Skip to main content

The Subductability of Continental Lithosphere: The Before and After Story

  • Chapter
  • First Online:
Arc-Continent Collision

Part of the book series: Frontiers in Earth Sciences ((FRONTIERS))

Abstract

The temporal evolution of internal forces in a collision environment controls first-order characteristics such as convergence rate, slab dip, subduction stall, and slab breakoff, amongst others. Foremost among these forces are the positive buoyancy provided by the subduction of felsic continental material and the negative buoyancy associated with the slab. In this work we use fully dynamic thermomechanical models coupled with thermodynamic/petrological formalisms to study the evolution of these forces during a continent–arc/microcontinent collision and their influence on the large-scale dynamics of the system. Two distinctive features of our models that allow a self-consistent assessment of collision dynamics are: (1) the use of a new thermodynamic database valid up to ~25–30 GPa that includes most of the major phases relevant to continental subduction, and (2) a fully dynamic approach in which no velocities are imposed to either force or stop subduction. The former allows realistic computations of the buoyancy forces driving the system as a function of P-T-composition. The latter assures that computed velocities emerge self-consistently in our simulations in response to the balance between internal forces in our numerical domain.

The main results from our experiments can be summarized as follows. (1) The delamination of the lithospheric mantle after a short episode of continental subduction is a viable scenario to end continental subduction; the associated evolution of convergence is comparable to those proposed for real collision setting. (2) We corroborate previous results showing that the main control on the dynamics and final configuration (type of slab breakoff) of the collision is the rheology and composition of the continental crust; strong mafic crusts favor deep subduction and recycling of significant volumes of continental material, while soft felsic crusts preclude them. (3) Subducted continental crust remains buoyant with respect to the surrounding mantle down to depths of ~250–300 km, thus allowing exhumation of deeply subducted crust as long as a detachment from the slab occurs. (4) Realistic compositional stratifications in the continental lithospheric mantle exert only a modest influence on the overall evolution of the collision system. (5) Subducted continental crust to depths >250–300 km becomes significantly denser than the surrounding mantle due to the appearance in the solid assemblage of high-density phases such as hollandite and stishovite; this provides extra negative buoyancy to the slab and precludes the exhumation of crustal components. This supports the idea of the existence of a “depth of no return” for continental material at around 250 km depth.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 169.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 219.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 219.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  • Afonso JC, Ranalli G (2004) Crustal and mantle strengths in continental lithosphere: is the jelly sandwich model obsolete? Tectonophysics 394:221–232

    Article  Google Scholar 

  • Afonso JC, Fernàndez M, Ranalli G, Griffin W, Connolly J, (2008a) Integrated geophysical-petrological modelling of the lithospheric-sublithospheric upper mantle: methodology and applications. Geochem Geophys Geosyst 9. doi:10.1029/2007GC001834

  • Afonso JC, Zlotnik S, Fernàndez M (2008b) The effects of compositional and rheological stratifications on small-scale convection under the oceans: implications for the thickness of oceanic lithosphere and seafloor flattening. Geophys Res Lett 35:L20308. doi:10.1029/2008GL035419

    Article  Google Scholar 

  • Afonso JC, Ranalli G, Fernndez M, Griffin W, O’Reilly S, Ulrich F (2010) On the vp/vs – mg correlation in mantle peridotites: implications for the identification of thermal and compositional anomalies in the upper mantle. Earth Planet Sci Lett 289:606–618

    Article  Google Scholar 

  • Ague J (2003) Fuid flow in the deep crust. In: Rudnick RL, Holland HD, Turekian KK (eds) The crust, vol 3, Treatise on geochemistry. Elsevier, Oxford, pp 195–228

    Google Scholar 

  • Akashi A, Nishihara Y, Takahashi E, Nakajima Y, Tange Y, Funakoshi K (2009) Orthoenstatite/clinoenstatite phase transformation in MgSiO3 at high-pressure and high-temperature determined by in situ X-ray diffraction: implications for nature of the X discontinuity. J Geophys Res 114:B04206. doi:10.1029/2008JB005894

    Article  Google Scholar 

  • Audetat A, Keppler H (2004) Viscosity of fluids in subduction zones. Science 303:513–516

    Article  Google Scholar 

  • Batchelor GK (1967) An introduction to fluid dynamics. Cambridge University Press, UK

    Google Scholar 

  • Beaumont C, Jamieson RA, Butler JP, Warren CJ (2009) Crustal structure: a key constraint on the mechanism of ultra-high-pressure rock exhumation. Earth Planet Sci Lett 287:116–129

    Article  Google Scholar 

  • Behn M, Kelemen P (2003) The relationship between seismic P-wave velocity and the composition of anhydrous igneous and meta-igneous rocks. Geochem Geophys Geosys 4:1041. doi:10.1029/2002GC000393

    Article  Google Scholar 

  • Burov E, Yamato P (2008) Continental plate collision, p-t-t-z conditions and unstable vs stable plate dynamics: insights from thermomechanical modelling. Lithos 103:178–204

    Article  Google Scholar 

  • Capitanio F, Morra G, Goes S, Weinberg R, Moresi L (2010) India-sia convergence driven by the subduction of the greater indian continent. Nature Geosci 3. doi: 10.1038/NGEO725

  • Carlson RW and Miller (2003) Bound water content of the lower oceanic crust estimated from modal analyses and seismic velocities of oceanic diabase and gabbro. Geophys Res Lett RG1001.

    Google Scholar 

  • Carlson RL, Miller D (2003) Mantle wedge water contents estimated from seismic velocities in partially serpentinized peridotites. Geophys Res Lett 30. doi:10.1029/2002GL016600

  • Chopin C (2003) Ultrahigh-pressure metamorphism: tracing continental crust into the mantle. Earth Planet Sci Lett 212:1–14

    Article  Google Scholar 

  • Christensen U, Yuen D (1985) Layered convection induced by phase transitions. J Geophys Res 90:10291–10300

    Article  Google Scholar 

  • Clift P, Vannucchi P (2004) Controls on tectonic accretion versus erosion in subduction zones: implications for the origin and recycling of the continental crust. Rev Geophys 42, RG2001. doi:10.1029/2003RG000127

  • Connolly JAD (2009) Geodynamic equation of state: what and how. Geochem Geophys Geosyst 10, Q10014. doi:10.1029/2009GC002540

  • Dahlen F.A, Suppe J (1988) Mechanics, growth, and erosion of mountain belts. In: Clark S.P, Jr., Burchfield B.C, and Suppe J (eds) Processes in continental lithospheric deformation. Geol. Soc. Am. Spec. Pap., 218:161–178

    Google Scholar 

  • Davis J, Blanckenburg F (1995) Slab breakoff: q model of lithosphere detachment and its test in the magmatism and deformation of collisional orogens. Earth Planet Sci Lett 129:85–102

    Article  Google Scholar 

  • Davis J, Blanckenburg F (1998) Thermal controls on slab breakoff and the rise of high-pressure rock during continental collisions. In: Hacker B, Liou J (eds) When continents collide. Dordrecht, Kluwer, pp 97–115

    Google Scholar 

  • Dewey JF (2005) Orogeny can be very short. Proc Natl Acad Sci USA 102:15286–15293

    Article  Google Scholar 

  • Dobrzhinetskaya LF, Green HW (2007) Experimental studies of mineralogical assemblages of metasedimentary rocks at earths mantle transition zone conditions. J Metam Geol 25:83–96

    Article  Google Scholar 

  • Ernst W, Liou J (2008) High and ultrahigh pressure metamorphism: past results and future prospects. Am Mineral 93:1771–1786

    Article  Google Scholar 

  • Faccenda M, Minelli G, Gerya T (2009) Coupled and decoupled regimes of continental collision: numerical modeling. Earth Planet Sci Let 278:337–349

    Article  Google Scholar 

  • Fernàndez M, Afonso J, Ranalli G (2010) The deep lithospheric structure of the Namibian volcanic margin. Tectonophysics 481:68–81

    Article  Google Scholar 

  • Fullea J, Fernàndez M, Afonso JC, Vergés J, Zeyen H (2010) The structure and evolution of the lithosphere-asthenosphere boundary beneath the Atlantic-Mediterranean transition region. Lithos 120(1–2):74–95

    Article  Google Scholar 

  • Geoffroy L (2005) Volcanic passive margins. C R Geosci 337:1395–1408

    Article  Google Scholar 

  • Gerya T, Yuen D, Maresch W (2004) Thermomechanical modelling of slab detachment. Earth Planet Sci Lett 226:101–106

    Article  Google Scholar 

  • Gerya T, Connolly J, Yuen D, Gorczyk W, Capel A (2006) Seismic implications of mantle wedge plumes. Phys Earth Planet Int 156:59–74

    Article  Google Scholar 

  • Griffin WL, O’Reilly S, Afonso J, Begg G (2009) The composition and evolution of lithospheric mantle: a re-evaluation and its tectonic implications. J Petrol 50:1185–1204

    Article  Google Scholar 

  • Green H, Dobrzhinetskaya L, Bozhilov K.N. (2000). Mineralogical and experimental evidence for very deep exhumation from subduction zones. J. Geodyn., 30:61–76

    Google Scholar 

  • Guillot S, Garzanti E, Baratoux D, Marquer D, Maheo G, Sigoyer J (2003) Reconstructing the total shortening history of the nw himalaya. Geochem Geophys Geosys 4:1064. doi:10.1029/2002GC000484

    Article  Google Scholar 

  • Haggerty S.E, Sautter V (1990) Ultradeep (greater than 300 kilometers), ultramafic upper mantle xenoliths. Science, 248:993–996

    Google Scholar 

  • Hart SR, Hauri EH, Oschman LA, Whitehead JA (1992) Mantle plumes and entrainment: isotopic evidence. Science 256:517–520

    Article  Google Scholar 

  • Hebert L, Antoshechkina P, Asimow P, Gurnis M (2009) Emergence of a low-viscosity channel in subduction zones through the coupling of mantle flow and thermodynamics. Earth Planet Sci Lett 278:243–256

    Article  Google Scholar 

  • Hirschmann MM (2006) Water, melting, and the deep Earth’s H2O cycle. Ann Rev Earth Planet Sci 34:629–653

    Article  Google Scholar 

  • Hirth G, Kohlstedt DL (2003) Rheology of the upper mantle and the mantle wedge: a view from the experimentalists. In: Eiler J (ed) Inside the subduction factory, vol 138, Geophysical monograph., pp 83–105

    Chapter  Google Scholar 

  • Holland T, Powell R (1998) An internally consistent thermodynamic data set for phases of petrological interest. J Metam Geol 16:309–343

    Article  Google Scholar 

  • Hyndman R, Peacock S (2003) Serpentinization of the forearc mantle. Earth Planet Sci Lett 212:417–432

    Article  Google Scholar 

  • Irifune T, Isshiki M (1998) Iron partitioning in a pyrolite mantle and the nature of the 410-km seismic discontinuity. Nature 392:702–705

    Article  Google Scholar 

  • Irifune T, Ringwood A, Hibberson W (1994) Subduction of continental crust and terrigenous and pelagic sediments: an experimental study. Earth Planet Sci Lett 126:351–368

    Article  Google Scholar 

  • Iwamori H, McKenzie D, Takahashi E (1995) Melt generation by isentropic mantle upwelling. Earth Planet Sci Lett 134:253–266

    Article  Google Scholar 

  • Jarvis G, McKenzie D (1980) Convection in a compressible fluid with infinite prandtl number. J Fluid Mech 96:515–583

    Article  Google Scholar 

  • Jordan TH (1978) Composition and development of the continental tectosphere. Nature 274:544–548

    Article  Google Scholar 

  • Karato S-I (2008) Deformation of Earth materials: an introduction to the rheology of the solid Earth. Cambridge University Press, Cambridge, UK, 463 p

    Book  Google Scholar 

  • Katz RF, Spiegelman M, Langmuir C (2003) A new parameterization of hydrous mantle melting. Geochem Geophys Geosyst 4(9):1073. doi:10.1029/2002GC000433

    Article  Google Scholar 

  • Keskin M (2007) Eastern Anatolia: a hotspot in a collision zone without a mantle plume. In: Foulger G, Jurdy D (eds) Plates, plumes and planetary processes. Geological Society of America Special Paper, vol 430. The Geological Society of America, pp 693–722, doi:10.1130/2007.2430

  • Kirby S.H., Kronenbrg A.K., (1987) Rheology of the lithosphere:selected topics. Rev. Geophys. 25:1219–1244

    Google Scholar 

  • Kirby S.H. (1983) Rheology of the lithosphere. Rev. Geophys., 21:1458–1487

    Google Scholar 

  • Kundu PK (1990) Fluid mechanics. Academic, California, USA

    Google Scholar 

  • Lange R, Charmichael I (1987) Densities of Na2O, K2O, CaO, MgO, FeO, Fe2O3, Al2O3, TiO2, SiO2 liquids: new measurements and derived partial molar properties. Geochim Cosmochim Acta 51(11):2931–2946

    Article  Google Scholar 

  • Lee C-TA, Chen W-P (2007) Possible density segregation of subducted oceanic lithosphere along a weak serpentinite layer and implications for compositional stratification of the earth’s mantle. Earth Planet Sci Lett 255:357–366

    Article  Google Scholar 

  • Li Z, Gerya T (2009) Polyphase formation and exhumation of high- to ultrahigh-pressure rocks in continental subduction zone: numerical modeling and application to the Sulu ultrahigh-pressure terrane in eastern china. J Geophys Res 114:B09406. doi:10.1029/2008JB005935

    Article  Google Scholar 

  • Liu L, Zhang J, Green H, Jin Z, Bozhilov K.N. (2007) Evidence of former stishovite in metamorphosed sediments, implying subduction to > 350 km. Earth Planet. Sci. Lett., 263:180–191

    Google Scholar 

  • Mackwell S.J., Zimmerman M.E., Kohlstedt D.L., (1998) Hightemperature deformation of dry diabase with application to tectonics on Venus. J. Geophys. Res. 103:975–984

    Google Scholar 

  • McDonough WF, Sun S (1995) The composition of the Earth. Chem Geol 120:223–253

    Article  Google Scholar 

  • Miyajima N, Yagi T, Hirose K et al (2001) Potential host phase of aluminum and potassium in the earths lower mantle. Am Mineral 86:740–746

    Google Scholar 

  • Moresi L, Dufur F, Mühlhaus H (2003) A lagrangian integration point finite element method for large deformation modeling of viscoelastic geomaterials. J Comp Phys 184:476–497

    Article  Google Scholar 

  • Nakagawa T, Tackley P, Deschamps F, Connolly J (2009) Incorporating self-consistently calculated mineral physics into thermochemical mantle convection simulations in a 3-D spherical shell and its influence on seismic anomalies in Earths mantle. Geochem Geophys Geosys 10:Q03004. doi:10.1029/2008GC002280

    Article  Google Scholar 

  • Nikolaeva K, Gerya T, Connolly J (2008) Numerical modelling of crustal growth in interoceanic volcanic arcs. Phys Earth Planet Int 171:336–356

    Article  Google Scholar 

  • Ono S (1998) Stability limits of hydrous minerals in sediment and mid-ocean ridge basalt compositions: implications for water transport in subduction zones. J Geophys Res 103:18253–18267

    Article  Google Scholar 

  • Ono S (1999) High temperature stability limit of phase egg, AlSiO3(OH). Contrib Mineral Petrol 137:83–89

    Article  Google Scholar 

  • Patiño Douce A, McCarthy T (1998) Melting of crustal rocks during continental collision and subduction. In: Hacker B, Liou J (eds) When continents collide. Dordrecht, Kluwer, pp 27–55

    Google Scholar 

  • Phipps Morgan J (2001) Thermodynamics of pressure release melting of a veined pudding mantle. Geochem Geophys Geosys 2, 2000GC000049

    Google Scholar 

  • Pitzer K, Sterner S (1994) Equations of state valid continuously from zero to extreme pressures for h2o and co2. J Chem Phys 101(4):3111–3116

    Article  Google Scholar 

  • Plank T, Langmuir C (1998) The chemical composition of subducting sediment and its consequences for the crust and mantle. Chem Geol 145:325–394

    Article  Google Scholar 

  • Ranalli G (1995) Rheology of the Earth, 2nd edn. Chapman & Hall, London

    Google Scholar 

  • Ranalli G, Pellegrini R, D’Offizi S (2000) Time dependence of negative buoyancy and the subduction of continental lithosphere. J Geodyn 30:539–555

    Article  Google Scholar 

  • Rapp R, Irifune T, Shimizu N, Nishiyama N, Norman MD, Inoue T (2008) Subduction recycling of continental sediments and the origin of geochemically enriched reservoirs in the deep mantle. Earth Planet Sci Lett 271:14–23

    Article  Google Scholar 

  • Reddy J, Gartling D (2000) The finite element method in heat transfer and fluid dynamics, 2nd edn. CRC Press, USA

    Google Scholar 

  • Renshaw C, Schulson E (2004) Plastic faulting: brittle-like failure under high confinement. J Geophys Res 109:B09207. doi:10.1029/2003JB002945

    Article  Google Scholar 

  • Ringwood AE (1982) Phase transformations and differentiation in subducted lithosphere: implications for mantle dynamics, basalt petrogenesis, and crustal evolution. J Geol 90:611–643

    Article  Google Scholar 

  • Rudnick RL, Gao S (2003) Composition of the continental crust. In: Rudnick RL, Holland HD, Turekian KK (eds) The crust, vol 3, Treatise on geochemistry. Elsevier, Oxford, pp 1–64

    Google Scholar 

  • Scambelluri M, Pettke T, van Roermund H.L.M. (2008) Majoritic garnets monitor deep subduction fluid flow and mantle dynamics. Geology, 36:59–62

    Google Scholar 

  • Schilling J, Zajac M, Evans R, Johnston T, White W, Devine J, Kingsley R (1983) Petrologic and geochemical variations along the Mid-Atlantic Ridge from 29 degrees N to 73 degrees N. Am J Sci 283:510–586

    Article  Google Scholar 

  • Schmidt M.W, Poli S (1998) Experimentally based water budgets for dehydrating slabs and consequences for arc magma generation. Earth Planet. Sci. Lett., 163:361–379

    Google Scholar 

  • Schubert G, Turcotte DL, Olson P (2001) Mantle convection in the Earth and planets. Cambridge University Press, Cambridge, UK

    Book  Google Scholar 

  • Sella GF, Dixon T, Mao A (2002) Revel: a model for recent plate velocities from space geodesy. J Geophys Res 107(B4), 2081, doi:10.1029/2000JB000033

  • Shimada M (1993) Lithosphere strength inferred from fracture strength of rocks at high confining pressures and temperatures. Tectonophysics 217:55–64

    Article  Google Scholar 

  • Sobolev S, Babeyko A (2005) What drives orogeny in the Andes? Geology 33:617–620

    Article  Google Scholar 

  • Song S, Zhang L, Niu Y. (2004) Ultra-deep origin of garnet peridotite from the North Qaidam ultrahigh-pressure belt, Northern Tibetan Plateau, NW China. Am. Min., 89:1330–1336

    Google Scholar 

  • Takahashi N, Kodaira S, Tatsumi Y, Kaneda Y, Suyehiro K (2008) Structure and growth of the Izu-Bonin-Mariana arc crust: 1 seismic constraint on crust and mantle structure of the Mariana arcback-arc system. J Geophys Res 113:B01104. doi:10.1029/2007JB005120

    Article  Google Scholar 

  • Tatsumi Y, Eggins S (1995) Subduction zone magmatism. Blackwell, Oxford, UK

    Google Scholar 

  • Tirone M, Ganguly J, Morgan J (2009) Modeling petrological geodynamics in the Earths mantle. Geochem Geophys Geosys 10:Q04012. doi:10.1029/2008GC002168

    Article  Google Scholar 

  • Toussaint G, Burov E, Avouac J-P (2005) Tectonic evolution of a continental collision zone: a thermomechanical numerical model. Tectonics, TC6003, doi:10.1029/2003TC001604

  • Turner S, Evans P, Hawkesworth C (2001) Ultrafast source-to-surface movement of melt at island arcs from 226ra-230th systematics. Science 292:1363–1366, doi: 10.1126/science.1059904

    Article  Google Scholar 

  • van Keken P, Karato S-I, Yuen D (1996) Rheological control of oceanic crust separation in the transition zone. Geophys Res Lett 23:1821–1824

    Article  Google Scholar 

  • Vry J, Powell R, Golden K, Petersen K (2009) The role of exhumation in metamorphic dehydration and fluid production. Nat Geosci 3:31–35

    Article  Google Scholar 

  • Wark D, Williams C, Watson E, Price J (2003) Reassesment of pore shapes in microstructurally equilibrated rocks, with implications for permeability of the upper mantle. J Geophys Res 108:2050. doi:10.1029/2001JB001575

    Article  Google Scholar 

  • Warren C, Beaumont C, Jamieson RA (2008) Formation and exhumation of ultra-high-pressure rocks during continental collision: role of detachment in the subduction channel. Geochem Geophys Geosys 9(4). doi:10.1029/2007GC001839

  • Woodland A, Angel R (1997) Reversal of the orthoferrosilite – high-P clinoferrosilite transition, a phase diagram for FeSiO3 and implications for the mineralogy of the Earth’s upper mantle. Eur J Mineral 9:245–254

    Google Scholar 

  • Workman R, Hart S, Jackson M, Regelous M, Farley KA, Blusztajn J, Kurz M, Staudigel H (2004) Recycled metasomatized lithosphere as the origin of the Enriched Mantle II (EM2) end-member: evidence from the Samoan volcanic chain. Geochem Geophys Geosyst 5, Q04008, doi:10.1029/2003GC000623

  • Wu Y, Fei Y, Jin Z, Liu X (2009) The fate of subducted upper continental crust: an experimental study. Earth Planet Sci Lett 282:275–284

    Article  Google Scholar 

  • Xu W, Lithgow-Bertelloni C, Stixrude L, Ritsema J (2008) The effect of bulk composition and temperature on mantle seismic structure. Earth Planet Sci Lett 275:70–79

    Article  Google Scholar 

  • Yamato P, Agard P, Burov E, Le Pourhiet L, Jolivet L, TiberiBurov C (2007) Burial and exhumation in a subduction wedge: mutual constraints from thermomechanical modeling and natural P-T-t data (schistes lustrs, western alps). J Geophys Res 112:B07410. doi:10.1029/2006JB004441

    Article  Google Scholar 

  • Yamato P, Burov E, Agard P, Le Pourhiet L, Jolivet L (2008) HP-UHP exhumation during slow continental subduction: self-consistent thermodynamically and thermomechanically coupled model with application to the western alps. Earth Planet Sci Lett 271:63–74

    Article  Google Scholar 

  • Ye K, Cong B, Ye D (2000) The possible subduction of continental material to depths greater than 200 km. Nature, 407:734–736

    Google Scholar 

  • Yoshioka S, Wortel M (1995) 3D numerical modeling of detachment of subducted lithosphere. J Geophys Res 100(20):223–244

    Google Scholar 

  • Zang SX, Weia R, Ninga J (2007) Effect of brittle fracture on the rheological structure of the lithosphere and its application in the ordos. Tectonophysics 429:267–285

    Article  Google Scholar 

  • Zhai S, Ito E (2008) Phase relations of CaAl4Si2O11 at high-pressure and high-temperature with implications for subducted continental crust into the deep mantle. Phys Earth Planet Int 167(3–4):161–167

    Article  Google Scholar 

Download references

Acknowledgments

We would like to thank the editors D. Brown and P. Ryan for inviting us to contribute to this special volume. We are also indebted to all the participants of the field trip IGCP 524 (Ireland) for stimulating discussions and explanations in the field. We thank S. Saxena and W. Yong for unselfishly sharing their databases and helping in the implementation of new end-members, and G. Ranalli, G. Houseman, and B. Kaus for their detailed reviews and suggestions. Discussions with F. Capitanio, M. Faccenda, S. Turner, T. Rushmere, K. Grant and B. Schaefer were particularly useful. We also thank the team at VPAC and the geodynamics research group at Monash for their support. Most figures were produced with the open-source visualization package Paraview. This work has been supported by Macquarie University (JCA) and Monash University (SZ) grants and by the Monash e-Research Centre and ITS-Research Support Services through the use of the Monash Sun Grid cluster. This is contribution 659 from the Australian Research Council National Key Centre for the Geochemical Evolution and Metallogeny of Continents (http://www.gemoc.mq.edu.au).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to J. C. Afonso .

Appendix A

Appendix A

1.1 A.1 Governing Equations

The dynamics of the problem analyzed in this paper is governed by the usual conservation equations of fluid dynamics (under the continuum hypothesis), namely the conservation of mass, momentum, and energy (cf. Batchelor 1967; Kundu 1990). The conservation of mass is the most obvious condition and requires

$$ \frac{{\partial \rho }}{{\partial t}} + \frac{{\partial (\rho {u_i})}}{{\partial {x_i}}} = 0 $$
(3.1)

where t is time, ρ bulk density, u i velocity, and x i refers to a Cartesian coordinate system. Equation (3.1) is also known as the continuity equation.

The conservation of momentum is the result of applying Newton´s law of motion to an infinitesimal fluid element. In its differential form it reads

$$ \rho \left( {\frac{{\partial {u_i}}}{{\partial t}} + {u_j}\frac{{\partial {u_i}}}{{\partial {x_j}}}} \right) = \rho g - \frac{{\partial P}}{{\partial {x_i}}} + \frac{{\partial {{\tau \prime }_{ij}}}}{{\partial {x_j}}} $$
(3.2)

where

$$ {\tau \prime _{ij}} = 2\mu \left[ {\frac{1}{2}\left( {\frac{{\partial {u_i}}}{{\partial {x_j}}} + \frac{{\partial {u_j}}}{{\partial {x_i}}}} \right) - \frac{1}{3}\left( {\frac{{\partial {u_i}}}{{\partial {x_i}}}} \right){\delta_{ij}}} \right] = 2\mu \left( {{\epsilon_{ij}} - \frac{1}{3}{\epsilon_{ii}}{\delta_{ij}}} \right) $$
(3.3)

is the deviatoric part of the total stress tensor \( {\sigma_{ij}} = - P{\delta_{ij}} + {\tau \prime _{ij}} \) for a Newtonian fluid, P is pressure (i.e., −1/3 σ ii ), g the acceleration of gravity, μ the dynamic shear viscosity, δ ij the Kronecker delta, and ε ij the strain rate tensor (cf. Batchelor 1967). Equation (3.2) is the 3D general form of the Navier-Stokes equation of motion. Note that the so-called Stokes assumption, in which the bulk viscosity is neglected, has been adopted in (3.3) (Kundu 1990).

The equation of conservation of energy relevant to our study includes five terms:

$$ \rho \frac{{DE}}{{Dt}} = {\sigma_{ij}}\frac{{\partial {u_i}}}{{\partial {x_j}}} + H + \rho Ls\frac{{D\chi }}{{Dt}} + \rho Lm\frac{{DF}}{{Dt}} - \frac{{\partial q}}{{\partial {x_i}}} $$
(3.4)

where E the is internal energy per unit mass, H the internal heat generation (volumetric) due to radioactive decay, Ls the latent heat of transformation arising from solid-solid phase changes, Lm the latent heat of melting, q the conductive heat flow, χ the mass fraction of the new phase across a divariant solid-state reaction, F the mass fraction of melt, and D/Dt refers to the material derivative. The first term in the right-hand side of (3.4) is a mechanical contribution due to compressions and deformations of a fluid element without change of its velocity (Batchelor 1967). Using the above definition of total stress tensor σ ij , (3.4) can be rewritten as

$$ \rho \frac{{DE}}{{Dt}} = - P\frac{{\partial {u_i}}}{{\partial {x_i}}} + \Phi + H + \rho Ls\frac{{D\chi }}{{Dt}} + \rho Lm\frac{{DF}}{{Dt}} - \frac{{\partial q}}{{\partial {x_i}}} $$
(3.5)

where

$$ \Phi = 2\mu {\left[{\epsilon_{ij}} - 1/3\frac{{\partial {u_i}}}{{\partial {x_i}}}{\delta_{ij}}\right]^2} = \tau_{ij}^\prime {\epsilon_{ij}} $$
(3.6)

is the viscous dissipation through shear deformation [i.e., there are no volume changes associated with the tensor Equation (3.6)], sometimes referred to as “shear heating” in the literature. Using the following thermodynamic identities

$$ TdS = dE + PdV $$
(3.7)
$$ dV = - \frac{{d\rho }}{{{\rho^2}}} $$
(3.8)

which imply

$$ \frac{{TDS}}{{Dt}} = \frac{{DE}}{{Dt}} - \frac{P}{{{\rho^2}}}\frac{{D\rho }}{{Dt}} $$
(3.9)

we can rewrite (3.5) as

$$ \rho T\frac{{DS}}{{Dt}} = \Phi + H + \rho Ls\frac{{D\chi }}{{Dt}} + \rho Lm\frac{{DF}}{{Dt}} - \frac{{\partial q}}{{\partial {x_i}}} $$
(3.10)

where we made use of (3.1) to eliminate the terms affected by P. Equation (3.7) implicitly includes all the “non-conventional” entropy production terms (e.g., due to phase changes) in the E term. Recalling that for a homogeneous fluid the rate of change of entropy is

$$ \frac{{DS}}{{Dt}} = \frac{{Cp}}{T}\frac{{DT}}{{Dt}} - \frac{\alpha }{\rho }\frac{{DP}}{{Dt}} $$
(3.11)

we obtain the more familiar form

$$ \eqalign{ \rho Cp\left( {\frac{{DT}}{{Dt}} - \frac{{T\alpha }}{{\rho Cp}}\frac{{DP}}{{Dt}}} \right) \cr \quad= \Phi + H + \rho Ls\frac{{D\chi }}{{Dt}} + \rho Lm\frac{{DF}}{{Dt}} - \frac{{\partial q}}{{\partial {x_i}}}} $$
(3.12)

where α is the coefficient of thermal expansion and Cp the isobaric specific heat capacity. Since we introduce latent heat effects directly into the energy equation, α and Cp do not need to be replaced by their “effective” or “apparent” counterparts in the phase change temperature range (e.g., Reddy and Gartling 2000; Li and Gerya 2009). Note that the second term on the left-hand side of (3.12) can be written as

$$ \frac{{T\alpha }}{{\rho Cp}}\frac{{DP}}{{Dt}} = {\left( {\frac{{\partial T}}{{\partial P}}} \right)_{\hskip-3ptS}}\frac{{DP}}{{Dt}} = \frac{{T\alpha }}{{\rho Cp}}\left( {\frac{{\partial P}}{{\partial t}} + {u_i}\frac{{\partial P}}{{\partial {x_i}}}} \right) $$
(3.13)

For the Earth’s mantle it is appropriate to neglect the term ∂P/∂t and assume that the vertical gradient in P is dominant and equal to ρg (i.e., hydrostatic profile; Schubert et al. 2001). Accordingly, (3.13) becomes

$$ \rho Cp\left( {\frac{{DT}}{{Dt}} - \frac{{{H_{ad}}}}{{\rho Cp}}} \right) = \Phi + H + \rho Ls\frac{{D\chi }}{{Dt}} + \rho Lm\frac{{DF}}{{Dt}} - \frac{{\partial q}}{{\partial {x_i}}} $$
(3.14)

where H ad  ≡ Tαρu z g is the energy gained/lost by adiabatic compression/decompression and u z is the vertical component of velocity. An alternative useful form of this latter relation is

$$ \eqalign{ \rho Cp\left( {\frac{{\partial T}}{{\partial t}} + {u_i}\frac{{\partial T}}{{\partial {x_i}}}} \right) \cr \quad= \frac{\partial }{{\partial {x_i}}}\left( { - {k_{ij}}\frac{{\partial T}}{{\partial {x_i}}}} \right) + \Phi + H + {H_{ad}} + \rho Ls\frac{{D\chi }}{{Dt}} + \rho Lm\frac{{DF}}{{Dt}}} $$
(3.15)

where k ij is the thermal conductivity tensor. Equations (3.1), (3.2), and (3.15) represent the fundamental set of equations governing the heat transfer and flow of a compressible Newtonian fluid. The insertion of an equation of state describing density as a function of other intensive variables (e.g., P, T, F, χ) completes the definition of the mathematical problem (except for the non-Newtonian case). In what follows, we describe a number of approximations and modifications to the fundamental set of equations.

1.2 A.2 Approximations

We adopt a modified version of the extended Boussinesq approximation (e.g., Christensen and Yuen 1985) in solving the system (3.1), (3.2), and (3.15). The Prandlt Pr and Match M numbers of a fluid determine the effects of inertia and compressibility on the flow (cf. Jarvis and McKenzie 1980; Schubert et al. 2001). For the Earth\prime s mantle, the product Pr M 2 is \( \ll 1 \) (Jarvis and McKenzie 1980), and therefore the material can be considered incompressible except in regions affected by phase transitions (see below). Equation (3.1) thus becomes

$$ \frac{{\partial {u_i}}}{{\partial {x_i}}} = 0 $$
(3.16)

Likewise, given the effectively infinite Prandlt number of the mantle, inertia terms in the momentum (3.2) are neglected. Using (3.16), we rewrite (3.2) as

$$ 0 = \rho g - \frac{{\partial P}}{{\partial {x_i}}} + \frac{\partial }{{\partial {x_j}}}\left[ {\mu \left( {\frac{{\partial {u_i}}}{{\partial {x_j}}} + \frac{{\partial {u_j}}}{{\partial {u_i}}}} \right)} \right] $$
(3.17)

Besides the inertia terms, the only difference between the (3.2) (compressible fluid) and (3.17) (incompressible fluid) is a term that includes the divergence ▽ = ∂/∂x i of the velocity field

$$ - \frac{\partial }{{\partial {x_j}}}\left[ {\frac{2}{3}\mu \left( {\frac{{\partial {u_i}}}{{\partial {x_i}}}} \right){\delta_{ij}}} \right] $$

The energy equation is made compatible with (3.16) by removing the term affected by the divergence of velocity in the viscous dissipation term (3.6), which becomes

$$ \Phi = 2\mu {\epsilon_{ij}}^2 $$
(3.18)

The latent heats L in the energy equation can be written as ΔS T, where ΔS is the entropy change of the phase transformation (solid-solid or solid-melt). Since entropy is an extensive quantity, the magnitude of ΔS in a rock experiencing a phase transition will depend on the actual amount of matter being transformed. For instance, although the ΔS associated with the spinel-garnet transformation is ~25 J Kg−1 K−1 in the system MgAl2O4 (spinel) + 4 MgSiO3 (orthopyroxene) \( \rightleftarrows \) Mg3Al2Si3O12 (pyrope) + Mg2SiO4 (forsterite), it becomes only ~4 J Kg−1 K−1 in a natural lherzolite with 4.5 wt% Al2O3 (see also Iwamori et al. 1995). Similarly, our thermodynamic calculations (see below) indicate that most metamorphic reactions typically involve ΔS < 8 J Kg−1 K−1. Such ΔS values result in local variations of the order of 5–10°C in the thermal structure and thus they can be ignored without losing any generality in large-scale models. On the other hand, reactions involving large mass fractions, such as the olivine-wadsleyite phase transitions in lherzolites (~60 vol% of the rock), can result in ΔS of the order of 30–45 J Kg−1 K−1. In this case, the local thermal perturbation close to the phase transition can be as high as 60–70°C for fast vertical flow, and therefore the contribution from latent heat release should not be neglected in the energy equation. We introduce latent heat effects by approximating the material derivative of χ by (no time effects)

$$ \frac{{D\chi }}{{Dt}} \equiv {u_i}\frac{{\partial \chi }}{{\partial {x_i}}} $$
(3.19)

where ∂χ/∂x i is assumed to be a step function (i.e., sharp transition) which is zero everywhere except at the transition, where it takes a value of 1/u i Δt. Appropriate values of ΔS are calculated as a function of bulk rock composition using a Gibbs free energy minimization formalism (see below).

The latent heat of melting is addressed in a similar manner, except that in this case the material derivative DF/Dt is a continuous function based on a parameterization of partial melting as a function of P, T, water content, and depletion (see below).

As stated above, the incompressibility assumption (3.16) is appropriate for modelling flow within the Earth as long as the Boussineq assumptions are fulfilled. This is not longer the case when density changes produced by mineral phase transitions are included in the model. Neglecting volume changes associated with phase transitions in the conservation equations has the net effect of “creating/destroying” mass when the material experiences a phase change (i.e., mass conservation is not ensured). This carries the obvious consequence of over/underestimating the buoyancy forces of the model, which in the case of a subducting plate has a major long-term effect on its buoyancy and thus on the dynamics of the model. We tackle this problem by modifying (3.16) and (3.17) to account for compressibility (i.e., non-zero divergence of the velocity field) in the regions affected by phase transitions. Equation (3.16) thus becomes

$$ \frac{{\partial {u_i}}}{{\partial {x_i}}} = - \frac{1}{\rho }\frac{{\partial \rho }}{{\partial {x_i}}}{u_i} \approx - \frac{1}{\rho }\frac{{\Delta \rho }}{{\Delta t}} $$
(3.20)

where the material change in density \( \tfrac{{\displaystyle\Delta \rho }}{{\displaystyle\Delta t}} \) is computed in each Lagrangian particle as the difference between the density in the current and previous time step (we only account for density changes due to mineral phase transitions, density variations related to melting are not included). The application of this non-zero divergence generates a localized velocity field that effectively compresses/expands the material at the time of the transition, therefore enforcing local mass conservation. In practice, since our numerical box is closed, we apply the mass correction only to lithospheric materials. This plays the role of simulating an “open” domain (i.e., sublithospheric convecting mantle can flow through the boundaries of the box to compensate for lithospheric volume changes). We note that this correction is not equivalent to modelling a fully compressible material, but rather numerically impose the changes in volume needed to conserve the mass in specific regions of the simulation domain, i.e., lithospheric materials. Similar strategies have been used by (e.g., Warren et al. 2008; Burov and Yamato 2008).

Equations (3.16) through (3.20) are solved using the finite-element platform Underworld (Moresi et al. 2003), which has been modified accordingly to account for all the processes described in this paper. The numerical accuracy of the mass correction term included in (3.20) is evaluated in each simulation by keeping track of the mass of the subducting plate trough time. However, this calculation is not trivial due to the difficulty in estimating the volumes of all bodies (i.e., the geometry of a body is defined by a set of moving particles that tend to mix with the particles of surrounding materials during the evolution of the model). We circumvent this problem by constructing Voronoi diagrams based on the particles and computing the mass of each particle as the product of its density times the area of the corresponding Voronoi cell. Therefore, the mass of a body is simply the sum of the mass of its constituents particles. We find that typical accumulated errors associated with mass conservation are <1%.

1.3 A.3 Rheological Relationship

Our numerical models have both plastic (brittle) and viscous (Newtonian and non-Newtonian) rheologies. The brittle behaviour of rocks is assumed to follow a modified version of the von Mises criterion, in which the material yields (plastically) when the following condition is met

$$ \sqrt {{{{\tau \prime }_{II}}}} \ge {c_0} + \mu (\rho gz) $$
(3.21)

where \( {\tau \prime _{II}} \) is the second invariant of the deviatoric stress tensor, c 0 is the “cohesive strength” and μ the coefficient of internal friction. Parameters c 0, and μ are specific for each material and vary among simulations. High-pressure failure of rocks (e.g., Shimada 1993; Renshaw and Schulson 2004) is modelled assuming a critical (constant) yield stress τ c between 250 and 350 MPa (See Fig. 3.2).

The assumed general flow law for both diffusion and dislocation mechanisms has the following form (e.g. Hirth and Kohlstedt 2003)

$$ \dot{\epsilon } = A{(\sigma )^n}{d^{ - m}}{\Psi_{{H_2}O}}\;\exp \left( { - \frac{{{E^*} + P{V^*}}}{{R\;T}}} \right) $$
(3.22)

where d is the average grain-size, σ the differential stress, A the pre-exponential factor, n the stress exponent, m the grain-size exponent, E * the activation energy, V * the activation volume, R the gas constant, and \( {\Psi_{{H_2}O}} \) a parameter dependent on the water content (see below). Note that the activation energy E * and volume V * already include the T and P dependence of OH dissolution in olivine (cf. Hirth and Kohlstedt 2003; Karato 2008). Applying the Levy-von Mises formalism to purely viscous fluids (Ranalli 1995; Karato 2008) allows defining the effective viscosity as

$$ \eta = \frac{1}{2}{({\tau \prime _E})^{1 - n}}{A^{ - 1}} = \frac{1}{2}{({\dot{\epsilon }_E})^{\tfrac{{1 - n}}{n}}}{A^{ - \tfrac{1}{n}}} $$
(3.23)

where τ E and \( {\dot{\epsilon }_E} \) are the “effective” second invariants of the deviatoric stress tensor \( [ = {(\tfrac{1}{2}{\tau \prime _{ij}}{\tau \prime _{ij}})^{1/2}}] \) and of the strain rate tensor \( [ = {(\tfrac{1}{2}{\dot{\epsilon }_{ij}}{\dot{\epsilon }_{ij}})^{1/2}}] \), respectively. Parameter A is a factor derived from the empirical Equation (3.22) and the type of experiment [for details see Chap. 4 of Ranalli (1995) and Chap. 3 of Karato (2008)]. All relevant parameters used to solve (3.21)–(3.23) are listed in Table 3.3.

Since diffusion and dislocation creep are thought to act simultaneously in the mantle, two different viscosities η diff and η disl are computed separately and then combined into an effective viscosity η eff. The latter is computed as the harmonic mean of η diff and η disl:

$$ \frac{1}{{{\eta_{\rm{eff}}}}} = \left( {\frac{1}{{{\eta_{\rm{diff}}}}} + \frac{1}{{{\eta_{\rm{disl}}}}}} \right) $$
(3.24)

This expression is truncated if the resulting viscosity is either greater or lower than two imposed cutoff values (1018 to 1024 Pa s). Although grain-size may change due to grain growth and dynamic recrystallization processes, its dependence on mantle flow conditions is poorly known and therefore we consider only constant grain sizes. For average grain sizes d ≳3.5 mm, dislocation creep represents an important component of the effective viscosity at depths < 200 km. For smaller values of d, diffusion creep becomes dominant. In this context, we note that synthetic seismological models of oceanic mantle suggest d ≳ 3.5 mm (Afonso et al. 2008a, b). All simulations shown in this work are carried out with a constant d = 5 mm.

1.4 A.4 Melting Model

We adopt a simple melting model based on the parameterizations of Katz et al. (2003). Melting of crustal components are not considered in the present work. The melt fraction of mantle rocks as a function of P, T, water content, and depletion ξ is expressed as

$$ {F_{(T,P,{H_2}O,\xi )}} = {\left( {\frac{{T - {T_s}}}{{{T_l} - {T_s}}}} \right)^{\beta 1}} $$
(3.25)

where

$$ {T_s} = {A_1} + {A_2}P + {A_3}{P^2} - \Delta {T_w} + \Delta {T_\xi } $$
(3.26)
$$ {T_l} = {B_1} + {B_2}P + {B_3}{P^2} - \Delta {T_w} + \Delta {T_\xi } $$
(3.27)

Parameters A n and B n are fitting parameters obtained from regressions through a large number of experimental data (Katz et al. 2003). ΔT w is a parameter that takes into account the temperature decrease in the solidus caused by the presence of water and has the simple form (Katz et al. 2003)

$$ \Delta {T_w} = KX_{{H_2}O}^\gamma $$
(3.28)

where K = 43°C wt%γ, γ = 0.75, and \( {X_{{H_2}O}} \) is the weight percent of water in the melt. The latter is obtained assuming batch melting and a representative bulk partition coefficient D w (Katz et al. 2003). Although it is expected that in natural systems D w will vary with P,T, and bulk composition, here we adopt a constant D w  = 0.008 as a representative average for our simulations; this value is in agreement with various experimental estimations (cf. Hirschmann 2006). The concomitant extraction of bound water during melting of hydrous assemblages is tracked using the same bulk partition coefficient D w as above.

In our simulations we assume “instantaneous” batch melting during each time step. Thus, all melt generated during a time step is extracted in the next time step, except for a small amount that remains in the solid assemblage as residual porosity (see below). The ΔT ξ parameter is introduced here to include the effect of incremental depletion on the solidus temperature. This effect, not considered in the original parameterization of Katz et al. (2003), arises due to the progressive extraction of fusible elements from the solid rock by partial melting, which in turn increases the solidus temperature of the solid residue. We adopt the following linear form

$$ \Delta {T_\xi } = 250F $$
(3.29)

where the result is in °C and F is in weight fraction (e.g., 0 < F < 1). The numerical value 250 is identical to that estimated by Phipps Morgan (2001) and used by Afonso et al. (2008a). The effect of melt depletion on bulk density is accounted for by tracking the total amount of melt extraction F T experienced by mantle particles and applying the following correction: ρ r  = ρ s  − ζ F T , where ρ r is the bulk density of the residual peridotite, ρ s is the bulk density of the unmelted (fertile) mantle, and ζ is a correction factor = 2 kg m−3 %F −1. For instance, if a mantle parcel has experienced a total of 10% melt extraction at a certain time of the simulation, the density of the residual solid assemblage will be 20 kg m−3 less than its fertile counterpart at identical P-T conditions.

The density of the partially melted rock is computed as

$$ {\rho_{eff}} = {\rho_s}(1 - F) + {\rho_m}F $$
(3.30)

where ρ s is the density of the solid aggregate and ρ m is the density of the melt. The former is retrieved from the energy minimization scheme; the latter is estimated as a linear function of P and T according to the relation

$$ {\rho_m} \;\quad=\quad\; {\rho_0} + {\left( {\frac{{\partial {\rho_0}}}{{\partial T}}} \right)_P}(T - {T_0}) + {\left( {\frac{{\partial {\rho_0}}}{{\partial P}}} \right)_T}(P - {P_0}) $$
(3.31)

where ρ 0 is a reference value at P 0 = 1 atm and T 0 = 800 K. The derivatives are estimated using the method of Lange and Charmichael (1987) for a melt with an average MgO content of 10 wt%. The results are (∂ρ 0/∂T) P ~ −10−4 kg m−3 K−1 and (∂ρ 0/∂P) T ~0.065 kg m−3 Gpa−1. We assume the existence of a critical residual porosity F c below which melts are immobile (Wark et al. 2003). Here we use a conservative value of 1% for F c . When the calculated F at a certain P-T-H2O condition becomes greater than F c , the “excess” melt (i.e., F − F c ) is extracted from the solid and transported instantaneously to the surface. This assumption is entirely justified by the high velocities of melts in the mantle wedge (e.g., Turner et al. 2001) in comparison to the duration of a typical time step in our simulations (50,000–200,000 years).

1.5 A.5 Darcy Flow of Water

Aqueous fluids released by dehydration reactions are assumed to migrate through the solid matrix as porus flow along the solid grain boundaries (Darcy\prime s flow). The effect of adding water into nominally-anhydrous minerals by diffusion is neglected here due to its minute contribution to the water balance compared to hydration/dehydration reactions.

Neglecting compaction effects and assuming that pressure variations from the lithostatic state are small (Burov and Yamato 2008), the fluid (melt or aqueous) velocity relative to the matrix velocity obeys the relation

$$ {V_f} - {V_s} = \frac{{\Delta \rho g{k_\phi }}}{{{\mu_f}\phi }} $$
(3.32)

V f and V s are the velocities of the fluid and solid matrix, respectively, φ the fluid fraction, μ f the dynamic viscosity of the fluid, and Δρ the density difference between the solid aggregate and the fluid. The permeability k φ is given by (Wark et al. 2003)

$$ {k_\phi } = \frac{{{d^2}{\phi^3}}}{{270}} $$
(3.33)

where d is the mean grain diameter (5 mm). As stated in (3.32), the densities and viscosities of the aqueous fluids are required to compute their velocities. Here we estimate the density with the Pitzer-Sterner equation of state for pure water (Pitzer and Sterner 1994). This assumption will underestimate the real density of fluids as the solubility of silica-rich components in water increases significantly at high pressures (Audetat and Keppler 2004), thus increasing the density of the fluid. The viscosities are somewhat more problematic due to their order-of magnitude variations over conditions pertaining to the mantle wedge (Audetat and Keppler 2004). In order to avoid overcomplicating the model with poorly constrained parameters, here we assume the viscosities of fluids to be constants and equal to 1 Pa s.

1.6 A.6 Thermodynamic Databases and Free Energy Minimization Strategy

We compute stable assemblages and all their relevant bulk properties (ρ, C p , α, bulk modulus, etc.) by free energy minimization using the software Perple_X (Connolly 2009). All calculations were performed in the system K2O–Na2O–TiO2–FeO–CaO–MgO–Al2O3–SiO2 + H2O (KNTFCMAS + H2O), which accounts for ≳99 wt% of the Earth´s crust and mantle. All stable assemblages and relevant physical properties were computed a priori as function of P, T, bulk composition, and water content and saved in “look-up” tables with a spacing of 75 MPa for P, 5.6 K for T, and 1 wt% for water content. Given the size of our numerical domain and the large range of compositions, we used two different thermodynamic databases and formalisms. For ultramafic upper mantle lithologies and for all crustal materials, we used an augmented-modified version of the Holland and Powell (1998) database (revised in 2002). The original data set has been augmented with the following phases: CAS* (CaAl4Si2O11), K-hollandite* (KAlSi3O8), Si-wadeite* (K2Si4O9), K-Cymrite* (KAlSi3O8⋅H2O), Mg-majorite (Mg4Si4O12), Na-majorite (Na2Al2Si4O12), Ca-perovskite* (CaSiO3), Mg-wadsleyite (Mg2SiO4), Fe-wadsleyite (Fe2SiO4), high-pressure (C2/c) clinoferrosilite (Fe4Si4O12), and high-pressure (C2/c) clinoenstatite (Mg4Si4O12); * denotes stoichiometric phases (i.e., not part of a solid-solutions). The so-called NAL and Egg phases as well as Ca-ferrite are stable at very high-pressures along subduction P-T paths (e.g., Ono 1999; Miyajima et al. 2001). However, they are not included in the present version of the database due to their small volume fraction in comparison with other dominant phases.

In order to obtain a reliable internally consistent data set within the system KNTFCMAS + H2O for the wide range 0 < P < 30 GPa and 273 < T < 2,000 K, the following phases in the original Holland and Powell (1998) database had to be slightly modified (always within experimental uncertainty): forsterite, fayalite, ferrosilite, enstatite, stishovite, tschermakite, Ca-tschermakite, Mg-tschermakite, grossular, pyrope, diopside, and A-phase. The modification involved mainly a refitting of the polynomials describing C p and α to make them compatible with high-pressure experiments, and in some cases a slight modification of the enthalpies of formation and molar volumes. However, we stress that the predictions of the original database at low and moderate (crustal and shallow upper mantle) T-P conditions were not affected to any significant extent by these modifications, since they mostly apply at high-pressure conditions. A full description of all solid-solution models and end-members will be published elsewhere (Afonso and Zlotnik, in preparation).

Our database has been calibrated based on experimental results in simplified systems. The thermodynamic parameters of the mentioned phases were refined in successive steps in a hierarchical manner. First, a consistent database was obtained for the MS, FS, MAS, and FAS systems. Succeeding refining stages considered the ternary systems KAS, CAS, FMS, CMS, and quaternary systems KAS + H2O, CMAS, FMAS. A subsequent comparison against experimental results in complex system such as NKCMFAS, KNTFCMAS + H2O, and KCMAS + H2O showed good agreement, and permitted a final refining of some phases and/or solutions (e.g., garnets and clinopyroxenes). Figure 3.15 shows a comparison of calculated phase diagrams for a peridotite using three different databases. It can be seen that at low P-T conditions our database (Fig. 3.15b) predicts similar equilibrium assemblages than the original database of Holland and Powell (1998) (Fig. 3.15a). At higher pressures, however, the original database fails to predict representative assemblages due to the lack of solid solution models for the solubility of pyroxenes into garnet, the transformation of olivine into wadsleyite, and the high-pressure monoclinic polymorph of orthopyroxene with C2/c symmetry. At high P-T conditions our phase diagram is therefore more similar to that predicted from the high-pressure database of Xu et al. (2008) (Fig. 3.15c).

Fig. 3.15
figure 15

Predicted phase diagrams of dry peridotite using the database of Holland and Powell (1998) (a), our augmented/modified version (b), and the database of Xu et al. (2008) (c). (d), (e) and (f) are the respective modal proportions along a 1330 °C adiabat.

Figure 3.16 shows the high T-P experimental results of Irifune and Isshiki (1998) together with predictions from our database for the same P-T and compositional conditions. Although the agreement with the experimental data is good, there are some differences at high pressures. Firstly, the experiments of Irifune and Isshiki (1998) did not constrain the amount of the monoclinic (clinopyroxene-structured) polymorph of orthopyroxene (e.g., Woodland and Angel 1997; Akashi et al. 2009), and therefore we cannot constrain well the effect that the presence (and disappearance) of the C2/c phase will have on the transfer of atomic species between garnet and clinopyroxene (see also Fig. 3.15). Although the behaviour of the C2/c phase in our calculations is in perfect agreement with experiments in the simple system FMS, it is still uncertain how this phase behaves in more complex systems such as the one in Fig. 3.16. Secondly, our database predicts high pressure garnets (P > 10 GPa) with relatively low contents in Na and high in Fe with respect to experiments on continental crust compositions (Irifune et al. 1994; Wu et al. 2009). Clinopyroxene, on the other hand, tend to be deficient in Fe and Mg. Although we have modified some end-member properties to make the database more reliable at high pressures, the above observations suggest that the solution models adopted in this study from experiments at low-moderate pressures require further refinements. Unfortunately, there are only a few high-P-T experiments in complex systems that could be used for this purpose, and not all of them are consistent (e.g., Irifune et al. 1994; Wu et al. 2009). Despite this limitation, Figs. 3.16 and 3.17 shows that our database reproduce high-P-T experimental observations satisfactorily for a wide range of compositions.

Fig. 3.16
figure 16

Comparison between experimental and theoretical phase proportions in dry pyrolite. Black circles with error bars are from Irifune and Ito (1998) based on high-T-P experiments. Red (grey) circles are predictions from our dataset. Stars are predictions from the dataset of Xu et al. (2008).

Fig. 3.17
figure 17

Comparison between experimental results (left column) and predictions from our dataset (right column). Experimental studies are (a) Irifune et al. (1994), (c) Ono (1998), (e) Wu et al. (2009), (g) Ono (1998).

The above discussion indicates that our database will predict reliable phase diagrams up to ≲30 GPa only for felsic, intermediate, and mafic compositions. For ultramafic compositions, however, predictions are only reliable up to ~15 GPa (i.e., above the wadsleyite-ringwoodite transition). Therefore, for mantle materials below this transition we use the thermodynamic database and formalism of Xu et al. (2008). This database is suitable for ultramafic compositions (but not for felsic, intermediate, or hydrated compositions) at transition zone and lower mantle conditions in the simplified system NCFMAS. For similar NCFMAS compositions, the differences in bulk ρ, α, and seismic velocities between the two databases are < 1%, ≲ 1.5, < 0.8%, respectively. Therefore, the discontinuity of such properties when passing from our database to that of Xu et al. (2008) is unlikely to have any significant effect on the dynamics of the system. In practice, we combine both databases at the Ol-Wads transition, where the properties of the system have a natural and significant discontinuity (i.e., the so-called 410d). We apply an interpolation function to smooth out the sharp contrasts in the physical properties.

Rights and permissions

Reprints and permissions

Copyright information

© 2011 Springer-Verlag Berlin Heidelberg

About this chapter

Cite this chapter

Afonso, J.C., Zlotnik, S. (2011). The Subductability of Continental Lithosphere: The Before and After Story. In: Arc-Continent Collision. Frontiers in Earth Sciences. Springer, Berlin, Heidelberg. https://doi.org/10.1007/978-3-540-88558-0_3

Download citation

Publish with us

Policies and ethics