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:

Correlation of porosity variations and rheological transitions on the southern Cascadia megathrust

Abstract

The unknown onshore extent of megathrust earthquake rupture in the Cascadia subduction zone represents a key uncertainty in earthquake hazard for the Pacific Northwest that is governed by the physical state and mechanical properties of the plate interface. The Cascadia plate interface is segmented into an interseismically locked zone located primarily offshore that is expected to rupture in large earthquakes, a region of aseismic slow slip at greater depth, and an intervening transition zone of uncertain rupture potential. Here we image the evolution of the ratio of seismic compressional to shear wave velocities from the locked zone to the transition zone, which is related to changes in fluid content of the plate boundary zone, using a dense onshore–offshore seismic dataset from southernmost Cascadia. The locked zone shows evidence of high fluid content implying a high porosity, yet the downdip transition zone shows an order of magnitude lower porosity. This strong variation is consistent with models that contain a ductile region between the earthquake rupture and slow slip zones that would inhibit onshore propagation of future large earthquake ruptures and hence reduce seismic hazard.

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: Tectonic background of the Cascadia subduction zone and map view of seismic events and stations in our study area.
Fig. 2: Cross-sections (AA′–DD′) of the Vp/Vs models from the synthetic test and real data inversion.
Fig. 3: Downdip variations in interseismic coupling and Vp/Vs.
Fig. 4: Seismic constraints on the stress state of the locked plate interface.
Fig. 5: Summary schematic showing observations and interpretations.

Similar content being viewed by others

Data availability

The CI and XQ seismic waveform data used for this study are available at the IRIS Data Management Center (https://doi.org/10.7914/SN/7D_2011 and https://doi.org/10.7914/SN/XQ_2007). The NCEDC seismic waveform data used for this study are available via the NCEDC official website http://ncedc.org.

Code availability

The seismic tomography software package tomoTD is available upon request.

References

  1. Wang, K. & Tréhu, A. M. Invited review paper: some outstanding issues in the study of great megathrust earthquakes—the Cascadia example. J. Geodyn. 98, 1–18 (2016).

    Article  Google Scholar 

  2. Hyndman, R. D. & Wang, K. The rupture zone of Cascadia great earthquakes from current deformation and the thermal regime. J. Geophys. Res. 100, 22133–22154 (1995).

    Article  Google Scholar 

  3. Hyndman, R. D. Downdip landward limit of Cascadia great earthquake rupture. J. Geophys. Res. 118, 5530–5549 (2013).

    Article  Google Scholar 

  4. McCaffrey, R., King, R. W., Payne, S. J. & Lancaster, M. Active tectonics of northwestern US inferred from GPS‐derived surface velocities. J. Geophys. Res. 118, 709–723 (2013).

    Article  Google Scholar 

  5. Schmalzle, G. M., McCaffrey, R. & Creager, K. C. Central Cascadia subduction zone creep. Geochem. Geophys. Geosyst. 15, 1515–1532 (2014).

    Article  Google Scholar 

  6. Pollitz, F. F. & Evans, E. L. Implications of the earthquake cycle for inferring fault locking on the Cascadia megathrust. Geophys. J. Int. 209, 167–185 (2017).

    Google Scholar 

  7. Bruhat, L. & Segall, P. Coupling on the northern Cascadia subduction zone from geodetic measurements and physics‐based models. J. Geophys. Res. 121, 8297–8314 (2016).

    Article  Google Scholar 

  8. Rogers, G. & Dragert, H. Episodic tremor and slip on the Cascadia subduction zone: the chatter of silent slip. Science 300, 1942–1943 (2003).

    Article  Google Scholar 

  9. Wech, A. G. & Creager, K. C. A continuum of stress, strength and slip in the Cascadia subduction zone. Nat. Geosci. 4, 624–628 (2011).

    Article  Google Scholar 

  10. Brown, J. R. et al. Deep low‐frequency earthquakes in tremor localize to the plate interface in multiple subduction zones. Geophys. Res. Lett. 36, L19306 (2009).

    Article  Google Scholar 

  11. Bostock, M. G., Royer, A. A., Hearn, E. H. & Peacock, S. M. Low frequency earthquakes below southern Vancouver Island. Geochem. Geophys. Geosyst. 13, Q11007 (2012).

    Article  Google Scholar 

  12. Petersen, M. D. et al. Documentation for the 2014 Update of the United States National Seismic Hazard Maps (US Geological Survey, 2014); https://doi.org/10.3133/ofr20141091

  13. Obara, K. & Kato, A. Connecting slow earthquakes to huge earthquakes. Science 353, 253–257 (2016).

    Article  Google Scholar 

  14. Nadeau, R. M. & Dolenc, D. Nonvolcanic tremors deep beneath the San Andreas Fault. Science 307, 389 (2005).

    Article  Google Scholar 

  15. Liu, Y. Numerical simulations on megathrust rupture stabilized under strong dilatancy strengthening in slow slip region. Geophys. Res. Lett. 40, 1311–1316 (2013).

    Article  Google Scholar 

  16. Gao, X. & Wang, K. Rheological separation of the megathrust seismogenic zone and episodic tremor and slip. Nature 543, 416–419 (2017).

    Article  Google Scholar 

  17. Tréhu, A. M., Braunmiller, J. & Davis, E. Seismicity of the central Cascadia continental margin near 44.5° N: a decadal view. Seismol. Res. Lett. 86, 819–829 (2015).

    Article  Google Scholar 

  18. Audet, P., Bostock, M. G., Christensen, N. I. & Peacock, S. M. Seismic evidence for overpressured subducted oceanic crust and megathrust fault sealing. Nature 457, 76–78 (2009).

    Article  Google Scholar 

  19. Abers, G. A. et al. Imaging the source region of Cascadia tremor and intermediate-depth earthquakes. Geology 37, 1119–1122 (2009).

    Article  Google Scholar 

  20. Peacock, S. M., Christensen, N. I., Bostock, M. G. & Audet, P. High pore pressures and porosity at 35 km depth in the Cascadia subduction zone. Geology 39, 471–474 (2011).

    Article  Google Scholar 

  21. Tauzin, B., Reynard, B., Perrillat, J. P., Debayle, E. & Bodin, T. Deep crustal fracture zones control fluid escape and the seismic cycle in the Cascadia subduction zone. Earth Planet. Sci. Lett. 460, 1–11 (2017).

    Article  Google Scholar 

  22. Audet, P., Bostock, M. G., Boyarko, D. C., Brudzinski, M. R. & Allen, R. M. Slab morphology in the Cascadia fore arc and its relation to episodic tremor and slip. J. Geophys. Res. 115, B00A16 (2010).

    Google Scholar 

  23. McGary, R. S., Evans, R. L., Wannamaker, P. E., Elsenbeck, J. & Rondenay, S. Pathway from subducting slab to surface for melt and fluids beneath Mount Rainier. Nature 511, 338–340 (2014).

    Article  Google Scholar 

  24. Evans, R. L., Wannamaker, P. E., McGary, R. S. & Elsenbeck, J. Electrical structure of the central Cascadia subduction zone: the EMSLAB Lincoln Line revisited. Earth Planet. Sci. Lett. 402, 265–274 (2014).

    Article  Google Scholar 

  25. Wannamaker, P. E. et al. Segmentation of plate coupling, fate of subduction fluids, and modes of arc magmatism in Cascadia, inferred from magnetotelluric resistivity. Geochem. Geophys. Geosyst. 15, 4230–4253 (2014).

    Article  Google Scholar 

  26. Audet, P. & Schaeffer, A. J. Fluid pressure and shear zone development over the locked to slow slip region in Cascadia. Sci. Adv. 4, eaar2982 (2018).

    Article  Google Scholar 

  27. Janiszewski, H. A. & Abers, G. A. Imaging the plate interface in the Cascadia seismogenic zone: new constraints from offshore receiver functions. Seismol. Res. Lett. 86, 1261–1269 (2015).

    Article  Google Scholar 

  28. Guo, H. & Zhang, H. Development of double-pair double difference earthquake location algorithm for improving earthquake locations. Geophys. J. Int. 208, 333–348 (2017).

    Article  Google Scholar 

  29. Guo, H., Zhang, H., Nadeau, R. M. & Peng, Z. High‐resolution deep tectonic tremor locations beneath the San Andreas Fault near Cholame, California, using the double-pair double-difference location method. J. Geophys. Res. 122, 3062–3075 (2017).

    Article  Google Scholar 

  30. McCrory, P. A., Blair, J. L., Waldhauser, F. & Oppenheimer, D. H. Juan de Fuca slab geometry and its relation to Wadati-Benioff zone seismicity. J. Geophys. Res. 117, B09306 (2012).

    Google Scholar 

  31. Beaudoin, B. C., Hole, J. A., Klemperer, S. L. & Tréhu, A. M. Location of the southern edge of the Gorda slab and evidence for an adjacent asthenospheric window: results from seismic profiling and gravity. J. Geophys. Res. 103, 30101–30115 (1998).

    Article  Google Scholar 

  32. Gulick, S. P., Meltzer, A. M. & Clarke, S. H. Jr Seismic structure of the southern Cascadia subduction zone and accretionary prism north of the Mendocino triple junction. J. Geophys. Res. 103, 27207–27222 (1998).

    Article  Google Scholar 

  33. den Hartog, S. A. & Spiers, C. J. A microphysical model for fault gouge friction applied to subduction megathrusts. J. Geophys. Res. 119, 1510–1529 (2014).

    Article  Google Scholar 

  34. Christensen, N. I. Poisson’s ratio and crustal seismology. J. Geophys. Res. 101, 3139–3156 (1996).

    Article  Google Scholar 

  35. Wang, X. Q. et al. High Vp/Vs ratio: saturated cracks or anisotropy effects? Geophys. Res. Lett. 39, L11307 (2012).

    Article  Google Scholar 

  36. Pimienta, L. et al. Anomalous Vp/Vs ratios at seismic frequencies might evidence highly damaged rocks in subduction zones. Geophys. Res. Lett. 45, 12210–12217 (2018).

    Article  Google Scholar 

  37. Takei, Y. Effect of pore geometry on Vp/Vs: from equilibrium geometry to crack. J. Geophys. Res. 107, 2043 (2002).

    Article  Google Scholar 

  38. Beeler, N. M., Hirth, G., Thomas, A. & Bürgmann, R. Effective stress, friction, and deep crustal faulting. J. Geophys. Res. 121, 1040–1059 (2016).

    Article  Google Scholar 

  39. den Hartog, S. A., Peach, C. J., de Winter, D. M., Spiers, C. J. & Shimamoto, T. Frictional properties of megathrust fault gouges at low sliding velocities: new data on effects of normal stress and temperature. J. Struct. Geol. 38, 156–171 (2012).

    Article  Google Scholar 

  40. den Hartog, S. A. & Spiers, C. J. Influence of subduction zone conditions and gouge composition on frictional slip stability of megathrust faults. Tectonophysics 600, 75–90 (2013).

    Article  Google Scholar 

  41. Rubinstein, J. L. et al. Non-volcanic tremor driven by large transient shear stresses. Nature 448, 579–582 (2007).

    Article  Google Scholar 

  42. Rubinstein, J. L., La Rocca, M., Vidale, J. E., Creager, K. C. & Wech, A. G. Tidal modulation of nonvolcanic tremor. Science 319, 186–189 (2008).

    Article  Google Scholar 

  43. Audet, P. & Bürgmann, R. Possible control of subduction zone slow-earthquake periodicity by silica enrichment. Nature 510, 389–392 (2014).

    Article  Google Scholar 

  44. Hyndman, R. D. & Peacock, S. M. Serpentinization of the forearc mantle. Earth Planet. Sci. Lett. 212, 417–432 (2003).

    Article  Google Scholar 

  45. Liu, Y. & Rice, J. R. Spontaneous and triggered aseismic deformation transients in a subduction fault model. J. Geophys. Res. 112, B09404 (2007).

    Google Scholar 

  46. Segall, P. & Bradley, A. M. Slow-slip evolves into megathrust earthquakes in 2D numerical simulations. Geophys. Res. Lett. 39, L18308 (2012).

    Article  Google Scholar 

  47. Oppenheimer, D. et al. The Cape Mendocino, California, earthquakes of April 1992: subduction at the triple junction. Science 261, 433–438 (1993).

    Article  Google Scholar 

  48. Murray, M. H., Marshall, G. A., Lisowski, M. & Stein, R. S. The 1992 M = 7 Cape Mendocino, California, earthquake: coseismic deformation at the south end of the Cascadia megathrust. J. Geophys. Res. 101, 17707–17725 (1996).

    Article  Google Scholar 

  49. Li, D., McGuire, J. J., Liu, Y. & Hardebeck, J. L. Stress rotation across the Cascadia megathrust requires a weak subduction plate boundary at seismogenic depths. Earth Planet. Sci. Lett. 485, 55–64 (2018).

    Article  Google Scholar 

  50. Chen, X. & McGuire, J. J. Measuring earthquake source parameters in the Mendocino triple junction region using a dense OBS array: implications for fault strength variations. Earth Planet. Sci. Lett. 453, 276–287 (2016).

    Article  Google Scholar 

  51. Wei, M. & McGuire, J. J. The Mw 6.5 offshore Northern California earthquake of 10 January 2010: ordinary stress drop on a high‐strength fault. Geophys. Res. Lett. 41, 6367–6373 (2014).

    Article  Google Scholar 

  52. Savard, G., Bostock, M. G. & Christensen, N. I. Seismicity, metamorphism, and fluid evolution across the northern Cascadia fore arc. Geochem. Geophys. Geosyst. 19, 1881–1897 (2018).

    Article  Google Scholar 

  53. Calvert, A. J., Bostock, M. G., Savard, G. & Unsworth, M. J. Cascadia low frequency earthquakes at the base of an overpressured subduction shear zone. Nat. Commun. 11, 3874 (2020).

    Article  Google Scholar 

  54. Nedimović, M. R., Hyndman, R. D., Ramachandran, K. & Spence, G. D. Reflection signature of seismic and aseismic slip on the northern Cascadia subduction interface. Nature 424, 416–420 (2003).

    Article  Google Scholar 

  55. Beeler, N. M., Hirth, G., Tullis, T. E. & Webb, C. H. On the depth extent of coseismic rupture. Bull. Seismol. Soc. Am. 108, 761–780 (2018).

    Article  Google Scholar 

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

    Article  Google Scholar 

  57. Liu, K., Levander, A., Zhai, Y., Porritt, R. W. & Allen, R. M. Asthenospheric flow and lithospheric evolution near the Mendocino Triple Junction. Earth Planet. Sci. Lett. 323, 60–71 (2012).

    Article  Google Scholar 

  58. Northern California Earthquake Data Center UC Berkeley Seismological Laboratory. Dataset (NCEDC, 2014); https://doi.org/10.7932/NCEDC

  59. Levander, A. Seismic and Geodetic Investigations of Mendocino Triple Junction Dynamics (International Federation of Digital Seismograph Networks, 2007); https://doi.org/10.7914/SN/XQ_2007

  60. Guo, H., Zhang, H. & Froment, B. Structural control on earthquake behaviors revealed by high-resolution Vp/Vs imaging along the Gofar transform fault, East Pacific Rise. Earth Planet. Sci. Lett. 499, 243–255 (2018).

    Article  Google Scholar 

  61. Waldhauser, F. & Schaff, D. P. Large-scale relocation of two decades of Northern California seismicity using cross correlation and double-difference methods. J. Geophys. Res. 113, B08311 (2008).

    Google Scholar 

  62. Waldhauser, F. Near-real-time double-difference event location using long-term seismic archives, with application to Northern California. Bull. Seismol. Soc. Am. 99, 2736–2848 (2009).

    Article  Google Scholar 

  63. Du, W. X., Thurber, C. H. & Eberhart-Phillips, D. Earthquake relocation using cross-correlation time delay estimates verified with the bispectrum method. Bull. Seismol. Soc. Am. 94, 856–866 (2004).

    Article  Google Scholar 

  64. Toomey, D. R. et al. The Cascadia initiative: a sea change in seismological studies of subduction zones. Oceanography 27, 138–150 (2014).

    Article  Google Scholar 

  65. Zhang, H. & Thurber, C. H. Double-difference tomography: ahe method and its application to the Hayward fault, California. Bull. Seismol. Soc. Am. 93, 1875–1889 (2003).

    Article  Google Scholar 

  66. Waldhauser, F. & Ellsworth, W. L. A double-difference earthquake location algorithm: method and application to the northern Hayward fault, California. Bull. Seismol. Soc. Am. 90, 1353–1368 (2000).

    Article  Google Scholar 

  67. Zhang, H., Nadeau, R. M. & Guo, H. Imaging the nonvolcanic tremor zone beneath the San Andreas fault at Cholame, California using station-pair double-difference tomography. Earth Planet. Sci. Lett. 460, 76–85 (2017).

    Article  Google Scholar 

  68. Zhang, H., Nadeau, R. M. & Toksoz, M. N. Locating nonvolcanic tremors beneath the San Andreas fault using a station‐pair double‐difference location method. Geophys. Res. Lett. 37, L13304 (2010).

    Article  Google Scholar 

  69. Um, J. & Thurber, C. A fast algorithm for two-point seismic ray tracing. Bull. Seismol. Soc. Am. 77, 972–986 (1987).

    Article  Google Scholar 

  70. Paige, C. C. & Saunders, M. A. LSQR: an algorithm for sparse linear equations and sparse least squares. ACM Trans. Math. Softw. 8, 43–71 (1982).

    Article  Google Scholar 

  71. Thurber, C. H. Hypocenter-velocity structure coupling in local earthquake tomography. Phys. Earth Planet. Inter. 75, 55–62 (1992).

    Article  Google Scholar 

  72. Hole, J. A., Beaudoin, B. C. & Klemperer, S. L. Vertical extent of the newborn San Andreas fault at the Mendocino triple junction. Geology 28, 1111–1114 (2000).

    Article  Google Scholar 

  73. Zelt, C. A. Lateral velocity resolution from three-dimensional seismic refraction data. Geophys. J. Int. 135, 1101–1112 (1998).

    Article  Google Scholar 

  74. Plourde, A. P., Bostock, M. G., Audet, P. & Thomas, A. M. Low‐frequency earthquakes at the southern Cascadia margin. Geophys. Res. Lett. 42, 4849–4855 (2015).

    Article  Google Scholar 

Download references

Acknowledgements

We thank J. Zhu and R. Evans for their efforts in helping us compare our results with unpublished MT models and constructive feedback on the manuscript. We thank P. Segall and N. Beeler for helpful discussions. We thank F. Waldhauser for supplying his NCEDC earthquake catalogue phase data and waveform cross-correlation data. We thank M. Bostock and A. Plourde for supplying their LFE phase data. We thank D. Shelly, J. Hardebeck and S. Hickman for their very helpful comments and suggestions. We thank J. Gao for his help on preparing Extended Data Fig. 8c. H.G. was supported by the China Scholarship Council award 201706340120, NSF EAR award 1520690 and the National Natural Science Foundation of China under grants U1839205 and 41861134009. J.J.M. was supported by NSF EAR award 1520690. H.Z. was supported by the Strategic Priority Research Program of the Chinese Academy of Sciences under grant XDB 41000000, the National Natural Science Foundation of China under grants U1839205 and 41861134009 and the Fundamental Research Funds for the Central Universities under grant WK2080000144.

Author information

Authors and Affiliations

Authors

Contributions

J.J.M. and H.Z. supervised the study. H.G. and J.J.M. conducted the seismic data analysis and designed the seismic tomographic inversion. H.G. and H.Z. developed the tomoTD method and software. H.G. conducted the seismic tomographic inversion and seismic model resolution tests. All the authors contributed to the interpretation of results and the writing of the manuscript.

Corresponding authors

Correspondence to Hao Guo, Jeffrey J. McGuire or Haijiang Zhang.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Geoscience thanks Michael Bostock and Gary Egbert for their contribution to the peer review of this work. Primary Handling Editor: Stefan Lachowycz.

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

Extended data

Extended Data Fig. 1 Model parameterization, initial earthquake locations, initial velocity models, and the trade-off curve analysis for selecting regularization parameters.

(ac) Initial earthquake locations (black dots) and tomographic grid nodes (white dots) in three directions. Green lines represent the axes of X = 0 and Y = 0 with their arrows pointing to the positive directions. Red triangles represent stations. (dg) Vp model from the active-source experiment72 and initial Vp, Vs, and Vp/Vs models at the cross-section of Y = 10 km. (hj) Trade-off curve analysis. (h) Selecting optimal damping value. Lines represent the inversions using smoothing values of 10 (bottom), 100 (middle), and 400 (top). To simplify the analysis for the selection of the damping, the smoothing values of Vp and Vs models are set to be same. Dots on each line represent the inversions using different damping values as labelled. Because the damping is applied to both the event location and velocity model parameters, the norm of all solutions is used for the analysis. The damping value we selected is 500. (i) Selecting optimal smoothing value of the Vp model. Lines represent the inversions using Vs model smoothing values of 10 (bottom), 60 (middle), and 150 (top). Dots on each line represent the inversions using different smoothing values of the Vp model as labelled. All the inversions use the selected optimal damping value. The norm of P-wave slowness parameters and the norm of P-wave data residual are used for the analysis. The optimal smoothing value of the Vp model is 60. (j) Selecting optimal smoothing value of the Vs model. Dots on the line represent the inversions using different smoothing values of the Vs model as labelled. All the inversions use the selected optimal damping and Vp model smoothing values. The norm of S-wave slowness parameters and the norm of S-wave data residual are used for the analysis. The optimal smoothing value of the Vs model is 60. For our final tomographic inversion, the smoothing values of the Vp and Vs models are slightly higher (80) and the damping values are varied around 500 for different iterations to maintain the reasonable condition number of the inversion system.

Extended Data Fig. 2 Earthquake and LFE relocations, location uncertainties, and station corrections.

(a) Map view and cross-section of earthquake (black dots) and LFE (red dots) relocations. Triangles represent the stations used to locate LFEs. (b) Location uncertainties of earthquakes (colored dots) and LFEs (colored stars), estimated from bootstrapping analysis. The median values of earthquake location uncertainties in longitude, latitude, and depth directions are 0.071, 0.058, and 0.090 km, respectively. The median values of LFE location uncertainties in longitude, latitude, and depth directions are 0.589, 0.582, and 1.161 km, respectively. (c, d) P-wave and S-wave station corrections from the tomographic inversion for stations (colored triangles) near the triple junction. Only stations with station-pair data are inverted for station corrections. Gray lines in (a) and (b) represent depth contours of the Gorda plate interface model of ref. 30 in 5 km increments. SAF, San Andreas Fault. MFZ, Mendocino Fault. The dark gray line represents the coast. The black sawtooth line represents the deformation front.

Extended Data Fig. 3 Final Vp and Vs models at the cross-sections of AA’-EE’.

(a) Vp model. (b) Vs model. Earthquakes (black dots) and LFEs (green dots) are within 5 and 20 km of the cross-sections, respectively. Solid lines near the surface represent the bathymetry. Solid lines at depth are 2 km above and 4 km below the plate interface model of ref. 30 (dashed lines). Inverted black triangles represent the deformation front and coast. Low-resolution regions, estimated from checkerboard resolution test, are masked to gray.

Extended Data Fig. 4 Checkerboard resolution test.

(ad) Input and recovered checkerboard Vp and Vs models and the corresponding Vp and Vs model resolvability at the cross-sections of Y = 0, 15, 25, and 45 km. In model resolvability panels, contours of the resolvability value of 0.7 are shown. Note that the ‘X’ in this figure refers to the east-west tomographic coordinate, not the ‘distance’ in the cross-section figures throughout the paper.

Extended Data Fig. 5 Synthetic test for estimating model resolution for the case of increased Vp/Vs plate boundary zone.

(ad) True (a, c) and recovered (b, d) Vs perturbation model (a, b) and Vp/Vs model (c, d) at the cross-sections of AA’-EE’. (e, f) Map view of the true and recovered Vp/Vs models of the plate boundary zone. The true model contains a 6-km-thick, low Vs, and high Vp/Vs layer below the plate interface.

Extended Data Fig. 6 Synthetic test for estimating model resolution for the case of decreased Vp/Vs plate boundary zone.

(a, b) True (a) and recovered (b) Vp/Vs models at the cross-sections of AA’-EE’. (c, d) Map view of the true (c) and recovered (d) Vp/Vs models of the plate boundary zone. The true model contains a 6-km-thick, high Vs, and low Vp/Vs layer below the plate interface.

Extended Data Fig. 7 Megathrust interseismic coupling models.

(ac) Interseismic coupling models calculated with different mantle viscosity values (a, 4.4 × 1018 Pa s; b, 4.0 × 1019 Pa s; c, 3.6 × 1020 Pa s), from ref. 6. Curves in the panels below map views show downdip variations in megathrust locking ratio along BB’, CC’, and DD’ and are colored by the Vp/Vs value of plate boundary zone. All the other elements in map views are the same as Fig. 3a, except that the LFEs are shown as orange dots.

Extended Data Fig. 8 Downdip variations in interseismic coupling, Vp/Vs, and resistivity models.

(a) Map view of the megathrust interseismic coupling model. (bc) Vp/Vs and resistivity model cross-sections along CC’ shown in (a). The resistivity model is taken from Fig. 5 of ref. 25. Blue stations in (a) and (c) represent the MT stations25 within our study area. All the other elements in this figure are the same as those in Fig. 3a,c.

Extended Data Fig. 9 Comparing real data inversion results using tomoDD and tomoTD methods.

Earthquake relocations (dots), Vp, Vs, and Vp/Vs models using (b) tomoDD and (c) tomoTD with catalog data and (d) the differences between tomoTD and tomoDD models at the cross-section of Y = 15 km, starting from the same initial locations and velocity models (a). (eg) The inversion results using different combinations of catalog differential data. (e) station-pair data only. (f) combining station-pair and event-pair differential data. (g) combining station-pair, event-pair, and double-pair differential data. Earthquakes within 10 km of the cross-section are shown. The line near the surface represents the bathymetry or topography. The gray regions near the surface in Vp/Vs models represent the seawater or air and are adjusted to roughly fit the topographic variation. Triangles represent the deformation front and the coast. The other two lines are 2 km above and 4 km below the plate interface of ref. 30, outlining a 6-km-thick layer at the top of the subducted Gorda plate.

Extended Data Fig. 10 Comparing synthetic data inversion results using tomoDD and tomoTD methods.

Vp, Vs, dVp (difference between the inverted and true Vp models), dVs (difference between the inverted and true Vs models), and Vp/Vs models at the vertical cross-section of Y = 15 km from (b, d) tomoDD and (c, e) tomoTD inversions with (b, c) noise-free and (d, e) noisy synthetic data. (a) Input true models. Initial models for all inversions are the same as the one used in the real data inversion (Extended Data Fig. 9a). The gray regions near the surface in dVp, dVs, and Vp/Vs models represent the seawater or air and are adjusted to roughly fit the topographic variation.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Guo, H., McGuire, J.J. & Zhang, H. Correlation of porosity variations and rheological transitions on the southern Cascadia megathrust. Nat. Geosci. 14, 341–348 (2021). https://doi.org/10.1038/s41561-021-00740-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41561-021-00740-1

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