TOPOLOGICAL APPROACH FOR THE CHARACTERIZATION OF PRESSURE INDUCED PHASE TRANSITIONS

Rheological structure and deformation of subducted slabs in the mantle transition zone: implications for mantle circulation and deep earthquakes

Crust : wet quartzite Arc root : dry olivine mantle = 2840 kg/m km = 3300 kg/m km (Arc root thickness) 280 km (Arc width)

Constraints on Mantle Structure from Surface Observables

Olivine-wadsleyite transition in the system (Mg,Fe) 2 SiO 4

Development of Anisotropic Structure by Solid-State Convection in the Earth s Lower Mantle

Influence of the Ringwoodite-Perovskite transition on mantle convection in spherical geometry as a function of Clapeyron slope and Rayleigh number

SUPPLEMENTARY INFORMATION

Post-perovskite 1. Galley Proofs

DETAILS ABOUT THE TECHNIQUE. We use a global mantle convection model (Bunge et al., 1997) in conjunction with a

Subduction II Fundamentals of Mantle Dynamics

Zhicheng Jing 3,, Yanbin Wang 3. Laboratoire magmas et volcans, CNRS, UMR 6524, IRD, Université Blaise Pascal, Clermont- Ferrand, France

Earth and Planetary Science Letters

Modeling the Thermal-Mechanical Behavior of Mid-Ocean Ridge Transform Faults

2.01 Overview Mineral Physics: Past, Present, and Future

Impact-driven subduction on the Hadean Earth

Seismology and Deep Mantle Temperature Structure. Thorne Lay

Effects of compositional and rheological stratifications on smallscale

Thermo-chemical structure, dynamics and evolution of the deep mantle: spherical convection calculations

Lateral, radial, and temporal variations in upper mantle viscosity and rheology under Scandinavia

Mars Internal Structure, Activity, and Composition. Tilman Spohn DLR Institute of Planetary Research, Berlin

Structural, vibrational and thermodynamic properties of Mg 2 SiO 4 and MgSiO 3 minerals from first-principles simulations

Zhu Mao. Department of Geological Sciences The University of Texas at Austin Austin, TX

SIO 224. Thermodynamics of phase transitions

Sorosilicates, Colors in Minerals (cont), and Deep Earth Minerals. ESS212 January 20, 2006

Supplement of The influence of upper-plate advance and erosion on overriding plate deformation in orogen syntaxes

Geodynamics of MARGINS

Slab stress and strain rate as constraints on global mantle flow

Anomalously thin transition zone and apparently isotropic upper mantle beneath Bermuda: Evidence for upwelling

Three-dimensional numerical simulations of thermo-chemical multiphase convection in Earth s mantle Takashi Nakagawa a, Paul J.

Beall et al., 2018, Formation of cratonic lithosphere during the initiation of plate tectonics: Geology, /g

Whole Mantle Convection

Time-dependent interaction between subduction dynamics and phase transition kinetics

The in uence of rheological weakening and yield stress on the interaction of slabs with the 670 km discontinuity

The anisotropic and rheological structure of the oceanic upper mantle from a simple model of plate shear

Modeling the interior dynamics of terrestrial planets

EMMR25 Mineralogy: Ol + opx + chlorite + cpx + amphibole + serpentine + opaque

Petrogenetic grid in the system MgO-SiO 2 -H 2 O up to 30 GPa, 1600 C: Applications to hydrous peridotite subducting into the Earth s deep interior

Pressure Volume Temperature Equation of State

Studies of Arc Volcanism and Mantle Behavior in Subduction Zones

Supplementary information on the West African margin

Earth and Planetary Science Letters

Localization of dislocation creep in the lower mantle: implications for the origin of seismic anisotropy

Influence of phase transformations on lateral heterogeneity and dynamics in Earth's mantle

The importance of the South-American plate motion and the Nazca Ridge subduction on flat subduction below South Peru

EART162: PLANETARY INTERIORS

Nature of heterogeneity of the upper mantle beneath the northern Philippine Sea as inferred from attenuation and velocity tomography

Rheology: What is it?

Water-induced convection in the Earth s mantle transition zone

Ch 6: Internal Constitution of the Earth

Inferring upper-mantle temperatures from seismic velocities

What is the LAB Dynamically: Lithosphere and Asthenosphere Rheology from Post-loading Deformation

Stability of hydrous silicate at high pressures and water transport to the deep lower mantle

The validity and usefulness of thermodynamic models commonly used

Subsolidus and melting experiments of a K-rich basaltic composition to 27 GPa: Implication for the behavior of potassium in the mantle

Raman spectroscopy at high pressure and temperature for the study of Earth's mantle and planetary minerals

Phase Transitions in the Lowermost Mantle

Trench motion-controlled slab morphology and stress variations: Implications for the isolated 2015 Bonin Islands deep earthquake

The influence of phase boundary deflection on velocity anomalies of stagnant slabs in the transition zone

Deep and persistent melt layer in the Archaean mantle

C3.4.1 Vertical (radial) variations in mantle structure

Physics of the Earth and Planetary Interiors

Garnet yield strength at high pressures and implications for upper mantle and transition zone rheology

Constraining the composition and thermal state of the mantle beneath Europe from inversion of long-period electromagnetic sounding data

Pressure dependence of electrical conductivity of (Mg,Fe)SiO 3 ilmenite

Computation of seismic profiles from mineral physics: the importance of the non-olivine components for explaining the 660 km depth discontinuity

Rheology. What is rheology? From the root work rheo- Current: flow. Greek: rhein, to flow (river) Like rheostat flow of current

Thermal-Mechanical Behavior of Oceanic Transform Faults

Chemical Composition of the Lower Mantle: Constraints from Elasticity. Motohiko Murakami Tohoku University

Journal of Applied Mathematics and Computation (JAMC), 2018, 2(7),

phase H in deep subducted oceanic crust

Supplementary Material for. Mantle induced subsidence and compression in SE Asia

Mantle Transition Zone Thickness in the Central South-American Subduction Zone

Experimental constraints on the strength of the lithospheric mantle

Transcription:

GSA Data Repository 2015085 Model 1 Model 2 Model 3 Model 4 Model 5 Model 6 Why cold slabs stagnate in the transition zone Scott D. King 1,2, Daniel J. Frost 2, and David C. Rubie 2 1 Department of Geosciences, Virginia Tech, Blacksburg, Virginia 24061, USA 2 Bayerisches Geoinstitut, Universität Bayreuth, D-95440 Bayreuth, Germany Supplemental Material This document contains supplemental material documenting the equilibrium formulation for the pyroxene-majorite transformation implemented in the convection code, a technical discussion of the transition temperature between equilibrium and nonequilibrium phase transformation based on the slow diffusion laboratory results of van Mierlo et al. (2013) and documentation of the composite rheology used in the convection modeling. We also include the phase diagram for the olivine-wadsleyite-ringwoodite transformation adapted from Frost (2008). Equilibrium Pyroxene-Majorite Transformation For the pyroxene to garnet-majorite transformation, we use a parameterized equilibrium thermodynamic formulation that assumes a 50:50 mixture of pyrope and pyroxene as a starting point using parameters from Stixrude and Lithgow-Bertelloni (2011). The formulation is as follows. For the Gibbs free energy, G, we have G = 126771+3.07T 7141P (1)

where P and T are pressure (in GPa) and temperature (in K). The mole fraction of majorite, Y, is then given by G Y exp (2) 8.314T and finally density (in kg/m 3 ) is given by px_ gt 40000 0.5a (0.5 Y)b Yc (3) where a = 11.611 0.0533P (4) b = 12.651 0.0724P (5) c = 11.723 0.0568P. (6) This is shown in Fig. S2 where we have scaled the density by T. To assess the effect of slow diffusion of majoritic garnet on the phase transformation at low temperatures, we applied a hyperbolic tangent function to Equation (3) as follows 40000 px_ gt 0.5a (0.5 Y)b Yc _tanh T T trans 1 (7) 100 where T trans is the transition temperature below which slow-diffusion occurs. Transition Temperature In order to model the effects of pyroxene metastability, we use a critical transition temperature. Below this temperature, pyroxene is considered to remain metastable up to a pressure of 18 GPa whereas above the transition temperature it is considered to transform rapidly to majoritic garnet. For comparison, in the case of the olivine-spinel transformation, the transition temperature has been estimated to be 900-950 K (Rubie and

Ross, 1994). For the majorite-forming reaction, relevant diffusion coefficients can be determined as a function of temperature from the activation energy and the preexponential factor determined by van Mierlo et al. (2013). Diffusion distances are then calculated assuming a timescale of 3 My for the descent of a slab through the transition zone and are compared with the expected grain size of 5-10 mm which gives the diffusion distance required for complete transformation. Diffusion distances thus calculated are 60 µm at 1000 K, 0.2 mm at 1200 K, 0.5 mm at 1300 K and 1.2 mm at 1400 K. Based on these results, we assume for most calculations that the transition is 1200 K because this results in an extent of transformation that is finite but small compared with the expected grain size. However, this value is on the conservative side and a value of 1300 K would also be reasonable. Using the latter would produce even stronger buoyancy/stagnation effects than are documented for 1200 K. Composite Rheology Following Hirth and Kohlstedt (2003) we define the effective viscosity as, 1 1 n d A p r n fh2 O n exp( ) 1 n 1 exp (E * PV * ) (8) nrt where A is a pre-exponential factor, n is the stress exponent, d is the grain-size, p is the grain-size exponent, fh 2 O is the water fugacity, r is the water fugacity exponent, φ is the melt fraction, α is a constant, is the strain-rate, E * is the activation energy, V * is the activation volume, R is the gas constant. This form of the equation assumes that the stress is given in MPa, grain size in μm and fh 2 O or C OH in H/ 10 6 Si. We do not consider melt and thus φ=0.

For diffusion creep, also called a linear rheology or Newtonian rheology, the stress exponent is one (i.e., n=1) and thus, the strain-rate-dependent term is always 1.0. The diffusion creep mechanism is a function of grain size (Hirth and Kohlstedt, 2003). Following Billen and Hirth (2007) we modify the activation volume for the lower mantle to be 1.5 10 6 m 3 /mole and modify the pre-exponential term so that there is a factor of 30 increase in lower mantle viscosity. Dislocation creep, or power-law or non-newtonian rheology is given by the same equation above (eqn 8), except that now the stress exponent is greater than 1 (i.e., n=3.5) and is independent of grain size (p=0). Following Billen and Hirth (2007) we define the composite rheology, comp, as comp 1.0 1.0 dif 1.0 disl, (9) where dif is diffusion creep rheology and disl is dislocation creep rheology. This weighted average is only used in the upper 400 km, as below this depth, the mantle is seismically isotropic and this suggests that diffusion creep dominates (King, 2007). For numerical stability we truncate the viscosity so that the maximum viscosity is no more than 10 5 times the background viscosity. This has been shown to be large enough viscosity contrast that truncation of larger viscosities does not affect the resulting flow (King, 2009).

References Akaogi, M.. and Ito, E., 1993, Heat capacity of MgSiO4 perovskite: Geophysical Research Letters, v. 20, p. 105 108. Bina, C. R., and Helffich, G., 1994, Phase transition Clapeyron slopes and transition zone seismic discontinuity topography: Journal of Geophysical Research, v. 99, p. 15,853 15,860. Billen, M. I., 2008, Modeling the dynamics of subducting slabs: Annual Reviews of Earth and Planetary Science, v 36, p. 325 356. Billen, M. I., and Hirth, G., 2007, Rheologic controls on slab dynamics: Geochemistry Geophysics Geosystems, v. 8, Q08012, doi:10.1029/2007gc001597. Fei, Y., Van Orman, J. Li, J., van Westrenen, W., Sanloup, C., et al., 2004, Experimentally determined postspinel transformation boundary in Mg2SiO4 using MgO as an internal pressure standard and its geophysical implications: Journal of Geophysical Research, v. 109, B02305. Frost, D.J., 2008, The upper mantle and transition zone: Elements, v. 4, p. 171 176, doi:10.2113/gselements.4.3.171. Hirth, G., and Kohlstedt, D., 2003, Rheology of the upper mantle and the mantle wedge: A view from the experimentalists, in Inside the Subduction Factory, Geophys. Monogr. Ser. 138 edited by J. Eiler, pp. 83 105 AGU, Washington, D. C. Irifune, T., Nishiyama, N., Kuroda, K., Inoue, T., Isshiki, T, M., et al., 1998, The postspinel phase boundary in Mg2 SiO4 determined by in situ X-ray diffraction: Science, v. 279, p. 1698 1700.

Ito, E., and Takahashi, E., 1989, Postspinel transformations in the system Mg2SiO4- Fe2Si04 and some geophysical implications: Journal of Geophysical Research, v. 94, p. 10,637 10,646. Katsura, T., Yamada, H., Shinmei, T., Kubo, A., Ono, S., et al., 2003, Post-spinel transition in Mg2SiO4 determined by high P-T in situ X-ray diffractometry: Physics of the Earth and Planetary Interiors, v. 136, p. 11 24. King, S.D., 2009, On topography and geoid from 2-D stagnant lid convection calculations: Geochemistry, Geophysics, Geosystems, v. 10, Q03002, doi:10.1029/2008gc002250. King, S.D., 2007, Mantle downwellings and the fate of subducting slabs: Constraints from seismology, geoid, topography, geochemistry, and petrology. in Treatise on Geophysics, Volume 7, Mantle Dynamics, pp. 325 370. Litasov, K.D., Ohtani, E., Sano, A., and Suzuki, A., 2005a, Wet subduction versus cold subduction: Geophysical Research Letters, v.32, L13312. Litasov, K.D., Ohtani, E., Sano, A., Suzuki, A., and Funakoshi K., 2005b, Earth and Planetary Science Letters, v. 238, p. 311 328. Rubie, D.C., and Ross, C.R., 1994, II, Kinetics of the olivine-spinel transformation in subducting lithosphere: Experimental constraints and implications for deep slab processes: Physics of the Earth and Planetary Interiors, v. 86, p. 223 241. Stixrude, L., and Lithgow-Bertelloni, C., 2011, Thermodynamics of mantle minerals II. Phase equilibria: Geophysical Journal International, v. 184, p. 1180 1213.

Figure S1: Phase diagram for the olivine mantle component as implemented in our subduction calculations. Light green is olivine, dark green is the transition zone (wadsleyite and ringwoodite) and orange is the bridgmanite plus ferropericlase component.

Figure S2: Relative density of the pyroxene-garnet phase transformation scaled by ϱαδt using equations (1)-(6). The non-equilibrium version is created by applying a hyperbolic tangent function to equation (7) as described in the text.

Model Earth Parameters Value reference density 3.3 10 3 kg/m 3 coefficient of thermal expansion 2.0 10 5 K 1 surface gravity 10 m/s 2 surface temperature 273 K convective temperature drop 1875 K depth of the mantle 2.890 10 6 m thermal diffusivity 10 6 m 2 /s reference viscosity 10 20 Pa s Rayleigh number 3.0 10 8 Table 1: Model Parameters

Clapeyron Slope Reference MPa/K -2.8 Ito and Takahashi (1989) -3.0 Akaogi and Ito (1993) -2.9 Irifune et al. (1998) -2.0 Bina and Helffich (1994) -0.4 to -2.0 Katsura et al. (2003) -1.3 Fei et al. (2004) -0.5 to -0.8 (dry) Litasov et al. (2005a) -2.0 (hydrus) Litasov et al. (2005b) Table 2: Published values of the Clapeyron Slope for the ringwoodite to perovskite plus ferropericlase transformaion. Study A n p E * V * (kj/mole) 10 6 m 3 /mole HK03 dry 1.5 10 9 1 3 375±50 2-10 HK03 wet 2.5 10 7 1 3 375±50 0-20 BH07 1.0 1 3 335 4 this study 1.0 1 3 335 4 Table 3: Diffusion Creep Model Parameters Study A n p E * V * (kj/mole) 10 _ 6 m 3 /mole HK03 dry 1.1 10 5 3.5±0.3 0 520±40 0-20 HK03 wet 1.6 10 3 3.5±0.3 0 480±40 22±11 BH07 90 10 21 3.5 0 480 11 this study 90 10 21 3.5 0 480 11 Table 4: Dislocation Creep Model Parameters