Skip to main content
Log in

Fast Computation of Global Sensitivity Kernel Database Based on Spectral-Element Simulations

  • Published:
Pure and Applied Geophysics Aims and scope Submit manuscript

Abstract

Finite-frequency sensitivity kernels, a theoretical improvement from simple infinitely thin ray paths, have been used extensively in recent global and regional tomographic inversions. These sensitivity kernels provide more consistent and accurate interpretation of a growing number of broadband measurements, and are critical in mapping 3D heterogeneous structures of the mantle. Based on Born approximation, the calculation of sensitivity kernels requires the interaction of the forward wavefield and an adjoint wavefield generated by placing adjoint sources at stations. Both fields can be obtained accurately through numerical simulations of seismic wave propagation, particularly important for kernels of phases that cannot be sufficiently described by ray theory (such as core-diffracted waves). However, the total number of forward and adjoint numerical simulations required to build kernels for individual source–receiver pairs and to form the design matrix for classical tomography is computationally unaffordable. In this paper, we take advantage of the symmetry of 1D reference models, perform moment tensor forward and point force adjoint spectral-element simulations, and save six-component strain fields only on the equatorial plane based on the open-source spectral-element simulation package, SPECFEM3D_GLOBE. Sensitivity kernels for seismic phases at any epicentral distance can be efficiently computed by combining forward and adjoint strain wavefields from the saved strain field database, which significantly reduces both the number of simulations and the amount of storage required for global tomographic problems. Based on this technique, we compute traveltime, amplitude and/or boundary kernels of isotropic and radially anisotropic elastic parameters for various (\(P\), \(S\), \(P_{\mathrm{diff}}\), \(S_{\mathrm{diff}}\), depth, surface-reflected, surface wave, S 660 S boundary, etc.) phases for 1D ak135 model, in preparation for future global tomographic inversions.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16

Similar content being viewed by others

References

  • Aki, K., & Richards, P. G. (2002). Quantitative Seismology (2nd ed.). Sausalito: University Science Books.

    Google Scholar 

  • Becker, T. W., Kellogg, J. B., Ekström, G., & O’Connell, R. J. (2003). Comparison of azimuthal seismic anisotropy from surface waves and finite strain from global mantle-circulation models. Geophysical Journal International, 155(2), 696–714. doi:10.1046/j.1365-246X.2003.02085.x.

    Article  Google Scholar 

  • Born, M., & Wolf, E. (1970). Principles of optics: Electromagnetic theory of propagation, interference and diffraction of light (4th ed.). Oxford: Pergamon Press.

    Google Scholar 

  • Capdeville, Y., Chaljub, E., & Montagner, J. P. (2003). Coupling the spectral element method with a modal solution for elastic wave propagation in global Earth models. Geophysical Journal International, 152(1), 34–67.

    Article  Google Scholar 

  • Carrington, L., Komatitsch, D., Laurenzano, M., Tikir, M.M., Michéa, D., Le Goff, N., Snavely, A., & Tromp, J. (2008). High-frequency simulations of global seismic wave propagation using SPECFEM3D-GLOBE on 62K processors. In: Proceedings of the 2008 ACM/IEEE conference on supercomputing, SC ’08 (pp. 60:1–60:11). Piscataway: IEEE Press. doi:10.1145/1413370.1413432

  • Chaljub, E., Komatitsch, D., Vilotte, J. P., Capdeville, Y., Valette, B., & Festa, G. (2007). Spectral-element analysis in seismology. In: Advances in geophysics (Vol. 48, pp. 365–419). Amsterdam: Elsevier

  • Chang, S. J., Ferreira, A. M. G., Ritsema, J., van Heijst, H. J., & Woodhouse, J. H. (2015). Joint inversion for global isotropic and radially anisotropic mantle structure including crustal thickness perturbations. Journal of Geophysical Research: Solid Earth, 120(6), 4278–4300. doi:10.1002/2014JB011824.

    Google Scholar 

  • Chen, M., & Tromp, J. (2007). Theoretical and numerical investigations of global and regional seismic wave propagation in weakly anisotropic earth models. Geophysical Journal International, 168(3), 1130–1152. doi:10.1111/j.1365-246X.2006.03218.x.

    Article  Google Scholar 

  • Chen, M., Huang, H., Yao, H., van der Hilst, R., & Niu, F. (2014). Low wave speed zones in the crust beneath SE Tibet revealed by ambient noise adjoint tomography. Geophysical Research Letters, 41(2), 334–340. doi:10.1002/2013GL058476.

    Article  Google Scholar 

  • Chen, M., Niu, F., Liu, Q., Tromp, J., & Zheng, X. (2015). Multiparameter adjoint tomography of the crust and upper mantle beneath East Asia: 1. Model construction and comparisons. Journal of Geophysical Research: Solid Earth, 120(3), 1762–1786. doi:10.1002/2014JB011638.

    Google Scholar 

  • Chen, P., Jordan, T. H., & Zhao, L. (2007). Full three-dimensional tomography: A comparison between the scattering-integral and adjoint-wavefield methods. Geophysical Journal International, 170(1), 175–181. doi:10.1111/j.1365-246X.2007.03429.x.

    Article  Google Scholar 

  • Chen, P., Jordan, T. H., & Lee, E. J. (2010). Perturbation kernels for generalized seismological data functionals (GSDF): Perturbation kernels for GSDF. Geophysical Journal International, 183(2), 869–883. doi:10.1111/j.1365-246X.2010.04758.x.

    Article  Google Scholar 

  • Colombi, A., Nissen-Meyer, T., Boschi, L., & Giardini, D. (2012). Seismic waveform sensitivity to global boundary topography. Geophysical Journal International, 191(2), 832–848. doi:10.1111/j.1365-246X.2012.05660.x.

    Article  Google Scholar 

  • Cupillard, P., Delavaud, E., Burgos, G., Festa, G., Vilotte, J. P., Capdeville, Y., et al. (2012). RegSEM: A versatile code based on the spectral element method to compute seismic wave propagation at the regional scale: RegSEM: A regional spectral element code. Geophysical Journal International, 188(3), 1203–1220. doi:10.1111/j.1365-246X.2011.05311.x.

    Article  Google Scholar 

  • Dahlen, F. A. (2005). Finite-frequency sensitivity kernels for boundary topography perturbations. Geophysical Journal International, 162(2), 525–540. doi:10.1111/j.1365-246X.2005.02682.x.

    Article  Google Scholar 

  • Dahlen, F. A., & Tromp, J. (1998). Theoretical Global Seismology. Princeton: Princeton University Press.

    Google Scholar 

  • Dahlen, F. A., Hung, S. H., & Nolet, G. (2000). Fréchet kernels for finite-frequency traveltimes—I. Theory. Geophysical Journal International, 141(1), 157–174. doi:10.1046/j.1365-246X.2000.00070.x.

    Article  Google Scholar 

  • de Vos, D., Paulssen, H., & Fichtner, A. (2013). Finite-frequency sensitivity kernels for two-station surface wave measurements. Geophysical Journal International, 194(2), 1042–1049. doi:10.1093/gji/ggt144.

    Article  Google Scholar 

  • Dziewoński, A. M., & Anderson, D. L. (1981). Preliminary reference Earth model. Physics of the Earth and Planetary Interiors, 25(4), 297–356. doi:10.1016/0031-9201(81)90046-7.

    Article  Google Scholar 

  • Ekström, G., & Dziewoński, A. M. (1998). The unique anisotropy of the Pacific upper mantle. Nature, 394(6689), 168–172. doi:10.1038/28148.

    Article  Google Scholar 

  • Favier, N., & Chevrot, S. (2003). Sensitivity kernels for shear wave splitting in transverse isotropic media. Geophysical Journal International, 153(1), 213–228.

    Article  Google Scholar 

  • Fichtner, A. (2009). Full seismic waveform inversion for structural and source parameters. PhD thesis, Ludwig-Maximilians-Universität München, Munich, Germany

  • Fichtner, A., & van Leeuwen, T. (2015). Resolution analysis by random probing. Journal of Geophysical Research: Solid Earth, 120(8), 5549–5573. doi:10.1002/2015JB012106.

    Google Scholar 

  • Fichtner, A., & Villaseñor, A. (2015). Crust and upper mantle of the western Mediterranean—Constraints from full-waveform inversion. Earth and Planetary Science Letters, 428, 52–62. doi:10.1016/j.epsl.2015.07.038.

    Article  Google Scholar 

  • Fichtner, A., Kennett, B. L. N., Igel, H., & Bunge, H. P. (2008). Theoretical background for continental- and global-scale full-waveform inversion in the time–frequency domain. Geophysical Journal International, 175(2), 665–685. doi:10.1111/j.1365-246X.2008.03923.x.

    Article  Google Scholar 

  • Fichtner, A., Igel, H., Bunge, H. P., & Kennett, B. L. (2009). Simulation and inversion of seismic wave propagation on continental scales based on a spectral-element method. Journal of Numerical Analysis, Industrial and Applied Mathematics, 4(1–2), 11–22.

    Google Scholar 

  • Fuji, N., Chevrot, S., Zhao, L., Geller, R. J., & Kawai, K. (2012). Finite-frequency structural sensitivities of short-period compressional body waves: 3-D Fréchet kernels for high frequencies. Geophysical Journal International, 190(1), 522–540. doi:10.1111/j.1365-246X.2012.05495.x.

    Article  Google Scholar 

  • Gokhberg, A., & Fichtner, A. (2016). Full-waveform inversion on heterogeneous HPC systems. Computers & Geosciences, 89, 260–268. doi:10.1016/j.cageo.2015.12.013.

    Article  Google Scholar 

  • Grand, S. P. (1994). Mantle shear structure beneath the Americas and surrounding oceans. Journal of Geophysical Research: Solid Earth, 99(B6), 11591–11621. doi:10.1029/94JB00042.

    Article  Google Scholar 

  • Grand, S. P. (2002). Mantle shear-wave tomography and the fate of subducted slabs. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 360(1800), 2475–2491. doi:10.1098/rsta.2002.1077.

    Article  Google Scholar 

  • Gu, Y. J., Dziewoński, A. M., Su, W., & Ekström, G. (2001). Models of the mantle shear velocity and discontinuities in the pattern of lateral heterogeneities. Journal of Geophysical Research, 106(B6):11169–11199. doi:10.1029/2001JB000340

  • Gu, Y. J., Dziewoński, A. M., & Ekström, G. (2003). Simultaneous inversion for mantle shear velocity and topography of transition zone discontinuities. Geophysical Journal International, 154(2), 559–583. doi:10.1046/j.1365-246X.2003.01967.x.

    Article  Google Scholar 

  • Guttenberg, B. (1960). The shadow of the Earth’s core. Journal of Geophysical Research, 65(3), 1013–1020. doi:10.1029/JZ065i003p01013.

    Article  Google Scholar 

  • Herraiz, M., & Espinosa, A. F. (1987). Coda waves: A review. Pure and Applied Geophysics, 125(4), 499–577. doi:10.1007/BF00879572.

    Article  Google Scholar 

  • Hosseini, K., & Sigloch, K. (2015). Multifrequency measurements of core-diffracted P waves (P\(_{{\rm diff}}\)) for global waveform tomography. Geophysical Journal International, 203(1), 506–521. doi:10.1093/gji/ggv298.

    Article  Google Scholar 

  • Houser, C., Masters, T. G., Shearer, P. M., & Laske, G. (2008). Shear and compressional velocity models of the mantle from cluster analysis of long-period waveforms. Geophysical Journal International, 174(1), 195–212. doi:10.1111/j.1365-246X.2008.03763.x.

    Article  Google Scholar 

  • Hudson, J. A., & Heritage, J. R. (1981). The use of the Born approximation in seismic scattering problems. Geophysical Journal of the Royal Astronomical Society, 66(1), 221–240. doi:10.1111/j.1365-246X.1981.tb05954.x.

    Article  Google Scholar 

  • Hung, S. H., Dahlen, F. A., & Nolet, G. (2000). Fréchet kernels for finite-frequency traveltimes—II. Examples. Geophysical Journal International, 141(1), 175–203. doi:10.1046/j.1365-246X.2000.00072.x.

    Article  Google Scholar 

  • Kárason, H., & van der Hilst, R. D. (2000). Constraints on mantle convection from seismic tomography. In M. A. Richards, R. G. Gordon, & R. D. van der Hilst (Eds.), The History and Dynamics of Global Plate Motion, Geophysical Monograph Series (Vol. 121, pp. 277–288). Washington, D. C.: American Geophysical Union.

    Chapter  Google Scholar 

  • Kennett, B. L. N., Engdahl, E. R., & Buland, R. P. (1995). Constraints on seismic velocities in the Earth from traveltimes. Geophysical Journal International, 122(1), 108–124. doi:10.1111/j.1365-246X.1995.tb03540.x.

    Article  Google Scholar 

  • Komatitsch, D., & Tromp, J. (1999). Introduction to the spectral element method for three-dimensional seismic wave propagation. Geophysical Journal International, 139, 806–822. doi:10.1046/j.1365-246x.1999.00967.x.

    Article  Google Scholar 

  • Komatitsch, D., & Tromp, J. (2002a). Spectral-element simulations of global seismic wave propagation—I. Validation. Geophysical Journal International, 149, 390–412. doi:10.1046/j.1365-246X.2002.01653.x.

    Article  Google Scholar 

  • Komatitsch, D., & Tromp, J. (2002b). Spectral-element simulations of global seismic wave propagation—II. Three-dimensional models, oceans, rotation and self-gravitation. Geophysical Journal International, 150, 303–318. doi:10.1046/j.1365-246X.2002.01716.x.

    Article  Google Scholar 

  • Komatitsch, D., Tsuboi, S., Ji, C., & Tromp, J. (2003). A 14.6 billion degrees of freedom, 5 teraflops, 2.5 terabyte earthquake simulation on the Earth simulator. In: SC ’03 Proceedings of the 2003 ACM/IEEE conference on supercomputing (p. 4). New York: ACM Press. doi:10.1145/1048935.1050155

  • Komatitsch, D., Liu, Q., Tromp, J., Süss, P., Stidham, C., & Shaw, J. H. (2004). Simulations of ground motion in the Los Angeles basin based upon the spectral-element method. Bulletin of the Seismological Society of America, 94(1), 187–206. doi:10.1785/0120030077.

    Article  Google Scholar 

  • Komatitsch, D., Michéa, D., & Erlebacher, G. (2009). Porting a high-order finite-element earthquake modeling application to NVIDIA graphics cards using CUDA. Journal of Parallel and Distributed Computing, 69(5), 451–460. doi:10.1016/j.jpdc.2009.01.006.

    Article  Google Scholar 

  • Komatitsch, D., Erlebacher, G., Göddeke, D., & Michéa, D. (2010). High-order finite-element seismic wave propagation modeling with MPI on a large GPU cluster. Journal of Computational Physics, 229(20), 7692–7714. doi:10.1016/j.jcp.2010.06.024.

    Article  Google Scholar 

  • Komatitsch, D., Vinnik, L. P., & Chevrot, S. (2010). \(SH_{{\rm diff}} - SV_{{\rm diff}}\) splitting in an isotropic Earth. Journal of Geophysical Research, 115(B7). doi:10.1029/2009JB006795.

  • Komatitsch, D., Xie, Z., Bozdağ, E., Sales de Andrade, E., Peter, D., Liu, Q., et al. (2016). Anelastic sensitivity kernels with parsimonious storage for adjoint tomography and full waveform inversion. Geophysical Journal International, 206(3), 1467–1478. doi:10.1093/gji/ggw224.

    Article  Google Scholar 

  • Krischer, L., Megies, T., Barsch, R., Beyreuther, M., Lecocq, T., Caudron, C., & Wassermann, J. (2015). ObsPy: A bridge for seismology into the scientific Python ecosystem. Computational Science & Discovery, 8(1):014003. doi:10.1088/1749-4699/8/1/014003

  • Kustowski, B., Ekström, & G., Dziewoński, A. M. (2008a). Anisotropic shear-wave velocity structure of the Earth’s mantle: A global model. Journal of Geophysical Research, 113(B6):B06306. doi:10.1029/2007JB005169

  • Kustowski, B., Ekström, G., & Dziewoński, A. M. (2008b). The shear-wave velocity structure in the upper mantle beneath Eurasia. Geophysical Journal International, 174(3), 978–992. doi:10.1111/j.1365-246X.2008.03865.x.

    Article  Google Scholar 

  • Lawrence, J. F., & Shearer, P. M. (2006). A global study of transition zone thickness using receiver functions. Journal of Geophysical Research: Solid Earth, 111(B6):B06307. doi:10.1029/2005JB003973

  • Lee, E. J., Chen, P., Jordan, T. H., Maechling, P. B., Denolle, M. A. M., & Beroza, G. C. (2014). Full-3-D tomography for crustal structure in Southern California based on the scattering-integral and the adjoint-wavefield methods. Journal of Geophysical Research: Solid Earth, 119(8), 6421–6451. doi:10.1002/2014JB011346.

    Google Scholar 

  • Lekić, V., Cottaar, S., Dziewoński, A. M., & Romanowicz, B. (2012). Cluster analysis of global lower mantle tomography: A new class of structure and implications for chemical heterogeneity. Earth and Planetary Science Letters, 357–358, 68–77. doi:10.1016/j.epsl.2012.09.014.

    Google Scholar 

  • Li, X. D., & Romanowicz, B. (1996). Global mantle shear velocity model developed using nonlinear asymptotic coupling theory. Journal of Geophysical Research: Solid Earth, 101(B10):22245–22272. doi:10.1029/96JB01306

  • Liu, Q., & Gu, Y. J. (2012). Seismic imaging: From classical to adjoint tomography. Tectonophysics, 566–567, 31–66. doi:10.1016/j.tecto.2012.07.006.

    Article  Google Scholar 

  • Liu, Q., & Tromp, J. (2006). Finite-frequency kernels based on adjoint methods. Bulletin of the Seismological Society of America, 96, 2383–2397. doi:10.1785/0120060041.

    Article  Google Scholar 

  • Liu, Q., & Tromp, J. (2008). Finite-frequency sensitivity kernels for global seismic wave propagation based upon adjoint methods. Geophysical Journal International, 174, 265–286. doi:10.1111/j.1365-246X.2008.03798.x.

    Article  Google Scholar 

  • Loken, C., Gruner, D., Groer, L., Peltier, R., Bunn, N., Craig, M., et al. (2010). SciNet: Lessons learned from building a power-efficient top-20 system and data centre. Journal of Physics: Conference Series, 256(012), 026. doi:10.1088/1742-6596/256/1/012026.

    Google Scholar 

  • Love, A. E. H. (1927). A Treatise on the Mathematical Theory of Elasticity (4th ed.). Cambridge: Cambridge University Press.

    Google Scholar 

  • Luo, Y., & Schuster, G. T. (1991). Wave-equation traveltime inversion. Geophysics, 56(5), 645–653. doi:10.1190/1.1443081.

    Article  Google Scholar 

  • Manners, U. J. (2008). Investigating the structure of the core-mantle boundary region using S and P diffracted waves. Ph.D. dissertation, University of California, San Diego

  • Marquering, H., Dahlen, F. A., & Nolet, G. (1999). Three-dimensional sensitivity kernels for finite-frequency traveltimes: The banana-doughnut paradox. Geophysical Journal International, 137(3), 805–815. doi:10.1046/j.1365-246x.1999.00837.x.

    Article  Google Scholar 

  • Masters, T. G., Laske, G., Bolton, H., & Dziewoński, A. M. (2000). The relative behavior of shear velocity, bulk sound speed, and compressional velocity in the mantle: Implications for chemical and thermal structure. In: Karato, S., Forte, A., Liebermann, R., Masters, T. G., Stixrude, L. (eds.) Geophysical Monograph Series (Vol. 117, pp 63–87). Washington, D. C.: American Geophysical Union

  • Mercerat, E. D., & Nolet, G. (2012). Comparison of ray- and adjoint-based sensitivity kernels for body-wave seismic tomography. Geophysical Research Letters, 39(12). doi:10.1029/2012GL052002

  • Meschede, M., & Romanowicz, B. (2015). Lateral heterogeneity scales in regional and global upper mantle shear velocity models. Geophysical Journal International, 200(2), 1076–1093. doi:10.1093/gji/ggu424.

    Article  Google Scholar 

  • Mégnin, C., & Romanowicz, B. (2000). The three-dimensional shear velocity structure of the mantle from the inversion of body, surface and higher-mode waveforms. Geophysical Journal International, 143(3), 709–728. doi:10.1046/j.1365-246X.2000.00298.x.

    Article  Google Scholar 

  • Montelli, R., Nolet, G., Dahlen, F. A., Masters, T. G., Engdahl, E. R., & Hung, S. H. (2004). Finite-frequency tomography reveals a variety of plumes in the mantle. Science, 303(5656), 338–343. doi:10.1126/science.1092485.

    Article  Google Scholar 

  • Moulik, P., & Ekström, G. (2014). An anisotropic shear velocity model of the Earth’s mantle using normal modes, body waves, surface waves and long-period waveforms. Geophysical Journal International, 199(3), 1713–1738. doi:10.1093/gji/ggu356.

    Article  Google Scholar 

  • Nissen-Meyer, T., Dahlen, F. A., & Fournier, A. (2007a). Spherical-earth Fréchet sensitivity kernels. Geophysical Journal International, 168(3), 1051–1066. doi:10.1111/j.1365-246X.2006.03123.x.

    Article  Google Scholar 

  • Nissen-Meyer, T., Fournier, A., & Dahlen, F. A. (2007b). A two-dimensional spectral-element method for computing spherical-earth seismograms— I. Moment-tensor source. Geophysical Journal International, 168(3), 1067–1092. doi:10.1111/j.1365-246X.2006.03121.x.

    Article  Google Scholar 

  • Nissen-Meyer, T., Fournier, A., & Dahlen, F. A. (2008). A 2-D spectral-element method for computing spherical-earth seismograms—II. Waves in solid–fluid media. Geophysical Journal International, 174(3), 873–888. doi:10.1111/j.1365-246X.2008.03813.x.

    Article  Google Scholar 

  • Nissen-Meyer, T., van Driel, M., Stãhler, S. C., Hosseini, K., Hempel, S., Auer, L., et al. (2014). AxiSEM: Broadband 3-D seismic wavefields in axisymmetric media. Solid Earth, 5(1), 425–445. doi:10.5194/se-5-425-2014.

    Article  Google Scholar 

  • Nolet, G. (2008). A breviary of seismic tomography: Imaging the interior of the earth and sun. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Panning, M. P., & Nolet, G. (2008). Surface wave tomography for azimuthal anisotropy in a strongly reduced parameter space. Geophysical Journal International, 174(2), 629–648. doi:10.1111/j.1365-246X.2008.03833.x.

    Article  Google Scholar 

  • Peter, D., Komatitsch, D., Luo, Y., Martin, R., Le, Goff N., Casarotti, E., et al. (2011). Forward and adjoint simulations of seismic wave propagation on fully unstructured hexahedral meshes: SPECFEM3D Version 2.0 ‘Sesame’. Geophysical Journal International, 186(2), 721–739. doi:10.1111/j.1365-246X.2011.05044.x.

    Article  Google Scholar 

  • Ritsema, J., Deuss, A., van Heijst, H. J., & Woodhouse, J. H. (2011). S40RTS: A degree-40 shear-velocity model for the mantle from new Rayleigh wave dispersion, teleseismic traveltime and normal-mode splitting function measurements. Geophysical Journal International, 184(3), 1223–1236. doi:10.1111/j.1365-246X.2010.04884.x.

    Article  Google Scholar 

  • Sales de Andrade, E., Liu, Q., Ma, Z., Manners, U. J., Lee-Varisco, E., & Masters, G. (2014). An updated one-degree seismic tomographic model based on a sensitivity kernel database. AGU Fall Meeting abstracts-s 33.

  • Schumacher, F., & Friederich, W. (2015). The modularized software package ASKI—Full waveform inversion based on waveform sensitivity kernels utilizing external seismic wave propagation codes, AGU Fall Meeting Abstracts.

  • Shen, Y., Zhang, Z., & Zhao, L. (2008). Component-dependent Fréchet sensitivity kernels and utility of three-component seismic records. Bulletin of the Seismological Society of America, 98(5), 2517–2525. doi:10.1785/0120070283.

    Article  Google Scholar 

  • Sieminski, A., Liu, Q., Trampert, J., & Tromp, J. (2007a). Finite-frequency sensitivity of body waves to anisotropy based upon adjoint methods. Geophysical Journal International, 171(1), 368–389. doi:10.1111/j.1365-246X.2007.03528.x.

    Article  Google Scholar 

  • Sieminski, A., Liu, Q., Trampert, J., & Tromp, J. (2007b). Finite-frequency sensitivity of surface waves to anisotropy based upon adjoint methods. Geophysical Journal International, 168(3), 1153–1174. doi:10.1111/j.1365-246X.2006.03261.x.

    Article  Google Scholar 

  • Sigloch, K., McQuarrie, N., & Nolet, G. (2008). Two-stage subduction history under North America inferred from multiple-frequency tomography. Nature Geoscience, 1(7), 458–462. doi:10.1038/ngeo231.

    Article  Google Scholar 

  • Simmons, N. A., Forte, A. M., Boschi, L., & Grand, S. P. (2010). GyPSuM: A joint tomographic model of mantle density and seismic wave speeds. Journal of Geophysical Research, 115(B12). doi:10.1029/2010JB007631.

  • Tape, C., Liu, Q., Maggi, A., & Tromp, J. (2009). Adjoint tomography of the Southern California crust. Science, 325(5943), 988–992. doi:10.1126/science.1175298.

    Article  Google Scholar 

  • The ObsPy Development Team (2016) ObsPy 1.0.1. doi:10.5281/zenodo.48254

  • Tian, Y., Sigloch, K., & Nolet, G. (2009). Multiple-frequency SH-wave tomography of the western US upper mantle. Geophysical Journal International, 178(3), 1384–1402. doi:10.1111/j.1365-246X.2009.04225.x.

    Article  Google Scholar 

  • To, A., Fukao, Y., & Tsuboi, S. (2011). Evidence for a thick and localized ultra low shear velocity zone at the base of the mantle beneath the central Pacific. Physics of the Earth and Planetary Interiors, 184(3–4), 119–133. doi:10.1016/j.pepi.2010.10.015.

    Article  Google Scholar 

  • Tong, P., Yang, D., & Hua, B. (2011). High accuracy wave simulation—Revised derivation, numerical analysis and testing of a nearly analytic integration discrete method for solving acoustic wave equation. International Journal of Solids and Structures, 48(1), 56–70. doi:10.1016/j.ijsolstr.2010.09.003.

    Article  Google Scholar 

  • Tromp, J., Tape, C., & Liu, Q. (2005). Seismic tomography, adjoint methods, time reversal and banana-doughnut kernels. Geophysical Journal International, 160(1), 195–216. doi:10.1111/j.1365-246X.2004.02453.x.

    Article  Google Scholar 

  • Tromp, J., Komatitsch, D., & Liu, Q. (2008). Spectral-element and adjoint methods in seismology. Communications in Computational Physics, 3(1), 1–32.

    Google Scholar 

  • Valenzuela, R. W., & Wysession, M. E. (1998). Illuminating the base of the mantle with diffracted waves. In M. Gurnis, M. E. Wysession, E. Knittle, & B. A. Buffett (Eds.), Geodynamics series (Vol. 28, pp. 57–71). Washington, D. C.: American Geophysical Union.

    Google Scholar 

  • Wen, L., Silver, P., James, D., & Kuehnel, R. (2001). Seismic evidence for a thermo-chemical boundary at the base of the Earth’s mantle. Earth and Planetary Science Letters, 189(3–4), 141–153. doi:10.1016/S0012-821X(01)00365-X.

    Article  Google Scholar 

  • Yang, H. Y., Zhao, L., & Hung, S. H. (2010). Synthetic seismograms by normal-mode summation: A new derivation and numerical examples: Normal-mode synthetic seismograms. Geophysical Journal International, 183(3), 1613–1632. doi:10.1111/j.1365-246X.2010.04820.x.

    Article  Google Scholar 

  • Yuan, H., & Romanowicz, B. (2010). Lithospheric layering in the North American craton. Nature, 466(7310), 1063–1068. doi:10.1038/nature09332.

    Article  Google Scholar 

  • Zaroli, C., Debayle, E., & Sambridge, M. (2010). Frequency-dependent effects on global S-wave traveltimes: Wavefront-healing, scattering and attenuation: Global multiple-frequency S-wave traveltimes. Geophysical Journal International, 182(2), 1025–1042. doi:10.1111/j.1365-246X.2010.04667.x.

    Article  Google Scholar 

  • Zhao, D. (2001). New advances of seismic tomography and its applications to subduction zones and earthquake fault zones: A review. The Island Arc, 10(1), 68–84. doi:10.1046/j.1440-1738.2001.00291.x.

    Article  Google Scholar 

  • Zhao, L., & Chevrot, S. (2011a). An efficient and flexible approach to the calculation of three-dimensional full-wave Fréchet kernels for seismic tomography—I. Theory. Geophysical Journal International, 185, 922–938. doi:10.1111/j.1365-246X.2011.04983.x.

    Article  Google Scholar 

  • Zhao, L., & Chevrot, S. (2011b). An efficient and flexible approach to the calculation of three-dimensional full-wave Fréchet kernels for seismic tomography—II. Numerical results. Geophysical Journal International, 185(2), 939–954. doi:10.1111/j.1365-246X.2011.04984.x.

    Article  Google Scholar 

  • Zhao, L., Jordan, T. H., & Chapman, C. H. (2000). Three-dimensional Fréchet differential kernels for seismic delay times. Geophysical Journal International, 141(3), 558–576. doi:10.1046/j.1365-246x.2000.00085.x.

    Article  Google Scholar 

  • Zhao, L., Chen, P., & Jordan, T. H. (2006). Strain Green’s tensors, reciprocity, and their applications to seismic source and structure studies. Bulletin of the Seismological Society of America, 96(5), 1753–1763. doi:10.1785/0120050253.

    Article  Google Scholar 

  • Zhou, Y., Nolet, G., Dahlen, F.A., & Laske, G. (2006). Global upper-mantle structure from finite-frequency surface-wave tomography. Journal of Geophysical Research, 111(B4). doi:10.1029/2005JB003677.

  • Zhou, Y., Liu, Q., & Tromp, J. (2011). Surface wave sensitivity: Mode summation versus adjoint SEM. Geophysical Journal International, 187(3), 1560–1576. doi:10.1111/j.1365-246X.2011.05212.x.

    Article  Google Scholar 

  • Zhu, H., Bozdağ, E., Peter, D., & Tromp, J. (2012). Structure of the European upper mantle revealed by adjoint tomography. Nature Geoscience, 5(7), 493–498. doi:10.1038/ngeo1501.

    Article  Google Scholar 

Download references

Acknowledgements

The authors wish to acknowledge the developers of the SPECFEM3D_GLOBE software and ObsPy for their continuing work. Computations were performed on the Sandybridge supercomputer at the SciNet HPC Consortium. SciNet is funded by the Canada Foundation for Innovation under the auspices of Compute Canada; the Government of Ontario; Ontario Research Fund—Research Excellence; and the University of Toronto. The ray paths shown in the figures in Sect. 5 are computed based on ObsPy (Krischer et al. 2015; The ObsPy Development Team 2016). The authors also recognize support from the NSERC G8 Research Councils Initiative on Multilateral Research Funding and the Discovery Grant No. 487237.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Elliott Sales de Andrade.

Appendices

Appendix A: Rotation of Elastic Tensor and Kernels

For any arbitrary point \(\mathbf {x}_\mathrm{p}\) at longitude \(\phi\) and colatitude \(\theta\), the rotation from the global coordinate system \((\hat{x}_1, \hat{x}_2, \hat{x}_3)\) to the local spherical coordinate system \((\hat{\theta }, \hat{\phi }, \hat{r})\) can be accomplished in two steps. First rotate the coordinate system with respect to \(\hat{x}_3\) and then about the new \(\hat{x}_2\), i.e.,

$$\begin{aligned} (\hat{x}_1,\hat{x}_2,\hat{x}_3)&\xrightarrow {\phi \text { angle w.r.t. }\hat{x}_3} (\hat{x}_1',\hat{x}_2',\hat{x}_3') \nonumber \\&\xrightarrow {\theta \text { angle w.r.t. }\hat{x}_2'} (\hat{x}_1'',\hat{x}_2'',\hat{x}_3'') = (\hat{\theta },\hat{\phi },\hat{r}). \end{aligned}$$
(34)

By defining \(C_t = \cos \theta\), \(S_t = \sin \theta\), \(C_\mathrm{p}=\cos \phi\), and \(S_\mathrm{p}=\sin \phi\) for shorter notation, the product of the above two rotation matrices gives the transformation from global to local spherical coordinates

$$\begin{aligned} {\varvec{M}}= {\varvec{M}}_2{\varvec{M}}_1=\begin{bmatrix} C_tC_\mathrm{p}&C_t S_\mathrm{p}&-S_t \\ -S_\mathrm{p}&C_\mathrm{p}&0 \\ S_tC_\mathrm{p}&S_tS_\mathrm{p}&C_t \end{bmatrix}. \end{aligned}$$
(35)

Therefore, the elastic tensor elements \({\mathsf {C}}_{i''j''k''l''}\) in the local spherical coordinate system can be calculated from the elastic tensor elements \({\mathsf {C}}_{ijkl}\) in the global coordinate system as

$$\begin{aligned} {\mathsf {C}}_{i''j''k''l''} = M_{i''i} M_{j''j} M_{k''k} M_{l''l} {\mathsf {C}}_{ijkl}. \end{aligned}$$
(36)

In the local spherical coordinate system, we can rewrite the \({\mathsf {C}}_{ijkl}\) tensor with the Voigt notation, following the notations in (Chen and Tromp 2007, Appendix A.1), as

$$\begin{aligned} {\mathsf {C}}^{\text {Voigt}}_{\text {spherical}}= & {} \begin{bmatrix} {\mathsf {c}}_{\theta \theta \theta \theta }&{\mathsf {c}}_{\theta \theta \phi \phi }&{\mathsf {c}}_{\theta \theta rr}&{\mathsf {c}}_{\theta \theta \phi r}&{\mathsf {c}}_{\theta \theta \theta r}&{\mathsf {c}}_{\theta \theta \theta \phi } \\&{\mathsf {c}}_{\phi \phi \phi \phi }&{\mathsf {c}}_{\phi \phi rr}&{\mathsf {c}}_{\phi \phi \phi r}&{\mathsf {c}}_{\phi \phi \theta r}&{\mathsf {c}}_{\phi \phi \theta \phi } \\&{\mathsf {c}}_{r r r r}&{\mathsf {c}}_{rr\phi r}&{\mathsf {c}}_{rr\theta r}&{\mathsf {c}}_{rr \theta \phi } \\&&{\mathsf {c}}_{\phi r \phi r }&{\mathsf {c}}_{\phi r \theta r}&{\mathsf {c}}_{\phi r \theta \phi }\\&&{\mathsf {c}}_{\theta r \theta r}&{\mathsf {c}}_{\theta r \theta \phi } \\&&&{\mathsf {c}}_{\theta \phi \theta \phi }, \end{bmatrix} \nonumber \\= & {} \begin{bmatrix} {\mathsf {C}}_1&{\mathsf {C}}_2&{\mathsf {C}}_3&{\mathsf {C}}_4&{\mathsf {C}}_5&{\mathsf {C}}_6 \\&{\mathsf {C}}_7&{\mathsf {C}}_8&{\mathsf {C}}_9&{\mathsf {C}}_{10}&{\mathsf {C}}_{11} \\&{\mathsf {C}}_{12}&{\mathsf {C}}_{13}&{\mathsf {C}}_{14}&{\mathsf {C}}_{15} \\&&{\mathsf {C}}_{16}&{\mathsf {C}}_{17}&{\mathsf {C}}_{18} \\&&{\mathsf {C}}_{19}&{\mathsf {C}}_{20} \\&&&{\mathsf {C}}_{21} \end{bmatrix}, \end{aligned}$$
(37)

where \({\mathsf {C}}_I\), \(I=1,\ldots , 21\) represent the 21 independent elements of the elastic tensor in the spherical coordinate system. Note the symmetry of the elastic tensor is assumed for the above expressions and only the upper triangle elements are listed. Also, note that \({\mathsf {C}}_I\)’s are different from C, one of the five independent elastic parameters for transversely isotropic media.

We define the sensitivity kernel for \({\mathsf {C}}_I\) as \(K_I, I=1,\ldots , 21\) (\(K_I\) and \(K_{ijkl}\) can be distinguished easily based on the number of indices). By considering the symmetry in \(K_{ijkl}\), it can be shown that

$$\begin{aligned} K_A&= K_1 + K_2 + K_7 \end{aligned}$$
(38a)
$$\begin{aligned} K_C&= K_{12} \end{aligned}$$
(38b)
$$\begin{aligned} K_F&= K_3 + K_8 \end{aligned}$$
(38c)
$$\begin{aligned} K_L&= K_{16} + K_{19} \end{aligned}$$
(38d)
$$\begin{aligned} K_N&= - 2K_2 + K_{21}. \end{aligned}$$
(38e)

Therefore, for any given point \(\mathbf {x}_\mathrm{p}\) in space, a rotation \({\mathsf {M}}\) about the center of the Earth that transforms it from the global coordinate system to the local spherical coordinate system is first applied to the global elastic tensor kernel \(K_{ijkl}\) to obtain \(K_I, I=1,\ldots\,21\), from which the kernels for the five transversely isotropic elastic parameters can be then constructed based on Eq. 38.

Appendix B: Rotation of an Arbitrary Source–Receiver Path onto the Equator

For an arbitrary source \(\mathbf {x}_\mathrm{s}\) at longitude \(\phi _\mathrm{s}\) and colatitude \(\theta _\mathrm{s}\), and receiver \(\mathbf {x}_\mathrm{r}\) at longitude \(\phi _\mathrm{r}\) and colatitude \(\theta _\mathrm{r}\), we seek rotations that transfer a general source–receiver great-circle path onto the equator. The same set of rotations can then be used to populate the sensitivity kernel values in the original global coordinate system by those extracted for equatorial source–receiver path from the strain database. We first rotate the global \(\hat{x}_1\) axis (at the intersection of prime meridian and the equator) onto the unit-normalized source position \(\hat{\mathbf {x}}_\mathrm{s}\) (\(=\mathbf {x}_\mathrm{s}/\Vert \mathbf {x}_\mathrm{s}\Vert\)) similar to the procedures in Appendix A,

$$\begin{aligned} (\hat{x}_1,\hat{x}_2,\hat{x}_3) \xrightarrow {\phi _\mathrm{s} \text { angle w.r.t. }\hat{x}_3} (\hat{x}'_1,\hat{x}'_2,\hat{x}'_3) \nonumber \\ \xrightarrow {-(90^\circ -\theta _\mathrm{s}) \text { angle w.r.t. }\hat{x}_2'} (\hat{r}_\mathrm{s},\hat{\theta }_\mathrm{s},\hat{\phi }_\mathrm{s}), \end{aligned}$$
(39)

corresponding to two rotation matrices. By assuming \(C_t=\cos \theta _\mathrm{s}\), \(S_t=\sin \theta _\mathrm{s}\), \(C_\mathrm{p}=\cos \phi _\mathrm{s}\), \(S_\mathrm{p}=\sin \phi _\mathrm{s}\), the product of these two rotations gives the rotation matrix from the global to a local spherical coordinates such that the unit-normalized source position \(\hat{\mathbf {x}}_\mathrm{s}\) becomes the new \(\hat{x}_1\),

$$\begin{aligned} {\varvec{R}}_\mathrm{s} = {\varvec{R}}_2 {\varvec{R}}_1 = \begin{bmatrix} S_tC_\mathrm{p}&S_t S_\mathrm{p}&C_t \\ -S_\mathrm{p}&C_\mathrm{p}&0 \\ -C_tC_\mathrm{p}&-C_tS_\mathrm{p}&S_t \end{bmatrix}. \end{aligned}$$
(40)

Thus, by assuming \(C_\mathrm{r}=\cos \theta _\mathrm{r}\), \(S_\mathrm{r}=\sin \theta _\mathrm{r}\), \(C_q=\cos \phi _\mathrm{r}\), and \(S_q=\sin \phi _\mathrm{r}\), \(C_{q-p}=\cos (\phi _\mathrm{r}-\phi _\mathrm{s})\), and \(S_{q-p}=\sin (\phi _\mathrm{r}-\phi _\mathrm{s})\), the unit-normalized receiver position \(\hat{\mathbf {x}}_\mathrm{r}\) after this rotation will have coordinates,

$$\begin{aligned} {\varvec{R}}_\mathrm{s} \hat{\mathbf {x}}_\mathrm{r} = \begin{bmatrix} S_tC_\mathrm{p}&S_t S_\mathrm{p}&C_t \\ -S_\mathrm{p}&C_\mathrm{p}&0 \\ -C_tC_\mathrm{p}&-C_tS_\mathrm{p}&S_t \end{bmatrix} \begin{bmatrix} S_\mathrm{r}C_q \\ S_\mathrm{r} S_q \\ C_\mathrm{r} \end{bmatrix}=\begin{bmatrix} C_t C_\mathrm{r} + S_t S_\mathrm{r} C_{q-p} \\ S_\mathrm{r} S_{q-p} \\ S_t C_\mathrm{r} - C_t S_\mathrm{r} C_{q-p} \end{bmatrix}. \end{aligned}$$
(41)

We can then rotate this new coordinate system with respect to \(\hat{r}_\mathrm{s}\) by an angle \(\Theta\) such that the receiver ends up on the equator, similar to those in Appendices C and D,

$$\begin{aligned} {\varvec{R}}_\mathrm{r}=\begin{bmatrix} 1&0&0 \\ 0&\cos \Theta&\sin \Theta \\ 0&-\sin \Theta&\cos \Theta \end{bmatrix}, \end{aligned}$$
(42)

which requires \(\Theta ={\text {arctan2}}(S_t C_\mathrm{r} - C_t S_\mathrm{r} C_{q-p}, S_\mathrm{r} S_{q-p})\). By assuming \(C_\theta =\cos \Theta\), \(S_\theta =\sin \Theta\), the full rotation matrix that needs to be applied to the original global coordinate becomes

$$\begin{aligned} {\varvec{R}}={\varvec{R}}_\mathrm{r}{\varvec{R}}_\mathrm{s}=\begin{bmatrix} S_tC_\mathrm{p}&S_t S_\mathrm{p}&C_t\\ -C_\theta S_\mathrm{p} - S_\theta C_t C_\mathrm{p}&C_\theta C_\mathrm{p} - S_\theta C_t S_\mathrm{p}&S_\theta S_t \\ -C_\theta C_t C_\mathrm{p} + S_\theta S_\mathrm{p}&-C_\theta C_t S_\mathrm{p} - S_\theta C_\mathrm{p}&C_\theta S_t \end{bmatrix}. \end{aligned}$$
(43)

To compute the kernel values on any grid point \(\mathbf {x}_\mathrm{p}\) for a general source–receiver pair, its coordinates in the rotated coordinate system, \(\mathbf {x}_\mathrm{p}^{\text {new}}\), with source–receiver along the equator can be calculated by multiplying the \({\varvec{R}}\) matrix. Hence, the kernel value at point \(\mathbf {x}_\mathrm{p}\) in the original coordinate system is the same as the kernel value at \(\mathbf {x}_\mathrm{p}^{\text {new}}\) for the equatorial source–receiver path which can be extracted and interpolated from our existing strain database.

Appendix C: Rotation of Moment Tensor Strain Field

We derive the set of rotations used in Sect. 3.1 to transform from the original coordinate system to a coordinate system aligned with the best double-couple orientation of a general moment tensor source, and then rotate the arbitrary point \(\mathbf {x}_\mathrm{p}\) to the equatorial plane. We assume a general moment tensor \({\mathsf {M}}\) is placed at the surface, and transform the original coordinate system such that the moment tensor becomes diagonal,

$$\begin{aligned} ({\mathsf {M}})^{\varvec{A}}= {\varvec{A}}{\mathsf {M}}{\varvec{A}}^\mathrm{T} = \begin{bmatrix} M_1&\\&M_2&\\&&M_3 \end{bmatrix}, \end{aligned}$$
(44)

where \(M_2 \le M_1 \le M_3\) are the eigenvalues of the symmetric moment tensor. As most earthquakes have predominantly double-couple source mechanisms (Aki and Richards 2002), it may be advantageous to further rotate the coordinate system such that the new \(\hat{x}_2\) and \(\hat{x}_3\) axes align with the fault plane normal \(\hat{n}\) and slip direction \(\hat{D}\) of the dominant double-couple component of the source (Fig. 17a),

$$\begin{aligned} ({\mathsf {M}})^{{\varvec{A}}{\varvec{B}}} = {\varvec{B}}{\varvec{A}}{\mathsf {M}}{\varvec{A}}^\mathrm{T} {\varvec{B}}^\mathrm{T}, \end{aligned}$$
(45)

where

$$\begin{aligned} {\varvec{B}}=\frac{1}{\sqrt{2}}\begin{bmatrix} \sqrt{2}&0&0 \\ 0&1&1 \\ 0&1&-1 \end{bmatrix} \end{aligned}$$
(46)

and

$$\begin{aligned} ({\mathsf {M}})^{{\varvec{A}}{\varvec{B}}} = \begin{bmatrix} M_1&0&0 \\ 0&(M_2+M_3)/2&(M_2-M_3)/2 \\ 0&(M_2-M_3)/2&(M_2+M_3)/2 \end{bmatrix}. \end{aligned}$$
(47)

For pure double-couple sources, we have \(M_1=0\), \(M_3 = -M_2 = M_\mathrm{s}\) where \(M_\mathrm{s}\) is the scalar moment of the moment tensor \({\mathsf {M}}\), and the moment tensor is reduced to

$$\begin{aligned} ({\mathsf {M}})^{{\varvec{A}}{\varvec{B}}} = M_\mathrm{s} \begin{bmatrix} 0&0&0 \\ 0&0&1 \\ 0&1&0 \end{bmatrix}. \end{aligned}$$
(48)

After the rotations \({\varvec{A}}\) and \({\varvec{B}}\), the original point \(\mathbf {x}_\mathrm{p}\) has been transformed to \((\mathbf {x}_\mathrm{p})^{{\varvec{A}}{\varvec{B}}}={\varvec{B}}{\varvec{A}}\mathbf {x}_\mathrm{p}\). Based on symmetry of the 1D background model, we plan to only store strain fields on the equatorial plane. Hence, we further transform \((\mathbf {x}_\mathrm{p})^{{\varvec{A}}{\varvec{B}}}\) through a rotation \({\varvec{C}}_\mathrm{s}\) with respect to the \((\hat{x}_1)^{{\varvec{A}}{\varvec{B}}}\) axis with an angle \(\Phi _\mathrm{s}\) such that it ends on the equatorial plane (shaded areas in Fig. 17b), i.e., \((\mathbf {x}_\mathrm{p})^{{\varvec{A}}{\varvec{B}}{\varvec{C}}_\mathrm{s}} = r_\mathrm{p}[\cos \Delta _\mathrm{s},\sin \Delta _\mathrm{s},0]^\mathrm{T}\), where if we define \((\mathbf {x}_\mathrm{p})^{{\varvec{A}}{\varvec{B}}} = [x_1,x_2,x_3]\), \(r_\mathrm{p}=(x_1^2+x_2^2+x_3^2)^{1/2}\) is the radius of point \(\mathbf {x}_\mathrm{p}\), and \(\Delta _\mathrm{s}\) is the epicentral distance of \(\mathbf {x}_\mathrm{p}\) from the source. It can be shown that the epicentral distance \(\Delta _\mathrm{s}\) and the rotation angle \(\Phi _\mathrm{s}\) can be computed from

$$\begin{aligned} \cos \Delta _\mathrm{s}&= x_1/r_\mathrm{p}, \end{aligned}$$
(49)
$$\begin{aligned} \cos \Phi _\mathrm{s}&= x_2/(r_\mathrm{p}\sin \Delta _\mathrm{s}),\end{aligned}$$
(50)
$$\begin{aligned} \sin \Phi _\mathrm{s}&= x_3/(r_\mathrm{p}\sin \Delta _\mathrm{s}),\quad \Delta _\mathrm{s} \in [0,\pi ] \end{aligned}$$
(51)

and the rotation matrix is given by

$$\begin{aligned} {\varvec{C}}_\mathrm{s}=\begin{bmatrix} 1&0&0 \\ 0&\cos \Phi _\mathrm{s}&\sin \Phi _\mathrm{s} \\ 0&-\sin \Phi _\mathrm{s}&\cos \Phi _\mathrm{s} \end{bmatrix}. \end{aligned}$$
(52)

Under this transformation the moment tensor source becomes

$$\begin{aligned}&({\mathsf {M}})^{{\varvec{A}}{\varvec{B}}{\varvec{C}}_\mathrm{s}} \nonumber \\&\qquad = {\varvec{C}}_\mathrm{s}({\mathsf {M}})^{{\varvec{A}}{\varvec{B}}} {\varvec{C}}^\mathrm{T}_\mathrm{s} \nonumber \\&\qquad = \begin{bmatrix} M_1&0&0 \\ 0&\frac{M_2+M_3}{2}+\frac{M_2-M_3}{2}\sin 2\Phi _\mathrm{s}&\frac{M_2-M_3}{2}\cos 2\Phi _\mathrm{s} \\ 0&\frac{M_2-M_3}{2}\cos 2\Phi _\mathrm{s}&\frac{M_2+M_3}{2}-\frac{M_2-M_3}{2}\sin 2\Phi _\mathrm{s} \end{bmatrix} \end{aligned}$$
(53)

as in Eq. 14.

Fig. 17
figure 17

Coordinate transformations applied to compute the strain field at point \(\mathbf {x}_\mathrm{p}\) due to a general point moment tensor on the surface (the parentheses used in coordinate transformation is omitted to reduce clutter): a The original coordinate system (black) is rotated by \({\varvec{A}}\) as in Eq. 44 and \({\varvec{B}}\) as in Eq. 46 such that the moment tensor aligns with the fault plane normal \(\hat{n}\) and the slip direction \(\hat{\mathbf {D}}\), producing the \((\hat{x}_i)^{{\varvec{A}}{\varvec{B}}}\) coordinate system (blue). b The \((\hat{x}_i)^{{\varvec{A}}{\varvec{B}}}\) coordinate system is rotated about the \((\hat{x}_1)^{{\varvec{A}}{\varvec{B}}}\)-axis by angle \(\Phi _\mathrm{s}\) such that the point of interest \((\mathbf {x}_\mathrm{p})^{{\varvec{A}}{\varvec{B}}}\) now exists on the equatorial plane (shaded area) as \((\mathbf {x}_\mathrm{p})^{{\varvec{A}}{\varvec{B}}{\varvec{C}}_\mathrm{s}}\). In the new \((\hat{x}_i)^{{\varvec{A}}{\varvec{B}}{\varvec{C}}_\mathrm{s}}\) coordinate system, the epicentral distance of the point \((\mathbf {x}_\mathrm{p})^{{\varvec{A}}{\varvec{B}}{\varvec{C}}_\mathrm{s}}\) from the source is \(\Delta _\mathrm{s}\)

Appendix D: Rotation of Point Force Strain Field

We derive the set of rotations used in Sect. 3.2 to transform the receiver onto the prime meridian and rotate the arbitrary point \(\mathbf {x}_\mathrm{p}\) to the equatorial plane. We first transform the original coordinate system such that the receiver \(\mathbf {x}_\mathrm{r}\) falls on the new \(\hat{x}_1\) axis (Fig. 18a) through a rotation matrix

$$\begin{aligned} {\varvec{D}}=\begin{bmatrix} \cos \Psi&\sin \Psi&0 \\ -\sin \Psi&\cos \Psi&0 \\ 0&0&1 \end{bmatrix}, \end{aligned}$$
(54)

where \(\Psi\) is the epicentral distance of the receiver from the source. We further transform the \((\mathbf {x}_\mathrm{p})^{\varvec{D}}\) point onto the equatorial plane (Fig. 18b) through rotation matrix \({\varvec{C}}_\mathrm{r}\) similar to Eqs. 51 and 52 in Appendix C,

$$\begin{aligned} {\varvec{C}}_\mathrm{r}=\begin{bmatrix} 1&0&0 \\ 0&\cos \Phi _\mathrm{r}&\sin \Phi _\mathrm{r} \\ 0&-\sin \Phi _\mathrm{r}&\cos \Phi _\mathrm{r} \end{bmatrix}, \end{aligned}$$
(55)

where

$$\begin{aligned} \cos \Delta _\mathrm{r}&= x_1/r_\mathrm{p}, \end{aligned}$$
(56)
$$\begin{aligned} \cos \Phi _\mathrm{r}&= x_2/(r_\mathrm{p}\sin \Delta _\mathrm{r}), \end{aligned}$$
(57)
$$\begin{aligned} \sin \Phi _\mathrm{s}&= x_3/(r_\mathrm{p}\sin \Delta _\mathrm{r}), \quad \Delta _\mathrm{r} \in [0,\pi ], \end{aligned}$$
(58)

\(\Delta _\mathrm{r}\) is the great-circle distance between \(\mathbf {x}_\mathrm{p}\) and the receiver \(\mathbf {x}_\mathrm{r}\), and \(\Phi _\mathrm{r}\) is the rotation angle about \((\hat{x}_1)^{\varvec{D}}\) axis (where the receiver resides) to transform \((\mathbf {x}_\mathrm{p})^{\varvec{D}}\) onto the equatorial plane.

Fig. 18
figure 18

The rotations applied to calculate the strain field due to a point force on the surface: a The original coordinate system (black) is rotated by \({\varvec{D}}\) as in Eq. 54 (i.e., \(\Psi\) angle about the \(\hat{x}_3\)-axis) such that the receiver at \(\mathbf {x}_\mathrm{r}\) coincides with the \((\hat{x}_1)^{\varvec{D}}\)-axis. b The \((\hat{x}_i)^{\varvec{D}}\) coordinate system (blue) is rotated about the \((\hat{x}_1)^{\varvec{D}}\)-axis by angle \(\Phi _\mathrm{r}\) such that the point of interest \((\mathbf {x}_\mathrm{p})^{\varvec{D}}\) now exists on the equatorial plane (shaded area) as \((\mathbf {x}_\mathrm{p})^{{\varvec{D}}{\varvec{C}}_\mathrm{r}}\). In the new \((\hat{x}_i)^{{\varvec{D}}{\varvec{C}}_\mathrm{r}}\) coordinate system, the great-circle distance of the point \((\mathbf {x}_\mathrm{p})^{{\varvec{D}}{\varvec{C}}_\mathrm{r}}\) from the receiver \(\mathbf {x}_\mathrm{r}\) is \(\Delta _\mathrm{r}\)

Appendix E: Algorithm Summary

Here we summarize the algorithms we proposed in Sect. 3.3.

  1. 1.

    Choose the cut-off period the Gaussian source time function, h for numerical simulations (e.g., 16 s), and the cut-off period l (e.g., 20 s) with which kernels can be constructed such that \(2l^2>3h^2\).

  2. 2.

    Compute the forward strain fields \(({\mathsf {\varepsilon }})^{\text {equator}}[t,r,\Delta ; {\mathsf {M}}^q,d,h]\) for a series of depth \(d=d_{\text {min}},\ldots , d_{\text {max}}\) and at every depth four fundamental mechanisms \({\mathsf {M}}^q\), \(q=1,\ldots ,4\), as well as the adjoint strain fields \(({\mathsf {\varepsilon }})^{\text {equator}}[t,r,\Delta ;\hat{x}_i, h]\) for three unit force vectors \(\hat{x}_i\), \(i=1,2,3\), based on 3D SEM simulations. Store both the strain and velocity fields only on the equatorial plane.

  3. 3.

    Loop over spatial points in the 3D volume. For each point \(\mathbf {x}_\mathrm{p}\), compute the kernel values based on the following steps:

    • Calculate rotation \({\varvec{A}}\), \({\varvec{B}}\) and \({\varvec{C}}_\mathrm{s}\) to transform the general moment tensor into a coordinate system associated with the best double-couple (\(\hat{n}\), \(\hat{D}\)) directions and then further rotate \(\mathbf {x}_\mathrm{p}\) onto the equatorial plane.

    • For the moment tensor forward strain field, extract the strain values at radius \(r_\mathrm{p}\) and epicentral distance \(\Delta _\mathrm{s}\) for the four fundamental mechanisms, \(({\mathsf {\varepsilon }})^{{\text {equator}}}[t,r_\mathrm{p},\Delta _\mathrm{s}; {\mathsf {M}}^q,d,h]\), \(q=1,\ldots , 4\), from the precomputed strain database. Compute the strain time series at \(\mathbf {x}_\mathrm{p}\) due to an arbitrary moment tensor source \({\mathsf {M}}\) based on the superposition in Eq. 16. Similarly, construct the corresponding velocity time series \(\dot{s}(t,\Psi ;{\mathsf {M}},g_\mathrm{h})\) at the receiver location.

    • For the adjoint strain field, first compute the rotation matrix \({\varvec{D}}\) and \({\varvec{C}}_\mathrm{r}\) to align the receiver with (0\(^{\circ }\) latitude, 0\(^{\circ }\) longitude) and then transform \(\mathbf {x}_\mathrm{p}\) onto the equatorial plane. Compute the corresponding components \(f_i'\) of the transformed force vector \((\mathbf {f})^{{\varvec{D}}{\varvec{C}}_\mathrm{r}}\). Extract the stored strain time series due to individual unit forces and sum them up as in Eq. 19 to obtain the adjoint strain field \(({\mathsf {\varepsilon }}^\dagger )^{{\varvec{D}}{\varvec{C}}_\mathrm{r}}\). Transform it back to obtain the adjoint strain time series at \(\mathbf {x}_\mathrm{p}\) due to an arbitrary force vector \(\mathbf {f}\) in the original coordinate system based on Eq. 20.

    • Convolve the two strain time series calculated from the last two steps. Window the velocity time series at the specific phase we are interested in, normalize it by \(N_t\), and convolve it with a Gaussian function with cut-off period of \(\sqrt{2l^2-3h^2}\) (e.g., 5.6 s). Further convolve these two resulting series and Fourier transform it back to the time domain.

    • The kernel values for the elastic tensor elements for point \(\mathbf {x}_\mathrm{p}\) is given by the values of the transformed time series at time \(t=T\) (i.e., the end of the simulation).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sales de Andrade, E., Liu, Q. Fast Computation of Global Sensitivity Kernel Database Based on Spectral-Element Simulations. Pure Appl. Geophys. 174, 2733–2761 (2017). https://doi.org/10.1007/s00024-017-1573-3

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00024-017-1573-3

Keywords

Navigation