Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Asthenospheric low-velocity zone consistent with globally prevalent partial melting

Abstract

The asthenosphere plays a fundamental role in present-day plate tectonics as its low viscosity controls how convection in the mantle below it is expressed at the Earth’s surface above. The origin of the asthenosphere, including the role of partial melting in reducing its viscosity and facilitating deformation, remains unclear. Here we analysed receiver-function data from globally distributed seismic stations to image the lower reaches of the asthenospheric low-seismic-velocity zone. We present globally widespread evidence for a positive seismic-velocity gradient at depths of ~150 km, which represents the base of a particularly low-velocity zone within the asthenosphere. This boundary is most commonly detected in regions with elevated upper-mantle temperatures and is best modelled as the base of a partially molten layer. The presence of the boundary showed no correlation with radial seismic anisotropy, which represents accumulated mantle strain, indicating that the inferred partial melt has no substantial effect on the large-scale viscosity of the asthenosphere. These results imply the presence of a globally extensive, partially molten zone embedded within the asthenosphere, but that low asthenospheric viscosity is controlled primarily by gradual pressure and temperature variations with depth.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Globally distributed receiver functions.
Fig. 2: The relationship between receiver-function phases and asthenospheric VSV.
Fig. 3: Modelling the PVG-150 with partial melt.
Fig. 4: The PVG-150, anisotropy, melt and viscosity.
Fig. 5: Interpretation of the observed PVG-150 and PVG-260 mantle velocity gradients.

Similar content being viewed by others

Data availability

Seismograms were downloaded from the IRIS Data Management Center (http://ds.iris.edu/ds/nodes/dmc/). The Sp receiver functions for all stations are available as raw data on the Figshare platform (https://doi.org/10.6084/m9.figshare.21706325). Geochemical data were downloaded from GEOROC (http://georoc.mpch-mainz.gwdg.de/Georoc) and are provided as Supplementary Table 1. Velocity model GLAD-M25 (ref. 22) was obtained from E. Bozdag. Velocity models SEMUCB-WM1 (ref. 21) and SEMUCB-UMQ59 were downloaded from the Berkeley Global Seismology Group (http://seismo.berkeley.edu/wiki_br/Main_Page). Velocity Model SAVANI-US42 was from T. W. Becker’s website (http://www-udc.ig.utexas.edu/external/becker/tdata.html). The attenuation model QRFSI12 (ref. 58) was from C. A. Dalton’s website (http://www.geo.brown.edu/research/Dalton/downloads.html); QsADR17 (ref. 60) was from E. Debayle’s website (http://perso.ens-lyon.fr/eric.debayle/). Other velocity models, including CAM2016 (refs. 2,40), S362ANI + M41 and SGLOBE-rani39 were from the IRIS archive (https://ds.iris.edu/ds/products/emc-earthmodels/).

Code availability

All computer codes used for data processing, analysis and plotting are available on request.

References

  1. Höink, T., Lenardic, A. & Richards, M. Depth-dependent viscosity and mantle stress amplification: implications for the role of the asthenosphere in maintaining plate tectonics. Geophys. J. Int. 191, 30–41 (2012).

    Article  Google Scholar 

  2. Priestley, K. & McKenzie, D. The relationship between shear wave velocity, temperature, attenuation and viscosity in the shallow part of the mantle. Earth Planet. Sci. Lett. 381, 78–91 (2013).

    Article  Google Scholar 

  3. Gung, Y., Panning, M. & Romanowicz, B. Global anisotropy and the thickness of continents. Nature 422, 707–711 (2003).

    Article  Google Scholar 

  4. Burgos, G. et al. Oceanic lithosphere–asthenosphere boundary from surface wave dispersion data. J. Geophys. Res. Solid Earth 119, 1079–1093 (2014).

    Article  Google Scholar 

  5. Beghein, C., Yuan, K., Schmerr, N. & Xing, Z. Changes in seismic anisotropy shed light on the nature of the Gutenberg discontinuity. Science 343, 1237–1240 (2014).

    Article  Google Scholar 

  6. Stixrude, L. & Lithgow-Bertelloni, C. Mineralogy and elasticity of the oceanic upper mantle: origin of the low-velocity zone. J. Geophys. Res. Solid Earth 110, B03204 (2005).

    Article  Google Scholar 

  7. Abers, G. A. et al. Reconciling mantle attenuation–temperature relationships from seismology, petrology, and laboratory measurements. Geochem. Geophys. Geosyst. 15, 3521–3542 (2014).

    Article  Google Scholar 

  8. Debayle, E., Bodin, T., Durand, S. & Ricard, Y. Seismic evidence for partial melt below tectonic plates. Nature 586, 555–559 (2020).

    Article  Google Scholar 

  9. Fischer, K. M. et al. A comparison of oceanic and continental mantle lithosphere. Phys. Earth Planet. Inter. 309, 106600 (2020).

  10. Holtzman, B. K. Questions on the existence, persistence, and mechanical effects of a very small melt fraction in the asthenosphere. Geochem. Geophys. Geosyst. 17, 470–484 (2016).

    Article  Google Scholar 

  11. Selway, K. & O’Donnell, J. P. A small, unextractable melt fraction as the cause for the low velocity zone. Earth Planet. Sci. Lett. 517, 117–124 (2019).

    Article  Google Scholar 

  12. Hansen, L. N., Qi, C. & Warren, J. M. Olivine anisotropy suggests Gutenberg discontinuity is not the base of the lithosphere. Proc. Natl Acad. Sci. USA 113, 10503–10506 (2016).

    Article  Google Scholar 

  13. Karato, S.-i On the origin of the asthenosphere. Earth Planet. Sci. Lett. 321, 95–103 (2012).

    Article  Google Scholar 

  14. Ma, Z. et al. Shear attenuation and anelastic mechanisms in the central Pacific upper mantle. Earth Planet. Sci. Lett. 536, 116148 (2020).

    Article  Google Scholar 

  15. Kind, R., Yuan, X. & Kumar, P. Seismic receiver functions and the lithosphere–asthenosphere boundary. Tectonophysics 536, 25–43 (2012).

    Article  Google Scholar 

  16. Liu, T. & Shearer, P. M. Complicated lithospheric structure beneath the contiguous US revealed by teleseismic S reflections. J. Geophys. Res. Solid Earth 126, e2020JB021624 (2020).

  17. Rychert, C. A., Laske, G., Harmon, N. & Shearer, P. M. Seismic imaging of melt in a displaced Hawaiian plume. Nat. Geosci. 6, 657–660 (2013).

    Article  Google Scholar 

  18. Wu, Z. et al. Lateral Structural Variation of the Lithosphere-Asthenosphere System in the Northeastern to Eastern Iranian Plateau and Its Tectonic Implications. J. Geophys. Res. Solid Earth 126, e2020JB020256 (2021).

    Article  Google Scholar 

  19. Hua, J., Fischer, K. M., Wu, M. & Blom, N. A. New approaches to multifrequency Sp stacking tested in the Anatolian region. J. Geophys. Res. Solid Earth 125, e2020JB020313 (2020).

    Article  Google Scholar 

  20. Miller, M. S., O’Driscoll, L. J., Butcher, A. J. & Thomas, C. Imaging Canary Island hotspot material beneath the lithosphere of Morocco and southern Spain. Earth Planet. Sci. Lett. 431, 186–194 (2015).

    Article  Google Scholar 

  21. French, S. W. & Romanowicz, B. Broad plumes rooted at the base of the Earth’s mantle beneath major hotspots. Nature 525, 95–99 (2015).

    Article  Google Scholar 

  22. Lei, W. et al. Global adjoint tomography—model GLAD-M25. Geophys. J. Int. 223, 1–21 (2020).

    Article  Google Scholar 

  23. Karato, S.-i Water distribution across the mantle transition zone and its implications for global material circulation. Earth Planet. Sci. Lett. 301, 413–423 (2011).

    Article  Google Scholar 

  24. Freitas, D. & Manthilake, G. Electrical conductivity of hydrous silicate melts: implications for the bottom-up hydration of Earth’s upper mantle. Earth Planet. Sci. Lett. 523, 115712 (2019).

    Article  Google Scholar 

  25. Pugh, S., Jenkins, J., Boyce, A. & Cottaar, S. Global receiver function observations of the X-discontinuity reveal recycled basalt beneath hotspots. Earth Planet. Sci. Lett. 561, 116813 (2021).

    Article  Google Scholar 

  26. Kawakatsu, H. Unexpected consequences of transverse isotropy. Bull. Seismol. Soc. Am. 111, 129–138 (2021).

    Article  Google Scholar 

  27. Dziewonski, A. M. & Anderson, D. L. Preliminary reference Earth model. Phys. Earth Planet. Inter. 25, 297–356 (1981).

    Article  Google Scholar 

  28. Hirth, G. & Kohlstedt, D. L. Water in the oceanic upper mantle: implications for rheology, melt extraction and the evolution of the lithosphere. Earth Planet. Sci. Lett. 144, 93–108 (1996).

    Article  Google Scholar 

  29. Mierdel, K., Keppler, H., Smyth, J. R. & Langenhorst, F. Water solubility in aluminous orthopyroxene and the origin of Earth’s asthenosphere. Science 315, 364–368 (2007).

    Article  Google Scholar 

  30. Hirschmann, M. M. Water, melting, and the deep Earth H2O cycle. Annu. Rev. Earth Planet. Sci. 34, 629–653 (2006).

    Article  Google Scholar 

  31. Jackson, I. & Faul, U. H. Grainsize-sensitive viscoelastic relaxation in olivine: towards a robust laboratory-based model for seismological application. Phys. Earth Planet. Inter. 183, 151–163 (2010).

    Article  Google Scholar 

  32. Havlin, C., Holtzman, B. K. & Hopper, E. Inference of thermodynamic state in the asthenosphere from anelastic properties, with applications to North American upper mantle. Phys. Earth Planet. Inter. 314, 106639 (2021).

    Article  Google Scholar 

  33. Takei, Y. Constitutive mechanical relations of solid–liquid composites in terms of grain-boundary contiguity. J. Geophys. Res. Solid Earth 103, 18183–18203 (1998).

    Article  Google Scholar 

  34. Hirth, G. & Kohlstedt, D. L. in Inside the Subduction Factory (ed. Eiler, J.) 83–106 (American Geophysical Union, 2003).

  35. Hammond, W. C. & Humphreys, E. D. Upper mantle seismic wave velocity: effects of realistic partial melt geometries. J. Geophys. Res. Solid Earth 105, 10975–10986 (2000).

    Article  Google Scholar 

  36. Lee, C.-T. A., Luffi, P., Plank, T., Dalton, H. & Leeman, W. P. Constraints on the depths and temperatures of basaltic magma generation on Earth and other terrestrial planets using new thermobarometers for mafic magmas. Earth Planet. Sci. Lett. 279, 20–33 (2009).

    Article  Google Scholar 

  37. Hirschmann, M. M. Partial melt in the oceanic low velocity zone. Phys. Earth Planet. Inter. 179, 60–71 (2010).

    Article  Google Scholar 

  38. Dasgupta, R. et al. Carbon-dioxide-rich silicate melt in the Earth’s upper mantle. Nature 493, 211–215 (2013).

    Article  Google Scholar 

  39. Chang, S. J., Ferreira, A. M., Ritsema, J., van Heijst, H. J. & Woodhouse, J. H. Joint inversion for global isotropic and radially anisotropic mantle structure including crustal thickness perturbations. J. Geophys. Res. Solid Earth 120, 4278–4300 (2015).

    Article  Google Scholar 

  40. Ho, T., Priestley, K. & Debayle, E. A global horizontal shear velocity model of the upper mantle from multimode Love wave measurements. Geophys. J. Int. 207, 542–561 (2016).

    Article  Google Scholar 

  41. Moulik, P. & Ekström, G. An anisotropic shear velocity model of the Earth’s mantle using normal modes, body waves, surface waves and long-period waveforms. Geophys. J. Int. 199, 1713–1738 (2014).

    Article  Google Scholar 

  42. Porritt, R. W., Becker, T. W., Boschi, L. & Auer, L. Multiscale, radially anisotropic shear wave imaging of the mantle underneath the contiguous United States through joint inversion of USArray and global data sets. Geophys. J. Int. 226, 1730–1746 (2021).

    Article  Google Scholar 

  43. Becker, T. W. Superweak asthenosphere in light of upper mantle seismic anisotropy. Geochem. Geophys. Geosyst. 18, 1986–2003 (2017).

    Article  Google Scholar 

  44. Ligorría, J. P. & Ammon, C. J. Iterative deconvolution and receiver-function estimation. Bull. Seismol. Soc. Am. 89, 1395–1400 (1999).

    Article  Google Scholar 

  45. Rychert, C. A. & Harmon, N. Stacked P-to-S and S-to-P receiver functions determination of crustal thickness, Vp, and Vs: the H-V stacking method. Geophys. Res. Lett. 43, 1487–1494 (2016).

    Article  Google Scholar 

  46. Hua, J., Fischer, K. M. & Savage, M. K. The lithosphere–asthenosphere boundary beneath the South Island of New Zealand. Earth Planet. Sci. Lett. 484, 92–102 (2018).

    Article  Google Scholar 

  47. Keith, C. M. & Crampin, S. Seismic body waves in anisotropic media: synthetic seismograms. Geophys. J. Int. 49, 225–243 (1977).

    Article  Google Scholar 

  48. Herzberg, C. & Gazel, E. Petrological evidence for secular cooling in mantle plumes. Nature 458, 619–622 (2009).

    Article  Google Scholar 

  49. Behn, M. D., Hirth, G. & Elsenbeck, J. R. II Implications of grain size evolution on the seismic structure of the oceanic upper mantle. Earth Planet. Sci. Lett. 282, 178–189 (2009).

    Article  Google Scholar 

  50. Turner, A. J., Katz, R. F. & Behn, M. D. Grain-size dynamics beneath mid-ocean ridges: implications for permeability and melt extraction. Geochem. Geophys. Geosyst. 16, 925–946 (2015).

    Article  Google Scholar 

  51. Connolly, J. The geodynamic equation of state: what and how. Geochem. Geophys. Geosyst. 10, Q10014 (2009).

    Article  Google Scholar 

  52. Takei, Y. Effect of pore geometry on VP/VS: from equilibrium geometry to crack. J. Geophys. Res. Solid Earth https://doi.org/10.1029/2001JB000522 (2002).

  53. Chantel, J. et al. Experimental evidence supports mantle partial melting in the asthenosphere. Sci. Adv. 2, e1600246 (2016).

    Article  Google Scholar 

  54. Katz, R. F., Spiegelman, M. & Langmuir, C. H. A new parameterization of hydrous mantle melting. Geochem. Geophys. Geosyst. https://doi.org/10.1029/2002GC000433 (2003).

  55. Hirschmann, M. M., Tenner, T., Aubaud, C. & Withers, A. C. Dehydration melting of nominally anhydrous mantle: the primacy of partitioning. Phys. Earth Planet. Inter. 176, 54–68 (2009).

    Article  Google Scholar 

  56. Kaminski, E., Ribe, N. M. & Browaeys, J. T. D-Rex, a program for calculation of seismic anisotropy due to crystal lattice preferred orientation in the convective upper mantle. Geophys. J. Int. 158, 744–752 (2004).

    Article  Google Scholar 

  57. Becker, T. W., Kustowski, B. & Ekström, G. Radial seismic anisotropy as a constraint for upper mantle rheology. Earth Planet. Sci. Lett. 267, 213–227 (2008).

    Article  Google Scholar 

  58. Dalton, C. A., Ekström, G. & Dziewoński, A. M. The global attenuation structure of the upper mantle. J. Geophys. Res. Solid Earth 113, B09303 (2008).

    Article  Google Scholar 

  59. Karaoğlu, H. & Romanowicz, B. Inferring global upper-mantle shear attenuation structure by waveform tomography using the spectral element method. Geophys. J. Int. 213, 1536–1558 (2018).

    Article  Google Scholar 

  60. Adenis, A., Debayle, E. & Ricard, Y. Attenuation tomography of the upper mantle. Geophys. Res. Lett. 44, 7715–7724 (2017).

    Article  Google Scholar 

  61. Courtillot, V., Davaille, A., Besse, J. & Stock, J. Three distinct types of hotspots in the Earth’s mantle. Earth Planet. Sci. Lett. 205, 295–308 (2003).

    Article  Google Scholar 

  62. Richards, M. A., Hager, B. H. & Sleep, N. H. Dynamically supported geoid highs over hotspots: observation and theory. J. Geophys. Res. Solid Earth 93, 7690–7708 (1988).

    Article  Google Scholar 

  63. Steinberger, B. Plumes in a convecting mantle: models and observations for individual hotspots. J. Geophys. Res. Solid Earth 105, 11127–11152 (2000).

    Article  Google Scholar 

  64. Bird, P. An updated digital model of plate boundaries. Geochem. Geophys. Geosyst. 4, 1027 (2003).

    Article  Google Scholar 

  65. Laske, G., Masters, G., Ma, Z. & Pasyanos, M. Update on Crust 1.0: a 1-degree global model of Earth’s crust. Geophys. Res. Abstr. 15, abstr. 2658 (2013).

  66. Hayes, G. P. et al. Slab2, a comprehensive subduction zone geometry model. Science 362, 58–61 (2018).

    Article  Google Scholar 

  67. Yamauchi, H. & Takei, Y. Polycrystal anelasticity at near-solidus temperatures. J. Geophys. Res. Solid Earth 121, 7790–7820 (2016).

    Article  Google Scholar 

  68. Lu, C., Grand, S. P., Lai, H. & Garnero, E. J. TX2019slab: a new P and S tomography model incorporating subducting slabs. J. Geophys. Res. Solid Earth 124, 11549–11567 (2019).

    Article  Google Scholar 

  69. Conrad, C. P. & Behn, M. D. Constraints on lithosphere net rotation and asthenospheric viscosity from global mantle flow models and seismic anisotropy. Geochem. Geophys. Geosys. 11, Q05W05 (2010).

    Article  Google Scholar 

  70. Stixrude, L. & Lithgow-Bertelloni, C. Thermodynamics of mantle minerals–II. Phase equilibria. Geophys. J. Int. 184, 1180–1213 (2011).

    Article  Google Scholar 

  71. McKenzie, D. The generation and compaction of partially molten rock. J. Petrol. 25, 713–765 (1984).

    Article  Google Scholar 

  72. Turcotte, D. L. & Schubert, G. Geodynamics (Cambridge Univ. Press, 2002).

  73. Plank, T. & Forsyth, D. W. Thermal structure and melting conditions in the mantle beneath the Basin and Range province from seismology and petrology. Geochem. Geophys. Geosyst. 17, 1312–1338 (2016).

    Article  Google Scholar 

  74. Herzberg, C. & Asimow, P. D. Petrology of some oceanic island basalts: PRIMELT2.XLS software for primary magma calculation. Geochem. Geophys. Geosyst. https://doi.org/10.1029/2008GC002057 (2008).

  75. Dixon, J. E. & Clague, D. A. Volatiles in basaltic glasses from Loihi Seamount, Hawaii: evidence for a relatively dry plume component. J. Petrol. 42, 627–654 (2001).

    Article  Google Scholar 

  76. Zhang, H. L., Cottrell, E., Solheid, P. A., Kelley, K. A. & Hirschmann, M. M. Determination of Fe3+/ΣFe of XANES basaltic glass standards by Mössbauer spectroscopy and its application to the oxidation state of iron in MORB. Chem. Geol. 479, 166–175 (2018).

    Article  Google Scholar 

  77. Bell, D. R., Rossman, G. R., Maldener, J., Endisch, D. & Rauch, F. Hydroxide in olivine: a quantitative determination of the absolute amount and calibration of the IR spectrum. J. Geophys. Res. Solid Earth 108, 2105 (2003).

    Article  Google Scholar 

  78. Paterson, M. & Olgaard, D. Rock deformation tests to large shear strains in torsion. J. Struct. Geol. 22, 1341–1358 (2000).

    Article  Google Scholar 

  79. Rudge, J. F. The viscosities of partially molten materials undergoing diffusion creep. J. Geophys. Res. Solid Earth 123, 10534–10562 (2018).

    Article  Google Scholar 

  80. Becker, T. W. On the effect of temperature and strain-rate dependent viscosity on global mantle flow, net rotation, and plate-driving forces. Geophys. J. Int. 167, 943–957 (2006).

    Article  Google Scholar 

  81. Turner, A. J., Katz, R. F., Behn, M. D. & Keller, T. Magmatic focusing to mid-ocean ridges: the role of grain-size variability and non-Newtonian viscosity. Geochem. Geophys. Geosyst. 18, 4342–4355 (2017).

    Article  Google Scholar 

  82. Frederiksen, A. & Bostock, M. Modelling teleseismic waves in dipping anisotropic structures. Geophys. J. Int. 141, 401–412 (2000).

    Article  Google Scholar 

  83. Holtzman, B. K. & Kendall, J. M. Organized melt, seismic anisotropy, and plate boundary lubrication. Geochem. Geophys. Geosyst. 11, Q0AB06 (2010).

    Article  Google Scholar 

Download references

Acknowledgements

We thank E. Bozdag for providing the global adjoint tomography model GLAD-M25, B. Romanowicz for the attenuation model SEMUCB-UMQ, M. Behn for his global flow model, C. A. Dalton for assistance with Perple_X and discussions about partial melting, Z. Ma for help with CRUST1.0 and S. P. Grand for discussions about paper structure. This work was supported by the US National Science Foundation grant EAR-1829401 received by K.M.F.

Author information

Authors and Affiliations

Authors

Contributions

J.H. initiated the project and conducted the data analysis, comparison and modelling in the paper. K.M.F. advised on the seismological aspects and overall conclusions of the paper. T.W.B. advised on the geodynamical modelling and seismology results. E.G. advised on its petrological components. G.H. advised on rheological modelling. Interpretation of the results reflects discussions among the authors. The manuscript was written by J.H. with contributions from K.M.F. and other authors.

Corresponding author

Correspondence to Junlin Hua.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Geoscience thanks Geeth Manthilake, Mingming Jiang and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Primary Handling Editor: Louise Hawkins, in collaboration with the Nature Geoscience team.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 Probabilities of the PVG-150 phase.

a. Example of estimating the probability of the PVG-150 phase from the Sp receiver function at station SMAI (Antarctic Peninsula) in network AI. The solid line shows the receiver function amplitude. The background shows the heat map for the 3,000 perturbed receiver functions. The dashed line shows one of the perturbed receiver functions, whose maximum amplitude is below the threshold amplitude for picking of 0.006 (dotted line). In this case, no PVG-150 phase is picked, although the original stacked receiver function has an amplitude much higher than the threshold. Among all perturbed receiver functions, 96.9% produce a picked PVG-150 phase. The other dotted line shows zero amplitude. b. Map of the depth of the PVG-150 phase for those with probabilities over 0.8 (Fig. 1a). c. Similar to b but for the maximum amplitude of the PVG-150 phase as defined in Extended Data Fig. 4d. d. Map of the probability of the PVG-150 phase obtained from weighted averaging of the probability at surrounding stations. e. Map of the uncertainty of the PVG-150 phase probability represented by the weighted standard deviation among surrounding stations. f. Stations (triangles) with similar PVG-150 phase probability relative to neighboring stations indicated by W3 (Eq. 4) values over 0.4. These stations have the strongest contributions to the average mantle profiles in Fig. 4a and Extended Data Fig. 8. Black triangles indicate stations with the PVG-150 (probability over 0.8), and white triangles show stations without the PVG-150 (probability below 0.2). Background shows average VSV at 110–130 km depths22, as in Fig. 1a.

Extended Data Fig. 2 The PVG-260 phase.

a. Map similar to Fig. 1a, but in this case PVGs were picked between 230 and 350 km depths, which is the potential depth range for X-discontinuities25. The background VSV22 is at 260 km depth. Hotspots61,62,63, subduction zones64 and the rim of cratons represented by the 180 LAB contour in ref. 2 are also plotted. b. Observed receiver functions, similar to Fig. 2a, but binned by VSV at 260 km depth. The PVG-260 phase is present across a wider range of VSV values than the PVG-150.

Extended Data Fig. 3 Relationships between receiver functions and mantle properties.

a-b. Observed receiver functions binned by average VSV at 110–130 km depths beneath the station, similar to Fig. 2a, but based on different reference velocity models: (a) SEMUCB-WM1 (ref. 21) (b) CAM2016 (ref. 2). c-d. Similar to Fig. 2a, with the same reference model22, but using only the 644 continental stations (c) or only the 74 oceanic stations (d). A station was designated as oceanic if its crustal type in CRUST1.0 (ref. 65) is normal oceanic, oceans 3 Myr and younger, melt affected ocean and oceanic plateaus, or oceanic plateau with continental crust. e. The same as Fig. 2a, but with a less saturated color scale to show the NVG phase better. f-h. Similar to Fig. 2a and panels c and d, respectively, but filtered with a 6–100 s bandpass filter instead of 10–100 s. i-j. Similar to Figs. 2a and 3c, respectively, but with stations close to subducting slabs removed. A station was designated as near a subducting slab if it is located within 1° of any slab depth that is shallower than 250 km based on Slab2 (ref. 66). k-l. Similar to Fig. 2a,b but plotted with respect to the average VP/VS at 110–130 km depths. m. Similar to Fig. 2b, but only with PVG-150 probabilities; Lines are for all stations combined, oceanic stations only, and continental stations only. The dashed lines in a-k mark the maximum (minimum for k) value with a clear PVG-150 phase; in c-i they are placed at the same VSV as in Fig. 2a in j at the same VS of 4.39 km/s as in Fig. 3c and in a, b and k at 4.29 km/s, 4.27 km/s and 1.763. The color bar at the bottom of the figure pertains to all panels except e, whose color bar is on its right side.

Extended Data Fig. 4 Properties of the PVG and NVG (LAB) phases.

a. Amplitude ratios between PVG-150 and LAB phases, and between LAB and Moho phases as a function of average VSV at 110–130 km depths22. The Moho amplitude is defined as the maximum receiver function amplitude above 80 km depth for each VSV, the LAB amplitude is the maximum negative amplitude between 50 and 150 km depths, and the PVG-150 amplitude is the maximum value between 100 and 200 km depths in Fig. 2a. b. The distribution of picked PVG-150 depths. c. The distribution of picked LAB depths. d. The PVG-150 amplitude distribution. e. The LAB amplitude distribution. f. The distribution of LVZ thickness from stations with both PVG and LAB phase picks. Lines show the median values for the distributions. All the distributions are based on non-cratonic stations (defined as having average VSV at 110–130 km depths below 4.46 km/s).

Extended Data Fig. 5 The effects of waveform frequency content.

Synthetic Sp receiver functions for the velocity models in c were filtered by 6–100 s (a) and 10–100 s (b) bandpass filters before deconvolution. Colors correspond to velocity gradients from the LVZ to the underlying asthenosphere over 10 to 50 km depth ranges (as shown in c). Horizontal dotted lines show the center of VS discontinuities. Vertical dashed lines mark zero amplitude for receiver functions. The vertical dotted-dashed lines in a & b show the median NVG and PVG amplitudes (Extended Data Fig. 4) of the observed phases when filtered over 10–100 s. This analysis shows that the 10–100 s filter receiver functions are less sensitive both to changes in the velocity gradient depth range and to small variations in the absolute depth of the velocity gradient. The 10–100 s filter was therefore chosen to measure globally-averaged PVG-150 properties. d) The effect of a range of filters on observed receiver function stacks at a single station (ANM, AK network), whose NVG and PVG amplitudes are close to the median of all stations (dashed lines). Colors correspond to receiver functions filtered over different period ranges. Filters containing shorter periods show greater separation between the LAB NVG and the PVG-150, providing additional evidence that the PVG-150 is not a side-lobe.

Extended Data Fig. 6 Ray parameter dependence of the Sp receiver functions.

a. Observed Sp receiver functions binned as a function of ray parameter. b. Synthetic Sp receiver functions for a mantle with an isotropic LVZ (LVZ VS lower than the underlying mantle). c. Synthetic Sp receiver functions where the LVZ is radially anisotropic with ξ equal to 1.1 but has no P anisotropy, and its VVoigt is the same as the underlying mantle. d. The LVZ contains P anisotropy (\(\varphi ^{ - 1} = {{{\mathrm{V}}}}_{{{{\mathrm{PH}}}}}^2/{{{\mathrm{V}}}}_{{{{\mathrm{PV}}}}}^2\) is 0.96) but no S anisotropy, and its Voigt average for P-wave velocity is the same as the underlying mantle. e. The LVZ contains both S (ξ equal to 1.1) and P anistropy (φ−1 equal to 1.04, the value suggested by PREM27 when ξ is 1.1), and the Voigt average for S-wave and P-wave velocities are the same as the underlying mantle. All anisotropic layers in c-e have an η of 0.9, consistent with PREM27 when ξ is 1.1.

Extended Data Fig. 7 Binned synthetic receiver functions.

a. Predicted receiver functions similar to Fig. 3b, but produced from velocities that are those in Fig. 3a minus the poroelasticity effect of partial melt33. b. Similar to a, but velocities also include the effects of a small amount of melt on diffusion creep viscosity10 which strongly influences anelasticity where melting occurs. c. Similar to Fig. 3b, but with a grain size of 1 mm (10 mm for Fig. 3b). d. Similar to a, but the anelasticity is based on ref. 67 to include pre-melting effects, and the poroelasticity effect is not included. Black dashed lines are the same as in Fig. 3.

Extended Data Fig. 8 Mantle property contrasts between areas with and without the PVG-150 phase.

The first three rows show differences in weighted averaged (Eq. 4) profiles of VVoigt (a), ξ (b), attenuation (c) and radial mantle flow velocity (d) for stations with (red lines) and without (blue lines) the PVG-150 phase. The fourth row (e) shows differences in weighted averaged profiles of VVoigt and ξ for different tectonic settings. Transparent areas show one standard deviation for the averaged profiles. Black dashed lines are reference values from PREM27. For VVoigt and ξ, columns are for velocity models SEMUCB-WM1 (ref. 21), CAM2016 (refs. 2,40), SVANI-US42, GLAD-M25 (ref. 22), S362ANI + M41 and SGLOBE-rani39. For attenuation, columns show models SEMUCB-UMQ59, QsADR17 (ref. 60) and QRFSI12 (ref. 58). For radial mantle flow velocity, the first column contains modeled radial flow velocity based on the approach in ref. 43 and the tomographic model in ref. 68, and the second column shows the modeled velocity from ref. 69. For tectonic settings70,71,72,73,74,75,76,77,78,79,80,81,82,83, we considered oceanic regions, continental regions, and regions excluding subducting slabs; these tectonic groups are defined the same as in Extended Data Fig. 3.

Extended Data Fig. 9 1D modeling of radial anisotropy assuming different effects of partial melt.

First and second rows illustrate the 1D models of mantle Couette (a) and Poiseuille flow (b) (1,450 °C cases). The first column shows the solution for mantle flow velocity, and the second column is the equivalent mantle viscosity. The 3rd to 5th columns show predicted ξ for different points in time: early in strain accumulation; when radial anisotropy has developed to resemble the observed profiles; and when anisotropy has reached its saturation level over a broad depth range. The time intervals of strain accumulation for these three stages are labelled in the bottom right corners. The sixth column shows the product of the fourth column convolved with a Gaussian whose standard deviation is 30 km to mimic the potential for reduced vertical resolution in tomography due to broad surface wave sensitivity kernels. The seventh column shows the radial anisotropy when grain size is 1 mm (10 mm for other columns); here the plate motion for Couette flow is set to 10 cm/yr (with a different x-axis), and the longer deformation times are labeled. Same as in Fig. 4b, blue lines are for the case when the effects of partial melt on viscosity are ignored; red dashed lines show when partial melt effects are only characterized by a moderate exponential term34; and red solid lines are the case when a factor of 5 viscosity increase is introduced across 150 km depth to represent stronger partial melting impacts on viscosity above the PVG-150.

Extended Data Fig. 10 Modelled 3D radial anisotropy with localized asthenospheric viscosity layering.

Model results in this figure are all from ref. 43. The first row shows the relative mantle viscosity at 134 km depth, and in the corresponding second and third columns, low-viscosity bands parallel or perpendicular to the mid-ocean ridges are present at depths between 100 and 150 km. Beneath 150 km depth, the reference viscosity distribution (first column) is assumed. The second and third rows show predicted ξ structures at 100 and 200 km depths for viscosity structures in the first row. At 100 km depth, radial anisotropy is increased in the low viscosity bands, and at 200 km depth, radial anisotropy is decreased beneath the low viscosity bands.

Supplementary information

Supplementary Information

Supplementary Methods 1–4, Discussion 1–4 and Figs. 1–3.

Supplementary Table 1

Information about the samples we used to calculate primary magma equilibration conditions

Supplementary Data 1

Global 10–100 s Sp receiver functions analysed in this study

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hua, J., Fischer, K.M., Becker, T.W. et al. Asthenospheric low-velocity zone consistent with globally prevalent partial melting. Nat. Geosci. 16, 175–181 (2023). https://doi.org/10.1038/s41561-022-01116-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41561-022-01116-9

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing