JGR Solid Earth - 2019 - Lu - TX2019slab A New P and S Tomography Model Incorporating Subducting Slabs
JGR Solid Earth - 2019 - Lu - TX2019slab A New P and S Tomography Model Incorporating Subducting Slabs
10.1029/2019JB017448
                                            Incorporating Subducting Slabs
Key Points:
• We developed new P and S
                                            Chang Lu1      , Stephen P. Grand1, Hongyu Lai2            , and Edward J. Garnero2
  tomography models incorporating           1
  3‐D subducting slabs
                                            Department of Geological Sciences, Jackson School of Geosciences, University of Texas at Austin, Austin, TX, USA,
                                            2
• Mislocation effects caused by             School of Earth and Space Exploration, Arizona State University, Tempe, AZ, USA
  subducting slabs were reduced by
  inverting for velocity and source
  location simultaneously                   Abstract Large numbers of earthquakes occur in subduction zones that are marked by dipping, narrow
• The spin transition and the
                                            high seismic velocity slabs. The existence of these fast velocity slabs can cause serious earthquake
  post‐perovskite phase transition
  may explain lower mantle                  mislocation problems that can bias estimates of seismic travel time residuals. This can affect the recovery of
  heterogeneities                           subducting slabs in tomography as well as introduce significant artifacts into lower mantle structure in
                                            tomography models. In order to better account for known subducting slabs, we performed a new P and S
Supporting Information:
                                            wave joint tomography inversion incorporating a three‐dimensional thermal model of subducting slabs in
• Supporting Information S1
                                            the starting model. In addition, velocity and source locations were inverted for simultaneously. Our new P
                                            and S models feature higher‐amplitude subducting slabs compared with previous global tomography results.
Correspondence to:                          The S to P heterogeneity ratio based on the new tomography model indicates that thermal elastic effects
C. Lu,                                      alone cannot explain all the heterogeneities in the lower mantle. Much of the observed abnormal S to P
[email protected]
                                            heterogeneity ratio can be explained by anelastic effects, the spin transition, and phase transitions of
                                            bridgmanite to post‐perovskite in the lower mantle.
Citation:
Lu, C., Grand, S. P., Lai, H., & Garnero,   Plain Language Summary                   Seismic tomography uses seismic travel time data to image deep
E. J. (2019). TX2019slab: A new P and S     earth velocity structure. However, it has been shown that the existence of subducting slabs can
tomography model incorporating
                                            significantly bias the imaging result. In order to reduce this effect, we produced a new P and S wave
subducting slabs. Journal of Geophysical
Research: Solid Earth, 124,                 tomography model that included theoretical three‐dimensional subducting slab structures in the starting
11,549–11,567. https://doi.org/10.1029/     model. The new model has higher‐amplitude subducting slabs compared with other models. Based on the
2019JB017448
                                            new model, we conclude that it is difficult to explain the P and S velocity anomalies found in the deep mantle
                                            by temperature variations alone without invoking complex mineral phase transitions, large anelastic effects,
Received 26 JAN 2019
Accepted 27 SEP 2019                        or chemical variations.
Accepted article online 22 OCT 2019
Published online 11 NOV 2019
                                            1. Introduction
                                            Subduction of oceanic lithosphere is believed to play a critical role in large‐scale mantle convection (e.g.,
                                            Billen, 2008; Kellogg et al., 1999). The investigation of deep subducting slabs, therefore, has been an active
                                            field in geophysics. It is widely accepted that subducting slabs are relatively cold and thus have high seismic
                                            velocity compared to surrounding mantle (e.g., Fukao & Obayashi, 2013; Zhao et al., 2017). Global seismic
                                            tomography, a method to image deep earth seismic velocity structure, has been used to constrain the loca-
                                            tion of subducting slabs. Generally, shear wave global tomography has limited ability to image short‐
                                            wavelength structures such as subducting slabs (e.g., French & Romanowicz, 2014; Grand, 2002;
                                            Kustowski et al., 2008; Moulik & Ekström, 2014; Panning & Romanowicz, 2006; Ritsema et al., 2011).
                                            Compressional wave global tomography has provided higher‐resolution images of upper mantle slabs
                                            (e.g., Amaru, 2007; Li et al., 2008; Obayashi et al., 2013; Simmons et al., 2012). However, detailed waveform
                                            modeling studies suggest that the amplitude of subducting slabs is underestimated in global P wave tomogra-
                                            phy models (Zhan et al., 2014). Tao et al. (2018) performed full‐waveform inversion using upper mantle tri-
                                            plicated waves to image the subducting slabs beneath Eastern Asia. Both of their P and S models show much
                                            higher‐amplitude velocity anomalies inside the slabs than in global tomography models. Other seismic mod-
                                            elling results, as well as theoretical thermal models of subducting slabs, also imply far stronger velocity
                                            anomalies within slabs relative to those seen in global tomography (Chen et al., 2007; Kawakatsu &
                                            Yoshioka, 2011; Syracuse et al., 2010; Wang et al., 2014). The discrepancies between detailed seismic studies
©2019. American Geophysical Union.          of slabs with global tomography models imply a potential problem with using tomography models to infer
All Rights Reserved.                        density anomalies for use in mantle convection studies.
LU ET AL.                                                                                                                                                11,549
                                                                                                                                                             21699356, 2019, 11, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JB017448 by Cochrane Saudi Arabia, Wiley Online Library on [16/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                         Journal of Geophysical Research: Solid Earth                                                       10.1029/2019JB017448
                                                                              The existence of sharp, dipping fast velocity anomalies can also bias earth-
                                                                              quake locations and impact seismic travel time data used in tomography
                                                                              (Creager & Jordan, 1984; Ding & Grand, 1994; Sleep, 1973). Therefore,
                                                                              the incorrect imaging of subducting slabs could degrade tomography
                                                                              models even far from subduction zones. Through synthetic testing, Lu
                                                                              and Grand (2016) found that the incorrect imaging of subducting slabs
                                                                              could introduce up to 0.5% amplitude false velocity anomalies in the lower
                                                                              mantle in global shear wave tomography. This is comparable to the ampli-
                                                                              tude of velocity anomalies found in the lower mantle. These artifacts may
                                                                              have significant implications for interpretation of lower mantle heteroge-
                                                                              neities. For example, some studies use the S to P heterogeneity ratio to
                                                                              identify potential chemically distinct heterogeneities (e.g., Koelemeijer
                                                                              et al., 2016; Masters et al., 2000; Saltzer et al., 2001; Tesoniero et al.,
                                                                              2016). Artifacts produced by incorrectly accounting for subducting slabs
                                                                              can impact the reliability of results using this approach.
                                                                          Lu and Grand (2016) compare several strategies to best account for sub-
Figure 1. Distribution of travel time residuals measured at station MAW for
                                                                          ducting slabs in global tomography. They suggest including a priori sub-
a representative 30 × 30‐km earthquake group used in the EHB data selec-
                                                                          ducting slabs in the starting model and performing structure and source
tion process. The group is centered at 7.3°S, 155.9°E, and 45‐km depth. The
station MAW is located in Antarctica 67.6°S, 62.9°E. The residuals outsidelocation jointly in tomographic inversions. In this study, we present a
±2σ were treated as outliers and eliminated from the inversion.           new P and S global tomography model using a three‐dimensional (3‐D)
                                                                          subducting slab structure in the initial model. This enables us to produce
                                        a global tomography model with more realistic subducting slabs. In the lower mantle, we evaluate the effect
                                        of subducting slabs on S to P heterogeneity ratio estimation.
LU ET AL.                                                                                                                                          11,550
                                                                                                                                                                 21699356, 2019, 11, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JB017448 by Cochrane Saudi Arabia, Wiley Online Library on [16/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                           Journal of Geophysical Research: Solid Earth                                                         10.1029/2019JB017448
LU ET AL.                                                                                                                                              11,551
                                                                                                                                               21699356, 2019, 11, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JB017448 by Cochrane Saudi Arabia, Wiley Online Library on [16/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
            Journal of Geophysical Research: Solid Earth                                                   10.1029/2019JB017448
                   Figure 3. Distribution of travel time residuals caused by the input slab model. Plotted are the differences between a 1‐D
                   prediction of travel times and predictions for the same 1‐D model with slabs embedded for all the data used in our
                   inversion. P wave residuals were derived from 3‐D ray tracing (Simmons et al., 2012), while the S wave residuals were
                   obtained from 3‐D ray tracing and SEM simulations (Komatitsch & Tromp, 2002a, 2002b).
                   the spherical tessellation grids, we compared 3‐D ray tracing results with SEM simulation results for a few
                   test earthquakes. We find that the when the tessellation recursion level is above 9 (corresponding to ~0.25
                   arc degree average node spacing), the difference between 3‐D ray tracing and SEM simulations agree to
                   within 0.2 s, which we consider acceptable for our purposes. The effect of subducting slabs on the seismic
                   travel time residuals are summarized in Figure 3. The largest travel time residual caused by our starting
                   model is ~8 s for P wave data while it is ~12 s for S wave data.
                   The “TX” data set contains upper mantle turning waves. As discussed in Grand (1994), due to the large het-
                   erogeneity in the shallow mantle, at a given distance the ray paths of S or SS waves can be quite different
                   depending on the specific region being sampled. Relative to standard 1‐D models the difference is especially
                   significant in cratonic regions. For this reason, different seismic models were used to determine ray paths for
                   upper mantle waves in Grand (1994, 2002). This process required detailed waveform analysis, and thus is not
                   suitable for large‐volume catalogue data. The EHB P data set contains upper mantle turning waves. In order
                   to correct the ray paths that sample cratonic regions for EHB data, we first derived a 3‐D S wave craton
                   model and then converted it to a P wave model (Figure S2). The fast velocity anomalies in our previous S
                   wave tomography at the shallowest depths (Lu & Grand, 2016) are used to determine the location of cratons
                   if they have Archean or Proterozoic crust (Laske et al., 2013). The fast velocity anomalies down to 210 km
                   which are close to these regions are used to define the craton roots. These cold cratons are believed to have
                   distinct chemical compositions relative to surrounding mantle and contain high percentages of Mg. Lee
                   (2003) showed that the effect of Mg# (100 × Mg/(Mg + Fe)) on Vp/Vs ratio was significantly larger than
                   the effect of temperature in peridotite. Therefore, we estimated the Mg# in cratons using a linear interpola-
                   tion based on our S wave craton model, assuming that normal mantle has a low (87) Mg# while the fastest
                   craton has the highest (94) Mg# (Lee, 2003). Using this approach, the average variation of Mg# in cratons is
                   ~3.5, which agrees with the value reported by Deschamps et al. (2002) and is slightly higher than predictions
                   by Forte and Perry (2000) and Perry et al. (2003). Then we adopted the linear relationship between Mg# and
                   Vp/Vs ratio reported by Lee (2003) to derive a P wave craton model. Using the P wave craton model, 3‐D ray
                   tracing (Simmons et al., 2012) was used to determine the ray paths for rays turning above 800‐km depth
                   (Figure S3). For rays turning below 800‐km depth, the effect of cratons on seismic ray paths is negligible.
                   2.5. Joint P and S Inversion Starting From 3‐D Slab
                   In our inversion, the mantle was divided into 99,146 blocks. The blocks are about 275 × 275 km in lateral
                   dimension and vary from 75 to 240 km in thickness. Both P and S wave sensitivity kernels were calculated
                   using the ray theory approximation based on a 1‐D velocity model. For S waves, we use a combined
                   TNA/SNA model (Grand & Helmberger, 1984), which is the same 1‐D model as we used in previous TX mod-
                   els (Grand, 2002). The AK‐135 model (Kennett et al., 1995) is used as the P wave starting model. The P and S
                   wave data are known to have very different ray coverage, especially in the shallow mantle. At shallow
                   depths, the P wave data have limited sampling of the oceans, while S wave data have much better resolution
                   there. Therefore, we correlated the P wave and S wave models by introducing another term XP/S into our
LU ET AL.                                                                                                                           11,552
                                                                                                                                    21699356, 2019, 11, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JB017448 by Cochrane Saudi Arabia, Wiley Online Library on [16/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
            Journal of Geophysical Research: Solid Earth                                             10.1029/2019JB017448
                   inversion, which is the ratio of the P and S wave slowness perturbations in each layer. This parameter can be
                   obtained using thermoelastic parameters from mineral physics measurements as a function of depth assum-
                   ing a thermal origin for mantle seismic heterogeneity (see Lu & Grand, 2016 for details). The XP/S in cratons
                   vary in 3‐D and are consistent with the values used in our forward modeling. Also, Lu and Grand (2016)
                   showed that correcting for earthquake mislocation can be better achieved if one inverts for velocity structure
                   and earthquake location simultaneously. Therefore, we include earthquake location in our inversion as well.
                   The linearized seismic tomography problem can be written as
                                                 2                           3         2         3
                                                     GP                 AP                  rP
                                                                              2    3
                                                 6                       λAS 7         6    λr S 7
                                                 6           λGS             7 mP      6         7
                                                 6                           76    7 6           7
                                                 6 DP                        76 mS 7 6       0 7
                                                 6                           76    7 6           7
                                                 6                           76 LP 7 ¼ 6         7                            (1)
                                                 6                           74    5 6           7
                                                 6           DS              7         6     0 7
                                                 4                           5 LS      4         5
                                                   λX     −λX X P=S                          0
                   where GP and GS are the P and S wave sensitivity kernel matrices and rP and rS are the corresponding travel
                   time residuals caused by velocity anomalies and earthquake mislocation. The travel time residuals r are the
                   leftover residuals after we removed the effect of subducting slabs, which is the most significant difference
                   from previous global tomography studies. We use λ to represent the relative weight of S wave data to P wave
                   data. mP and mS are the P wave and S wave slowness perturbations and L represents the relocation para-
                   meters to be inverted that include changes in event latitude, longitude, and depth, as well as origin time.
                   A is the relocation matrix formed of the partial derivatives for these parameters. D is a smoothing operator,
                   which is a Laplacian filter with 76% of the weight applied to horizontal nearby blocks and 24% of the weight
                   applied to the vertical nearby blocks. A weighting term λX is used to control how strong the connection
                   between the P and S models is enforced. Through a trial and error process, λX was chosen to be 500 which
                   was the maximum value before the connection term began to decrease the variance reduction of the S wave
                   data (less than 0.5%; Figure S4).
                   3. Results
                   3.1. TX2019slab Model
                   We show our new model (TX2019slab) in horizontal slices at selected depths in Figure 4. The blocks used in
                   the inversion and the input slab model have been resampled onto a 1° × 1° grid when making the plot. Our P
                   and S models generally agree with previous tomography results at large scale. At 150‐km depth, our model
                   shows fast velocity anomalies in cratonic regions as expected. Slow velocity anomalies are found beneath
                   mid‐ocean ridges as well as the East African Rift. Due to poor data coverage in the oceans, most previous
                   global P wave tomography models do not have slow mid‐ocean ridges (e.g., Amaru, 2007; Li et al., 2008;
                   Obayashi et al., 2013). Our shallow oceanic P structure is mainly constrained by the S wave data and the rela-
                   tionship between P and S anomalies we introduced into the inversion. In the upper mantle and transition
                   zone, short‐wavelength subducting slabs appear to be the most heterogeneous structures in our model.
                   Interestingly, the slab signature near 600‐km depth is much broader than at 300‐km depth. In the mid‐lower
                   mantle, as in previous tomography studies, two elongated fast velocity anomalies in P and S are seen beneath
                   North/South America and South Asia. The locations of these fast anomalies generally agree with previous
                   studies (e.g., Amaru, 2007; French & Romanowicz, 2014; Grand, 2002; Li et al., 2008; Moulik & Ekström,
                   2014; Obayashi et al., 2013; Ritsema et al., 2011). Beginning in the mid‐lower mantle, slow anomalies are
                   seen in both P and S beneath the south central Pacific and Africa. They increase in strength and size with
                   depth and at the bottom of the mantle are quite broad. These two structures have been called Large Low
                   Shear Velocity Provinces (LLSVPs; see Garnero et al., 2016 for a review). Note that the P wave amplitude
                   is muted at the base of the mantle inside the Pacific LLSVP.
                   Our model yields a 33.8% overall variance reduction for the P wave data and 91.4% for S wave data. In com-
                   parison, we derived another P and S model using the same data, inversion method, and regularization but
                   starting from a 1‐D velocity model. We refer to this model as TX2019. The variance reduction, using model
                   TX2019, is slightly less than 33.8% for the P wave data and 91.3% for the S wave data. The inversion starting
LU ET AL.                                                                                                                 11,553
                                                                                                                                                21699356, 2019, 11, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JB017448 by Cochrane Saudi Arabia, Wiley Online Library on [16/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
            Journal of Geophysical Research: Solid Earth                                                    10.1029/2019JB017448
                   Figure 4. Lateral velocity variations in model TX2019slab at representative depths. Each row corresponds to the model at a
                   particular depth, and the two columns show (left) P wave and (right) S wave models, respectively. The color scales change
                   for each plot according to the amplitude (X) labeled at the bottom right corner. Purple lines in the second row show the
                   locations of cross sections in Figures 7 and 8.
                   from a model with 3‐D slabs results in a slightly better data fit than the inversion starting from the 1‐D model
                   although the difference is minimal. This is a clear indication of the nonuniqueness of models in global
                   seismic tomography. The addition of slabs in our starting model, however, results in a model more
                   consistent with regional studies and geodynamic models without sacrificing fit to global data.
LU ET AL.                                                                                                                            11,554
                                                                                                                                                 21699356, 2019, 11, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JB017448 by Cochrane Saudi Arabia, Wiley Online Library on [16/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
            Journal of Geophysical Research: Solid Earth                                                            10.1029/2019JB017448
                   coefficient (CC) between the input (min) and output models (mout) at each depth has been calculated and
                   labeled in Figure 5. The CC of two scalar fields can be calculated as
                                                                             N               
                                                                         ∑i¼1 X 1;i X 2;i
                                                          CC ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
                                                                                   2ffiqffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
                                                                                                          2ffi                              (2)
                                                                    N                     N 
                                                                ∑i¼1 X 1;i              ∑i¼1 X 2;i
                   where X1,i and X2,i are the ith elements in X1 and X2 after having mean values removed, respectively. We also
                   calculated amplitude recovery (AR) to illustrate the resolution of the inversion. AR is defined as the ratio of
                   the root‐mean‐square (RMS) amplitudes of min and mout as
                                                          rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
                                                                                                           
                                                                  N                            N
                                                     AR ¼     ∑i¼1 mout 2 = ∑i¼1 min 2 ×100%                                               (3)
                   The CCs vary from 0.68 to 0.98, while amplitude recoveries are generally below 83% because of the regular-
                   ization. Because of the S to PP wave velocity scaling used in the inversion, the shallow oceanic regions in the
                   P wave model show some structures simply scaled from the S wave model. However, these regions are still
                   the least resolved regions in the mantle. The derived P wave model in these regions highly depends on the S
                   wave model as well as the scaling relationship used in the inversion.
                   4. Discussion
                   Compared with previous global tomography studies, the most significant difference in our model is that we
                   performed our inversion starting with a 3‐D subducting slab model and invert for velocity and earthquake
                   location simultaneously. We compare our new model to previously published models in vertical cross sec-
                   tions across four major convergent plate boundaries. Figure 7 compares P models across the sections
                   (Amaru, 2007; Li et al., 2008; Obayashi et al., 2013), and Figure 8 compares S models (French &
                   Romanowicz, 2014; Moulik & Ekström, 2014; Ritsema et al., 2011). The cross‐section locations are shown in
                   Figure 4.
                   For the P wave models, similar fast anomalies are seen in all four tomography models (Figure 7). Across the
                   northern Honshu arc, our model shows stagnant slab above the 660 discontinuity, and agrees with all the
                   other models. The cross section across the western Java arc, beneath which the Indo‐Australian plate is sub-
                   ducting, shows large fast anomalies in the uppermost lower mantle, with an extension of fast velocity into
                   the deeper mantle to the north. Similar features can be seen in the other models, although our new model
                   has a larger, stronger anomaly in the deepest mantle. The Tonga and South American cross sections show
                   more differences among models. This is likely due to worse azimuthal station coverage around these regions.
                   The cross sections across the Tonga arc show very complex structures. In our model, a stagnant slab above
                   the 660 discontinuity is seen but there is also an anomaly in the lower mantle with a gap between the two.
                   This is most similar to the GAPP4 model except in that model there is no gap between the deeper structure
                   and the anomalies in the transition zone. Some studies (e.g., Bonnardot et al., 2009; Brudzinski & Chen,
                   2005) suggest that the stagnant slab above the 660 is the southwestward flattening of the downgoing slab
LU ET AL.                                                                                                                            11,555
                                                                                                                                           21699356, 2019, 11, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JB017448 by Cochrane Saudi Arabia, Wiley Online Library on [16/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
            Journal of Geophysical Research: Solid Earth                                                 10.1029/2019JB017448
                   Figure 5. Output from a checkerboard test for (left column) P wave and (right column) S wave data at selected depths.
                   Correlation coefficients (CCs) and amplitude recovery (AR) between the input checkerboard and the inversion results
                   are labeled on each plot.
                   beneath the Lau basin, while the slab below the 660 is the westward extension of the downgoing slab from
                   the southern Tonga trench (see Fukao & Obayashi, 2013, Figure 11). Richards et al. (2011), however, treat
                   the stagnant slab above the 660 as a slab remnant detached from the Vanuatu trench. In South America,
                   both our model and UU07P show a dipping slab in the upper mantle but the slab in our model penetrates
                   deeper into the lower mantle. Possibly due to poor ray coverage, models GAPP4 and MIT08P have less clear
                   subducting slabs in the upper mantle and little continuation of slab into the lower mantle.
                   Compared with the other models, the most significant difference with our model is the higher amplitude of
                   velocity anomalies within subducting slabs. In our model, the amplitude of P wave velocity anomalies inside
                   the slab in the upper mantle and transition zone are mostly higher than 2% while in the other models the
                   anomalies are less than 1.5%. The difference is more significant in the regions where ray coverage is limited.
                   The difference, of course, is because we include slabs in our starting model (Figure S6). The differences this
LU ET AL.                                                                                                                       11,556
                                                                                                                                               21699356, 2019, 11, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JB017448 by Cochrane Saudi Arabia, Wiley Online Library on [16/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
            Journal of Geophysical Research: Solid Earth                                                   10.1029/2019JB017448
                   Figure 6. Representative epicenter shifts derived from velocity and source location joint inversion. The circles show the
                   original earthquake epicenters. The red lines show the direction and amplitude of epicenter shifts. Only earthquakes
                   with more than 250 travel time residuals and a focal depth greater than 100 km are shown for better visualization.
                   causes elsewhere in the model are smaller in amplitude but can still be significant. Given that the slabs in our
                   model are consistent with regional studies and theoretical modeling, and that the data variance reduction is
                   the same or slightly improved over a model derived without the slabs in the starting model, we feel that our
                   new model has some advantages over previous models.
                   The comparison of our model with other S wave models also shows similar features (Figure 8). Generally, S
                   wave models are longer wavelength and narrow slabs, such as in our model, are not well resolved. The slabs
                   we start with are consistent with geologic inferences and also result in a slightly higher variance reduction of
                   Figure 7. Comparison of cross sections across four major subduction zones in representative P wave tomography models.
                   Each column shows specific tomography models, including TX2019_slab (this study), GAPP4 (Obayashi et al., 2013),
                   MIT08P (Li et al., 2008), and UU07P (Amaru, 2007). Each row corresponds to cross‐section locations indicated in Figure 4.
                   Solid black lines show 410‐, 660‐, and 1,000‐km depth, respectively.
LU ET AL.                                                                                                                           11,557
                                                                                                                                          21699356, 2019, 11, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JB017448 by Cochrane Saudi Arabia, Wiley Online Library on [16/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
            Journal of Geophysical Research: Solid Earth                                                10.1029/2019JB017448
                  Figure 8. Comparison of cross sections across major subduction zones in representative S wave tomography models. Each
                  column shows specific tomography models, including TX2019_slab (this study), S40rts (Ritsema et al., 2011), UCBSEM
                  (French & Romanowicz, 2014), and S362+M (Moulik & Ekström, 2014). Each row corresponds to cross‐section locations
                  indicated in Figure 4. Solid black lines show 410‐, 660‐, and 1,000‐km depth, respectively.
                   the S wave data. Our S wave model shows similar features as seen in our P wave model, including dipping,
                   stagnant, and penetrated slabs (Figure S6). The other three S wave models, however, only show very long‐
                   wavelength fast velocity anomalies, which makes it challenging to identify the location and shape of sub-
                   ducting slabs. Also, the amplitude of the subducting slabs in our S wave model are mostly higher than 3%,
                   which are again much higher than in the other models.
                   The S to P heterogeneity ratio, defined as
                                                                                  ∂lnV S
                                                                    RV S =V P ¼                                                    (4)
                                                                                  ∂lnV P
                   has been widely used as an important diagnostic parameter to determine the compositional state of the
                   Earth's mantle, especially in the deeper mantle (e.g., Della Mora et al., 2011; Houser et al., 2008; Karato &
                   Karki, 2001; Koelemeijer et al., 2016; Masters et al., 2000; Robertson & Woodhouse, 1996; Saltzer et al.,
                   2001; Tesoniero et al., 2016). Several approaches have been used to calculate the S to P mantle heterogeneity
                   ratio (see Tesoniero et al., 2016 for a summary). It has been shown that different approaches can lead to dif-
                   ferent heterogeneity ratios, even using the same tomography model (Koelemeijer et al., 2016). We used two
                   methods to determine the 1‐D S to P heterogeneity as a function a depth using our TX2019slab model
                   (Figures 9 and S7). In the first method, a point to point division is computed except for regions where Vp
                   or Vs variations are less than 0.1%. Regions that have opposite sign Vp and Vs anomalies, which contribute
                   to about 5% of the volume of the mantle, were also excluded. The median values at each depth are chosen
                   as the 1‐D heterogeneity ratio. In the second method, we divided the RMS of the velocity variations for P
                   and S at each depth after excluding small or opposite variations as in the first method (Figure 9).
                   Throughout most of the mantle, the S to P ratio derived by RMS division is higher than by point to point divi-
                   sion. A similar observation is seen in the comparison done by Koelemeijer et al. (2016) using tomography
                   model KRDH16. This is due to the fact that the RMS value is more sensitive to outliers than the median
                   value. The LLSVPs have large shear anomalies with smaller P anomalies and are the primary cause of the
                   difference between the two methods in the deepest mantle. Even though the two methods described above
                   lead to different 1‐D S to P ratios, both S to P ratio models show similar trends.
LU ET AL.                                                                                                                      11,558
                                                                                                                                                                   21699356, 2019, 11, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JB017448 by Cochrane Saudi Arabia, Wiley Online Library on [16/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                         Journal of Geophysical Research: Solid Earth                                                           10.1029/2019JB017448
Figure 9. (a) S to P heterogeneity ratio derived using the TX2019slab model as a function of depth. Two S to P ratio profiles are derived by point to point (P2P)
division and RMS value division, respectively. (b) Distribution of P2P division results at selected depths.
                                        Lu and Grand (2016) showed that earthquake mislocation caused by subducting slabs could bias the tomo-
                                        graphy results in the lower mantle. This bias has the potential to further affect our estimation of the S to P
                                        ratio. To evaluate this effect, we also derived the S to P heterogeneity ratio for model TX2019 model using the
                                        two approaches described above (Figure 10). We find very little difference in the S to P heterogeneity ratio
                                        between the slab model and the model without starting slabs. We also calculated the point to point S to P
                                        ratio differences between the TX2019slab and TX2019 models and plot them in map view in Figure 11.
                                        The S to P ratio can differ by more than 0.5 between the models in specific regions. This is a significant dif-
                                        ference since the lateral average S to P ratio in the lower mantle is between 1.5 and 3.5 in our model. Our
                                        analysis shows that the 1‐D average of P to S heterogeneity ratio is unlikely to be significantly affected by
                                        the bias introduced by unmodeled subducting slabs. However, the bias may be quite large for the S to P het-
                                        erogeneity ratio on a regional scale.
                                        In the deeper mantle, our 1‐D S to P heterogeneity ratio profiles show an increase with depth, which is con-
                                        sistent with most previous tomography studies (Antolik et al., 2003; Della Mora et al., 2011; Ishii & Tromp,
                                        1999; Koelemeijer et al., 2016; Masters et al., 2000; Mosca et al., 2012; Resovsky & Trampert, 2003;
                                        Romanowicz, 2001; Su & Dziewonski, 1997; Tesoniero et al., 2016) (Figure 12). We followed Koelemeijer
                                        et al. (2016) and used the RMS division method for all the models in Figure 12. The S to P heterogeneity ratio
                                        profiles derived using body wave seismic data alone are plotted separately (Figure 12a) from results that also
                                        used normal model data (Figure 12b). Among these profiles, the RV S =V P range from about 1.5 to 4 throughout
                                        the lower mantle. The RV S =V P profiles vary more dramatically in the deep lower mantle than at shallower
                                        depths among the different models. In particular, most of the RV S =V P profiles agree with each other above
                                        1,500‐km depth. These models were produced using different data sets and different inversion strategies;
                                        therefore, the similarity indicates a convergence of RV S =V P estimations in this depth range. In the deep lower
                                        mantle, RV S =V P in KRDH16 (Koelemeijer et al., 2016) increases rapidly with depth and reaches ~4 near 2,450
                                        km, then decreases again to the CMB. This feature is not seen in the other models. Most of the other models,
                                        including TX2019slab, show a gradual increasing RV S =V P with depth in this depth range. In contrast, there are
                                        several models have a small decrease above the CMB, including DBTNG11 (Della Mora et al., 2011), R01
                                        (Romanowicz, 2001), TCB16 (Tesoniero et al., 2016), and SFBG10 (Simmons et al., 2010).
                                        Figure 12c compares our RV S =V P in the lower mantle with predictions assuming that seismic heterogeneity
                                        is due to thermal variations alone. All the predicted RV S =V P profiles in Figure 12c assume a simple
LU ET AL.                                                                                                                                               11,559
                                                                                                                                               21699356, 2019, 11, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JB017448 by Cochrane Saudi Arabia, Wiley Online Library on [16/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
            Journal of Geophysical Research: Solid Earth                                                   10.1029/2019JB017448
                   Figure 10. Comparison of S to P heterogeneity ratios using TX2019slab (solid lines) and TX2019 (dash lines). S to P ratio
                   profiles are derived by point to point (P2P) division (green lines) and RMS value division (red lines), respectively.
                   “pyrolite” mineral assemblage model (e.g., Stixrude & Lithgow‐Bertelloni, 2012) without invoking more
                   complex proposed scenarios in the lower mantle such as spin transition effects (Lin et al., 2013; Wu &
                   Wentzcovitch, 2014) or the phase transition from bridgmanite to post‐perovskite (pPv; Murakami et al.,
                   2004; Oganov & Ono, 2004; Tsuchiya et al., 2004). The estimate of RV S =V P by Karato and Karki (2001) is
                   lower than values we obtained using the RMS division method from about 1,500‐km depth and below
                   Figure 11. The difference in point to point S to P heterogeneity ratio between the TX2019slab and TX2019 models in map
                   view at representative depths. The S to P heterogeneity ratio was calculated in each individual block. Although the 1‐D
                   average S to P heterogeneity ratio is not significantly different between these two models, the incorrect imaging of sub-
                   ducting slabs could bias the tomography results in the lower mantle in specific regions, as shown by high‐amplitude
                   regions in this figure.
LU ET AL.                                                                                                                           11,560
                                                                                                                                                                        21699356, 2019, 11, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JB017448 by Cochrane Saudi Arabia, Wiley Online Library on [16/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                          Journal of Geophysical Research: Solid Earth                                                              10.1029/2019JB017448
Figure 12. Comparison of S to P heterogeneity ratios in the lower mantle from (a) pure body wave seismic study, (b) normal mode data involved seismic study, and
(c) mineral physics predictions assuming a thermal cause for heterogeneity in the lower mantle. Results in this study are shown using red solid (RMS division
method) and green solid lines (point to point method). S to P heterogeneity ratio derived from mineral physics modeling from this study and used in the P and S joint
inversion is also shown as a red dot dash line. (a) Results from previous pure body wave studies include AGED03 (Antolik et al., 2003), HMSL08 (Houser et al.,
2008), SD97 (Su & Dziewonski, 1997), SHK01 (Saltzer et al., 2001), DBTNG11 (Della Mora et al., 2011), SFBG10 (Simmons et al., 2010), and TCB16 (Tesoniero et al.,
2016). All the S to P heterogeneity ratios were obtained using the RMS value division method. (b) Results from models that used normal mode data include MLDB00
(Masters et al., 2000), IT99 (Ishii & Tromp, 1999), R01 (Romanowicz, 2001), KRDH16 (Koelemeijer et al., 2016), and MCDRT12 (Mosca et al., 2012). All the S to P
heterogeneity ratios again were obtained using the RMS value division method. (c) Thermally induced S to P heterogeneity ratios derived from mineral physics
modeling including the effect of anelasticity (Trampert et al., 2001; Karato, 1993; Karato & Karki, 2001; Yang et al., 2016) are shown in colored dot dash lines.
                                         but is similar to values we derived using point to point division. However, the Karato and Karki (2001)
                                         estimates are significantly higher than other mineral physics predictions (Karato, 1993; Trampert et al.,
                                         2001; Yang et al., 2016) as well as ours. Two studies have argued that Karato and Karki (2001)
                                         overestimated the effect of anelasticity in the lower mantle, which would overpredict RV S =V P since the
                                         temperature derivative of shear wave velocity is more sensitive to the anelasticity effect (Brodholt et al.,
                                         2007; Matas & Bukowinski, 2007). If this is the case, our derived RV S =V P implies factors other than
                                         simple thermal variations, starting from near 1,500‐km depth, contribute to seismic heterogeneity. We
                                         also plotted the predicted RV S =V P from mineral physics modeling in this study in Figures 12a and 12b.
                                         All of the seismically derived RV S =V P profiles have higher values than mineral physics predictions in the
                                         deep lower mantle, which implies the existence of more complex thermal‐chemical heterogeneities,
                                         although the amount and depth of these heterogeneities still vary among models and methods for
                                         estimating 1‐D RV S =V P .
                                         Anelasticity in the Earth for P and S waves is quantified by quality factors QP and QS, respectively. The qual-
                                         ity factor Q due to viscoelastic relaxation can be written as (Karato, 1993)
                                                                                                                   
                                                                                                         α αH *
                                                                                       Qðω; T Þ ¼ Q0 ω exp                                                       (5)
                                                                                                           RT
                                         where ω is the seismic frequency, α is the exponent describing the frequency dependence of the attenuation,
                                         T is the temperature, R is the gas constant, H* is the activation enthalpy, and Q0 is a normalization constant
                                         that can be constrained using seismically observed attenuation. Assuming that bulk attenuation is
LU ET AL.                                                                                                                                                    11,561
                                                                                                                                                    21699356, 2019, 11, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JB017448 by Cochrane Saudi Arabia, Wiley Online Library on [16/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
            Journal of Geophysical Research: Solid Earth                                                               10.1029/2019JB017448
                   negligible, the relationship between elastic (“el”) and anelastic (“anel”) values of RV S =V P can be given by
                   (Matas & Bukowinski, 2007)
                                                                                                                       
                                                                                            AP     3V P 2
                                                RV S =V P   anel
                                                                   ¼ RV S =V P   el
                                                                                      þ                   −RV S =V P el
                                                                                                                                              (6)
                                                                                        V P;T þ AP 4V S 2
where
                                                                                        αH * cot ðαπ=2Þ
                                                                         AP ¼ −                                                               (7)
                                                                                          2QP RT 2
                                                                                           ∂lnV P el
                                                                             V P;T ¼                                                          (8)
                                                                                             ∂T
                   The measurements of several critical anelasticity‐related parameters, including H*and α, still vary a lot
                   among different studies (see Brodholt et al. (2007) and Matas and Bukowinski (2007) for summaries).
                   Therefore, the predicted RV S =V P from mineral physics after taking anelasticity into account may contain sig-
                   nificant uncertainties. In Figure 13, we evaluated the potential effect of anelasticity on predicted RV S =V P to
                   check whether the uncertainty in anelasticity could explain the apparent discrepancies between seismic
                   observed and mineral physics predicted RV S =V P in the deep lower mantle (Figure 12c). To make the compar-
                   ison more diagnostic, we plotted the variation of RV S =V P as a function of S velocity perturbation at selected
                   depths instead of using a 1‐D profile. We varied the anelasticity‐related parameters within their possible
                   ranges to fully explore the possible anelasticity effect. Figures 13b (and 13c only show estimates at 2,000‐
                   km depth but we found similar results at all other depths. According to our test, when the anelasticity effect
                   is strong, mineral physics predicted RV S =V P becomes larger, but the amount of variation is limited. For exam-
                   ple, the maximum RV S =V P is ~2.5 at 2,000‐km depth, which is comparable to our observed 1‐D RV S =V P in
                   TX2019slab (Figure 12). But this maximum value requires the combination of extreme values for multiple
                   parameters (i.e., H* = 800 kJ/mol, α = 0.4), that Brodholt et al. (2007) claim unlikely. For better visualization,
                   we binned and averaged seismically derived RV S =V P based on a 0.2% S velocity perturbation interval and
                   applied a moving average filter in Figure 13a for comparison. At all selected depths, the seismically derived
                   RV S =V P profiles show “V” shapes: RV S =V P values increase with increasing S wave perturbation, both in the
                   positive and negative directions. The RV S =V P for large S wave perturbation regions (larger than 1% for exam-
                   ple) can easily be over 2.8 and even reach 5. Tesoniero et al. (2016) has shown that the “V” shape is a common
                   feature in all the tomography models they tested and argued that the V shape could be a result of errors in
                   seismic tomography. The effect of small errors on RV S =V P values can be amplified when the true velocity per-
                   turbation is small because of the division between S and P wave anomalies (see Tesoniero et al. (2016) for
                   more discussion). If Tesoniero et al. (2016) are correct, it means that the RV S =V P in the large S wave perturba-
                   tion regions are more reliable than in the small‐velocity perturbation regions. In other words, the actual
                   RV S =V P could be higher than our 1‐D seismic observation in Figure 12, which means that anelasticity effects
                   alone cannot explain the observed RV S =V P assuming that heterogeneity is due to thermal effects in a chemi-
                   cally homogeneous mantle. More complex thermal‐chemical structures are needed to explain our seismic
                   observed RV S =V P in the deep lower mantle.
                   An interesting feature in our S to P heterogeneity profile is that the ratio increases starting from 1,400‐km
                   depth and reaches a local maximum around 1,800‐km depth, stays relatively constant to 2,100‐km depth,
                   then decreases before a large jump near the CMB. Both our derived S to P heterogeneity ratios have this pat-
                   tern, although the S to P ratio derived by RMS division shows this feature more clearly. Model SD97 (Su &
                   Dziewonski, 1997) and KRDH16 (Koelemeijer et al., 2016) show a similar trend above 2,100‐km depth but
                   RV S =V P continues to increase with depth below 2,100 km. Several other models have similar peaks but at dee-
                   per depth, including SFBG10 (Simmons et al., 2010), SHK01 (Saltzer et al., 2001), IT99 (Ishii & Tromp, 1999),
                   (Romanowicz, 2001), and MLDB00 (Masters et al., 2000). Wu and Wentzcovitch (2014) reported that the P
                   wave velocity of pyrolitic lower mantle becomes insensitive to temperature variations due to the spin transi-
                   tion between ~1,400‐ and ~2,100‐km depth. This would increase the S to P heterogeneity there even if varia-
                   tions were solely due to thermal effects. Thus, the abnormal RV S =V P we observe between 1,400 and 2,100 km
LU ET AL.                                                                                                                               11,562
                                                                                                                                                                        21699356, 2019, 11, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JB017448 by Cochrane Saudi Arabia, Wiley Online Library on [16/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                          Journal of Geophysical Research: Solid Earth                                                              10.1029/2019JB017448
Figure 13. Comparison of seismically observed S to P heterogeneity ratios versus S velocity perturbations and mineral physics predictions assuming different ane-
lasticity parameters. (a) Results calculated for TX2019slab model at selected depths. The average scaling factors are calculated for each 0.2% velocity perturbation
                                                                                                                                                                   *
interval. A three‐point moving‐average filter was applied before plotting for better visualization. (b) Mineral physics predictions at 2,000‐km depth with varying H .
α is fixed at 0.2 and Q0 is derived following Matas and Bukowinski (2007) method assuming average lateral temperature variation dT/T = 0.4 and seismic observed
quality factor Qs = 350. (c) Similar as (b) but α is fixed at 0.4.
                                         may partially be due to the spin transition in a mantle with heterogeneity dominated by temperature
                                         variations. In Figure 14 we compare the predicted RV S =V P from Wu and Wentzcovitch (2014) to our
                                         results. Our result matches the mineral physics prediction for an aggregate consisting of 5 wt %
                                         ferropericlase (Mg0.92Fe0.08)SiO3 and the rest bridgmanite the best, which generally agrees with a
                                         bridgmanite‐enriched mantle model proposed by Murakami et al. (2012). However, the modeling result of
                                         Wu and Wentzcovitch (2014) may oversimplify the effect of the spin transition in the mantle since it does
                                         not consider several critical effects such as iron‐partitioning variations crossing the spin transition (e.g.,
                                                                            Kobayashi et al., 2005; Lin et al., 2013), and the spin transition in Fe‐
                                                                            bearing bridgmanite (e.g., Chantel et al., 2012; Fu et al., 2018). We leave
                                                                            to further work a more detailed analysis of the effect of the spin
                                                                            transition on the S to P heterogeneity ratio.
                                                                                Karato and Karki (2001) claim that RV S =V P cannot exceed 2.7 for an iso-
                                                                                chemical deep lower mantle. Our model, as well as several other mod-
                                                                                els shown in Figure 12, shows the bottom 200 km of the mantle to
                                                                                have heterogeneity ratios well above 2.7. One major limitation of the
                                                                                mineral physics predictions we use is that they did not consider poten-
                                                                                tial phase transitions occurring in the lower mantle. Bridgmanite could
                                                                                transform to pPv in the deepest ~400‐km mantle (Murakami et al.,
                                                                                2004; Oganov & Ono, 2004; Tsuchiya et al., 2004). Compared with
                                                                                bridgmanite, pPv has similar P velocity and higher S velocity
                                                                                (Tsuchiya et al., 2004; Wookey et al., 2005). Wentzcovitch et al.
                                                                                (2006) reported a very large RV S =V P (>6) caused by the pPv transition
                                                                                along their predicted phase boundary (Figure 14). Lateral variation in
                                                                                phase abundances can explain the high RV S =V P we found in our model
                                                                                in the bottom few hundred kilometers of the mantle. The amplitude
                                                                                differences between the results of Wentzcovitch et al. (2006) and seis-
Figure 14. Comparison of the S to P heterogeneity ratios in the TX2019slab
model and mineral physics predictions including the effects of the spin         mic observations could be caused by uncertainties in the mineral phy-
transition and the pPv phase transition. The mineral physics predictions        sics modeling, such as the pressure and temperature sensitivity of the
accounting for the spin transition (dash lines) are adopted from Wu and         bridgmanite to pPv phase transition (Cobden et al., 2015), and the pos-
Wentzcovitch (2014). Thermally induced S to P heterogeneity ratios in           sible presence of subducted MORB or iron at the lowest mantle
aggregates along adiabatic (Adi) and superadiabatic (SAdi) geotherms are
                                                                                (Grocholski et al., 2012; Tateno et al., 2007).
shown. Aggregates consist of 5 or 10 wt % ferropericlase (Mg0.92Fe0.08)SiO3
and bridgmanite. The predicted S to P heterogeneity ratio due to the bridg-     Figure 15 shows radial depth profiles of the correlation coefficient (CC)
manite to pPv transition along the phase boundary (solid black line) is         between P and S wave perturbations, the RMS velocity anomalies for P
shown with uncertainty (shaded purple area) from Wentzcovitch et al.
                                                                                and S, and the CC between bulk sound and shear speed perturbations.
(2006). The light yellow background color indicates the proposed depth
range for the spin transition effect and light background below shows the       P and S velocity heterogeneities are highly correlated throughout the
depths where the pPv phase transition could occur.                              mantle in our model (Figure 15a). The RMS amplitude of S wave
LU ET AL.                                                                                                                                                    11,563
                                                                                                                                             21699356, 2019, 11, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JB017448 by Cochrane Saudi Arabia, Wiley Online Library on [16/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
            Journal of Geophysical Research: Solid Earth                                                  10.1029/2019JB017448
                   Figure 15. Depth profiles calculated for the TX2019slab model. (a) Correlation between P and S wave heterogeneities. (b)
                   RMS amplitude of P and S wave anomalies. (c) Correlation between bulk and shear wave anomalies. Bulk and shear
                   speeds become decorrelated in the shallow lower mantle (~700‐ to ~1,500‐km depth) and anticorrelated below 1,500‐km
                   depth.
                   perturbations is higher than P waves at all depths, which agrees with previous studies (e.g., Koelemeijer
                   et al., 2016; Simmons et al., 2010; Figure 15b). The depth profile of the correlation between bulk sound
                   and shear speeds shows a more complex variation (Figure 15c). In the upper mantle, the bulk sound
                   speed correlates well with shear speed. In the shallow and mid‐lower mantle, the bulk sound
                   anomalies are almost decorrelated with S wave anomalies. Sound and shear speeds become
                   anticorrelated below ~1,500‐km depth. Similar anticorrelation of sound and shear speeds in the deep
                   lower mantle has been reported in several other tomography studies (e.g., Koelemeijer et al., 2016;
                   Masters et al., 2000; Simmons et al., 2010; Su & Dziewonski, 1997). The anticorrelation between
                   shear wave and bulk sound speed anomalies has been used to argue for chemical variations in the
                   deep mantle. However, previous investigations have shown that both the spin transition effect (Wu &
                   Wentzcovitch, 2014) and the phase transition from bridgmanite to pPv (Wookey et al., 2005) can
                   cause anticorrelation between sound and shear speeds in the deep mantle. Thus, the anticorrelation
                   we observe in our model could still be explained by thermal effects alone for an isochemical mantle,
                   although it likely requires a strong effect of the spin transition.
                   It is still challenging to draw concrete conclusions about the thermal‐chemical structure in the lower mantle
                   using P‐S tomography results alone. Although the observed RV S =V P is higher in the deep mantle than pre-
                   dicted by simple calculations assuming that heterogeneity is due to thermal effects alone, the effects of the
                   spin transition, phase changes, and anelasticity can increase the scaling beyond what has been considered
                   in the past for an isochemical mantle. Detailed evaluation of several critical topics are still lacking to defini-
                   tively rule out a thermal explanation for mantle heterogeneity. These include the effect of iron‐partitioning
                   variations crossing the spin transition (e.g., Kobayashi et al., 2005; Lin et al., 2013), the spin transition in Fe‐
                   bearing bridgmanite (e.g., Chantel et al., 2012; Fu et al., 2018), the pressure and temperature sensitivity of
                   the bridgmanite to pPv phase transition (Cobden et al., 2015), and the vertical smearing effect in tomography
                   models.
LU ET AL.                                                                                                                         11,564
                                                                                                                                                                                              21699356, 2019, 11, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JB017448 by Cochrane Saudi Arabia, Wiley Online Library on [16/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                             Journal of Geophysical Research: Solid Earth                                                                           10.1029/2019JB017448
                                          5. Conclusions
                                          In this study, we produce a new P and S global tomography model. The most significant improvement in our
                                          inversion, relative to our past work, is that we include a priori 3‐D subducting slabs in our starting model and
                                          inverted for velocity and source parameters simultaneously. Both of our P and S wave models feature higher‐
                                          amplitude subducting slabs compared with previous global tomography results, which better matches obser-
                                          vations using other approaches. We also calculated the S to P heterogeneity ratio using our model. Although
                                          the 1‐D S to P heterogeneity ratio is not significantly different using models with starting slabs relative to
                                          models without, we found that the incorrect imaging of subducting slabs could bias the tomography results
                                          in the lower mantle in specific regions. Our derived S to P ratio features a broad peak at around 1,800‐ to
                                          2,100‐km depth, which is consistent with mineral physics predictions of the spin transition effect. The high
                                          S to P ratio right above CMB can be due to bridgmanite to pPv phase change. However, our observations
                                          alone cannot rule out the possibility of the existence of chemical heterogeneities in the deep lower mantle.
Acknowledgments                           References
The TX2019slab model is available at
IRIS EMC website (http://ds.iris.edu/     Amaru, M. L. (2007). Global travel time tomography with 3‐D reference model (Doctoral dissertation). Utrecht, Netherlands: Utrecht
ds/products/emc‐earthmodels/). We            University.
thank ISC (International Seismological    Antolik, M., Gu, Y. J., Ekström, G., & Dziewonski, A. M. (2003). J362D28: a new joint model of compressional and shear velocity in the
Centre, http://www.isc.ac.uk/isc‐ehb/),      Earth's mantle. Geophysical Journal International, 153(2), 443–466.
IRIS DMC (Incorporated Research           Billen, M. I. (2008). Modeling the dynamics of subducting slabs. Annual Review Earth Planetary Sciences, 36, 325–356.
Institutions for Seismology, Data         Bonnardot, M. A., Régnier, M., Christova, C., Ruellan, E., & Tric, E. (2009). Seismological evidence for a slab detachment in the Tonga
Management Center, https://ds.iris.          subduction zone. Tectonophysics, 464(1), 84–99. https://doi.org/10.1016/j.tecto.2008.10.011
edu/ds/nodes/dmc/), ORFEUS                Brodholt, J. P., Helffrich, G., & Trampert, J. (2007). Chemical versus thermal heterogeneity in the lower mantle: The most likely role of
(Observations & Research Facilities for      anelasticity. Earth and Planetary Science Letters, 262(3), 429–437. https://doi.org/10.1016/j.epsl.2007.07.054
European Seismology, https://www.         Brudzinski, M. R., & Chen, W.‐P. (2005). Earthquakes and strain in subhorizontal slabs. Journal of Geophysical Research, 110(B8), B08303.
orfeus‐eu.org/), NECDC (Northern             https://doi.org/10.1029/2004JB003470
California Earthquake Data Center,        Chantel, J., Frost, D. J., McCammon, C. A., Jing, Z., & Wang, Y. (2012). Acoustic velocities of pure and iron‐bearing magnesium silicate
http://ncedc.org/), F‐net (F‐net             perovskite measured to 25 GPa and 1200 K. Geophysical Research Letters, 39, L19307. https://doi.org/10.1029/2012GL053075
Broadband Seismograph Network,            Chen, M., Tromp, J., Helmberger, D., & Kanamori, H. (2007). Waveform modeling of the slab beneath Japan. Journal of Geophysical
http://www.fnet.bosai.go.jp/), and           Research, 112, B02305. https://doi.org/10.1029/2006JB004394
CNSN (Canadian National Seismic           Cobden, L., Thomas, C., & Trampert, J. (2015). Seismic Detection of Post‐perovskite Inside the Earth. In A. Khan, & F. Deschamps (Eds.),
Network, http://www.                         The Earth's Heterogeneous Mantle (pp. 391–440). Switzerland: Springer.
earthquakescanada.nrcan.gc.ca/            Creager, K. C., & Jordan, T. H. (1984). Slab penetration into the lower mantle. Journal of Geophysical Research, 89, 3031–3049. https://doi.
stndon/CNDC/) for providing seismic          org/10.1029/JB089iB05p03031
data used in this study. We thank Georg   Della Mora, S., Boschi, L., Tackley, P., Nakagawa, T., & Giardini, D. (2011). Low seismic resolution cannot explain S/P decorrelation in the
Stadler, Michael Gurnis, and Ellen           lower mantle. Geophysical Research Letters, 38, L12303. https://doi.org/10.1029/2011GL047559
Syracuse for providing thermal slab       Deschamps, F., Trampert, J., & Snieder, R. (2002). Anomalies of temperature and iron in the uppermost mantle inferred from gravity
model. We also thank Nathan                  data and tomographic models. Physics of the Earth and Planetary Interiors, 129(3), 245–264. https://doi.org/10.1016/S0031‐9201(01)
Simmons, Alessandro Forte, Jung‐Fu           00294‐1
Lin, Thorsten Becker, and Suyu Fu for     Ding, X. Y., & Grand, S. P. (1994). Seismic structure of the deep Kurile subduction zone. Journal of Geophysical Research, 99(B12),
the valuable discussions. This research      23,767–23,786. https://doi.org/10.1029/94JB02130
was supported by National Science         Engdahl, E. R., van der Hilst, R. D., & Buland, R. (1998). Global teleseismic earthquake relocation with improved travel times and proce-
Foundation (NSF) grants EAR‐1648770          dures for depth determination. Bulletin of the Seismological Society of America, 88(3), 722–743.
and EAR‐1648817 and by the Jackson        Forte, A. M., & Perry, H. C. (2000). Geodynamic evidence for a chemically depleted continental tectosphere. Science, 290(5498), 1940–1944.
School of Geosciences at the University      https://doi.org/10.1126/science.290.5498.1940
of Texas at Austin.                       French, S. W., & Romanowicz, B. A. (2014). Whole‐mantle radially anisotropic shear velocity structure from spectral‐element waveform
                                             tomography. Geophysical Journal International, 199(3), 1303–1327. https://doi.org/10.1093/gji/ggu334
                                          Fu, S., Yang, J., Zhang, Y., Okuchi, T., McCammon, C., Kim, H.‐I., et al. (2018). Abnormal elasticity of Fe‐bearing bridgmanite in the
                                             Earth's lower mantle. Geophysical Research Letters, 45(10), 4725–4732. https://doi.org/10.1029/2018GL077764
                                          Fukao, Y., & Obayashi, M. (2013). Subducted slabs stagnant above, penetrating through, and trapped below the 660 km discontinuity.
                                             Journal of Geophysical Research: Solid Earth, 118, 5920–5938. https://doi.org/10.1002/2013JB010466
                                          Gaherty, J. B., Lay, T., & Vidale, J. E. (1991). Investigation of deep slab structure using long‐period S waves. Journal of Geophysical Research,
                                             96(B10), 16349–16367. https://doi.org/10.1029/91JB01483
                                          Garnero, E. J., McNamara, A. K., & Shim, S.‐H. (2016). Continent‐sized anomalous zones with low seismic velocity at the base of Earth's
                                             mantle. Nature Geoscience, 9, 481–489. https://doi.org/10.1038/ngeo2733
                                          Grand, S. P. (1994). Mantle shear structure beneath the Americas and surrounding oceans. Journal of Geophysical Research, 99,
                                             11,591–11,621. https://doi.org/10.1029/94JB00042
                                          Grand, S. P. (2002). Mantle shear–wave tomography and the fate of subducted slabs. Philosophical Transactions of the Royal Society of
                                             London A: Mathematical, Physical and Engineering Sciences, 360(1800), 2475–2491.
                                          Grand, S. P., & Helmberger, D. V. (1984). Upper mantle shear structure of North America. Geophysical Journal International, 76(2), 399–438
                                             . https://doi.org/10.1111/j.1365‐246X.1984.tb05053.x
                                          Grocholski, B., Catalli, K., Shim, S.‐H., & Prakapenka, V. (2012). Mineralogical effects on the detectability of the postperovskite boundary.
                                             Proceedings of the National Academy of Science, 109(7), 2275–2279.
                                          Hayes, G. P., Wald, D. J., & Johnson, R. L. (2012). Slab1. 0: A three‐dimensional model of global subduction zone geometries. Journal of
                                             Geophysical Research, 117(B1), B01302. https://doi.org/10.1029/2011JB008524
                                          Houser, C., Masters, G., Shearer, P., & 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. https://doi.org/10.1111/j.1365‐246X.2008.03763.x
LU ET AL.                                                                                                                                                                         11,565
                                                                                                                                                                     21699356, 2019, 11, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JB017448 by Cochrane Saudi Arabia, Wiley Online Library on [16/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
            Journal of Geophysical Research: Solid Earth                                                                    10.1029/2019JB017448
                   Ishii, M., & Tromp, J. (1999). Normal‐mode and free‐air gravity constraints on lateral variations in velocity and density of Earth's mantle.
                      Science, 285(5431), 1231–1236. https://doi.org/10.1126/science.285.5431.1231
                   Karato, S.‐i. (1993). Importance of anelasticity in the interpretation of seismic tomography. Geophysical Research Letters, 20(15), 1623–1626.
                      https://doi.org/10.1029/93GL01767
                   Karato, S.‐i., & Karki, B. B. (2001). Origin of lateral variation of seismic wave velocities and density in the deep mantle. Journal of
                      Geophysical Research, 106(B10), 21,771–21,783. https://doi.org/10.1029/2001JB000214
                   Kawakatsu, H., & Yoshioka, S. (2011). Metastable olivine wedge and deep dry cold slab beneath southwest Japan. Earth and Planetary
                      Science Letters, 303(1), 1–10. https://doi.org/10.1016/j.epsl.2011.01.008
                   Kellogg, L. H., Hager, B. H., & van der Hilst, R. D. (1999). Compositional Stratification in the Deep Mantle. Science, 283(5409), 1881–1884.
                      https://doi.org/10.1126/science.283.5409.1881
                   Kennett, B. L. N., Engdahl, E. R., & Buland, R. (1995). Constraints on seismic velocities in the Earth from traveltimes. Geophysical Journal
                      International, 122(1), 108–124. https://doi.org/10.1111/j.1365‐246X.1995.tb03540.x
                   Kobayashi, Y., Kondo, T., Ohtani, E., Hirao, N., Miyajima, N., Yagi, T., et al. (2005). Fe‐Mg partitioning between (Mg, Fe)SiO3 post‐
                      perovskite, perovskite, and magnesiowüstite in the Earth's lower mantle. Geophysical Research Letters, 32, L19301. https://doi.org/
                      10.1029/2005GL023257
                   Koelemeijer, P., Ritsema, J., Deuss, A., & van Heijst, H. J. (2016). SP12RTS: a degree‐12 model of shear‐ and compressional‐wave velocity
                      for Earth's mantle. Geophysical Journal International, 204(2), 1024–1039. https://doi.org/10.1093/gji/ggv481
                   Komatitsch, D., & Tromp, J. (2002a). Spectral‐element simulations of global seismic wave propagation—I. Validation. Geophysical Journal
                      International, 149(2), 390–412.
                   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(1), 303–318.
                   Kustowski, B., Ekström, G., & Dziewoński, A. (2008). Anisotropic shear‐wave velocity structure of the Earth's mantle: A global model.
                      Journal of Geophysical Research, 113, B06306. https://doi.org/10.1029/2007JB005169
                   Lai, H., Garnero, E. J., Grand, S. P., Porritt, R. W., & Becker, T. W. (2019). Global travel time dataset from adaptive empirical wavelet
                      construction. Geochemistry, Geophysics, Geosystems, 20. https://doi.org/10.1029/2018GC007905
                   Laske, G., Masters, G., Ma, Z., & Pasyanos, M. (2013). Update on CRUST1. 0‐A 1‐degree global model of Earth's crust. Paper presented at
                      EGU General Assembly, Vienna, Austria.
                   Lee, C. T. (2003). Compositional variation of density and seismic velocities in natural peridotites at STP conditions: Implications for seismic
                      imaging of compositional heterogeneities in the upper mantle. Journal of Geophysical Research, 108(B9), 2441, ECV 6‐1‐6‐20. https://doi.
                      org/10.1029/2003JB002413
                   Li, C., van der Hilst, R. D., Engdahl, E. R., & Burdick, S. (2008). A new global model for P wave speed variations in Earth's mantle.
                      Geochemistry, Geophysics, Geosystems, 9, Q05018. https://doi.org/10.1029/2007GC001806
                   Lin, J.‐F., Speziale, S., Mao, Z., & Marquardt, H. (2013). Effects of the electronic spin transitions of iron in lower mantle minearls: impli-
                      cations for deep mantle geophysics and geochemistry. Reviews of Geophysics, 51, 244–275. https://doi.org/10.1002/rog.20010
                   Lu, C., & Grand, S. P. (2016). The effect of subducting slabs in global shear wave tomography. Geophysical Journal International, 205(2),
                      1074–1085. https://doi.org/10.1093/gji/ggw072
                   Masters, G., Laske, G., Bolton, H., & Dziewonski, A. (2000). The Relative Behavior of Shear Velocity, Bulk Sound Speed, and Compressional
                      Velocity in the Mantle: Implications for Chemical and Thermal Structure. In S. i. Karato, A. Forte, R. Liebermann, G. Masters, & L.
                      Stixrude (Eds.), Earth's Deep Interior: Mineral Physics and Tomography From the Atomic to the Global Scale, Geophysical Monograph
                      Series (Vol. 117, pp. 63–87). Washington, DC: American Geophysical Union.
                   Matas, J., & Bukowinski, M. S. T. (2007). On the anelastic contribution to the temperature dependence of lower mantle seismic velocities.
                      Earth and Planetary Science Letters, 259(1), 51–65. https://doi.org/10.1016/j.epsl.2007.04.028
                   Mosca, I., Cobden, L., Deuss, A., Ritsema, J., & Trampert, J. (2012). Seismic and mineralogical structures of the lower mantle from prob-
                      abilistic tomography. Journal of Geophysical Research, 117, B06304. https://doi.org/10.1029/2011JB008851
                   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.
                   Murakami, M., Hirose, K., Kawamura, K., Sata, N., & Ohishi, Y. (2004). Post‐perovskite phase transition in MgSiO3. Science, 304(5672),
                      855–858. https://doi.org/10.1126/science.1095932
                   Murakami, M., Ohishi, Y., Hirao, N., & Hirose, K. (2012). A perovskitic lower mantle inferred from high‐pressure, high‐temperature sound
                      velocity data. Nature, 485, 90–94.
                   Obayashi, M., Yoshimitsu, J., Nolet, G., Fukao, Y., Shiobara, H., Sugioka, H., et al. (2013). Finite frequency whole mantle P wave
                      tomography: Improvement of subducted slab images. Geophysical Research Letters, 40, 5652–5657. https://doi.org/10.1002/
                      2013GL057401
                   Oganov, A. R., & Ono, S. (2004). Theoretical and experimental evidence for a post‐perovskite phase of MgSiO3 in Earth's D″ layer. Nature,
                      430(6998), 445–448.
                   Panning, M., & Romanowicz, B. (2006). A three‐dimensional radially anisotropic model of shear velocity in the whole mantle. Geophysical
                      Journal International, 167(1), 361–379. https://doi.org/10.1111/j.1365‐246X.2006.03100.x
                   Perry, H. K. C., Forte, A. M., & Eaton, D. W. S. (2003). Upper‐mantle thermochemical structure below North America from seismic–
                      geodynamic flow models. Geophysical Journal International, 154(2), 279–299. https://doi.org/10.1046/j.1365‐246X.2003.01961.x
                   Resovsky, J., & Trampert, J. (2003). Using probabilistic seismic tomography to test mantle velocity‐density relationships. Earth and
                      Planetary Science Letters, 215(1), 121–134. https://doi.org/10.1016/S0012‐821X(03)00436‐9
                   Richards, S., Holm, R., & Barber, G. (2011). When slabs collide: A tectonic assessment of deep earthquakes in the Tonga‐Vanuatu region.
                      Geology, 39(8), 787–790. https://doi.org/10.1130/G31937.1
                   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, 1223–1236. https://doi.org/10.1111/j.1365‐246X.2010.04884.x
                   Robertson, G., & Woodhouse, J. (1996). Ratio of relative S to P velocity heterogeneity in the lower mantle. Journal of Geophysical Research,
                      101(B9), 20041–20052. https://doi.org/10.1029/96JB01905
                   Romanowicz, B. (2001). Can we resolve 3D density heterogeneity in the lower mantle? Geophysical Research Letters, 28(6), 1107–1110.
                      https://doi.org/10.1029/2000GL012278
                   Saltzer, R. L., van der Hilst, R. D., & Karason, H. (2001). Comparing P and S wave heterogeneity in the mantle. Geophysical Research Letters,
                      28(7), 1335–1338. https://doi.org/10.1029/2000GL012339
LU ET AL.                                                                                                                                                11,566
                                                                                                                                                                     21699356, 2019, 11, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JB017448 by Cochrane Saudi Arabia, Wiley Online Library on [16/10/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
            Journal of Geophysical Research: Solid Earth                                                                    10.1029/2019JB017448
                   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, B12310. https://doi.org/10.1029/2010JB007631
                   Simmons, N. A., Myers, S. C., Johannesson, G., & Matzel, E. (2012). LLNL‐G3Dv3: Global P wave tomography model for improved regional
                      and teleseismic travel time prediction. Journal of Geophysical Research, 117, B10302. https://doi.org/10.1029/2012JB009525
                   Sleep, N. H. (1973). Teleseismic P‐wave transmission through slabs. Bulletin of the Seismological Society of America, 63(4), 1349–1373.
                   Stadler, G., Gurnis, M., Burstedde, C., Wilcox, L. C., Alisic, L., & Ghattas, O. (2010). The Dynamics of Plate Tectonics and Mantle Flow:
                      From Local to Global Scales. Science, 329(5995), 1033–1038. https://doi.org/10.1126/science.1191223
                   Stixrude, L., & Lithgow‐Bertelloni, C. (2012). Geophysics of Chemical Heterogeneity in the Mantle. Annual Review of Earth and Planetary
                      Sciences, 40(1), 569–595. https://doi.org/10.1146/annurev.earth.36.031207.124244
                   Su, W.‐j., & Dziewonski, A. M. (1997). Simultaneous inversion for 3‐D variations in shear and bulk velocity in the mantle. Physics of the
                      Earth and Planetary Interiors, 100(1), 135–156. https://doi.org/10.1016/S0031‐9201(96)03236‐0
                   Syracuse, E. M., van Keken, P. E., & Abers, G. A. (2010). The global range of subduction zone thermal models. Physics of the Earth and
                      Planetary Interiors, 183(1), 73–90. https://doi.org/10.1016/j.pepi.2010.02.004
                   Tao, K., Grand, S. P., & Niu, F. (2018). Seismic structure of the upper mantle beneath eastern Asia from full waveform seismic tomography.
                      Geochemistry, Geophysics, Geosystems, 19, 2732–2763. https://doi.org/10.1029/2018GC007460
                   Tateno, S., Hirose, K., Sata, N., & Ohishi, Y. (2007). Solubility of FeO in (Mg, Fe)SiO3 perovskite and the post‐perovskite phase transition.
                      Physics of the Earth and Planetary Interiors, 160(3), 319–325.
                   Tesoniero, A., Cammarano, F., & Boschi, L. (2016). S‐to‐P heterogeneity ratio in the lower mantle and thermo‐chemical implications.
                      Geochemistry, Geophysics, Geosystems, 17, 2522–2538. https://doi.org/10.1002/2016GC006293
                   Trampert, J., Vacher, P., & Vlaar, N. (2001). Sensitivities of seismic velocities to temperature, pressure and composition in the lower mantle.
                      Physics of the Earth and Planetary Interiors, 124(3), 255–267. https://doi.org/10.1016/S0031‐9201(01)00201‐1
                   Tsuchiya, T., Tsuchiya, J., Umemoto, K., & Wentzcovitch, R. M. (2004). Phase transition in MgSiO3 perovskite in the earth's lower mantle.
                      Earth and Planetary Science Letters, 224(3‐4), 241–248.
                   van Keken, P. E., Hacker, B. R., Syracuse, E. M., & Abers, G. A. (2011). Subduction factory: 4. Depth‐dependent flux of H2O from sub-
                      ducting slabs worldwide. Journal of Geophysical Research, 116, B01401. https://doi.org/10.1029/2010JB007922
                   Vidale, J. E. (1987). Waveform effects of a high‐velocity, subducted slab. Geophysical Research Letters, 14(5), 542–545. https://doi.org/
                      10.1029/GL014i005p00542
                   Wang, T., Revenaugh, J., & Song, X. (2014). Two‐dimensional/three‐dimensional waveform modeling of subducting slab and transition
                      zone beneath Northeast Asia. Journal of Geophysical Research: Solid Earth, 119, 4766–4786. https://doi.org/10.1002/2014jb011058
                   Wentzcovitch, R. M., Tsuchiya, T., & Tsuchiya, J. (2006). MgSiO3 postperovskite at D conditions. Proceedings of the National Academy of
                      Sciences of the United States of America, 103(3), 543–546. https://doi.org/10.1073/pnas.0506879103
                   Wookey, J., Stackhouse, S., Kendall, J.‐M., Brodholt, J., & Price, G. D. (2005). Efficacy of the post‐perovskite phase as an explanation for
                      lowermost‐mantle seismic properties. Nature, 438(7070), 1004.
                   Wu, Z., & Wentzcovitch, R. M. (2014). Spin crossover in ferropericlase and velocity heterogeneities in the lower mantle. Proceedings of the
                      National Academy of Sciences, 111(29), 10,468–10,472. https://doi.org/10.1073/pnas.1322427111
                   Yang, J., Lin, J.‐F., Jacobsen, S. D., Seymour, N. M., Tkachev, S. N., & Prakapenka, V. B. (2016). Elasticity of ferropericlase and seismic
                      heterogeneity in the Earth's lower mantle. Journal of Geophysical Research: Solid Earth, 121, 8488–8500. https://doi.org/10.1002/
                      2016JB013352
                   Zhan, Z., Helmberger, D. V., & Li, D. (2014). Imaging subducted slab structure beneath the Sea of Okhotsk with teleseismic waveforms.
                      Physics of the Earth and Planetary Interiors, 232, 30–35. https://doi.org/10.1016/j.pepi.2014.03.008
                   Zhao, D., Fujisawa, M., & Toyokuni, G. (2017). Tomography of the subducting Pacific slab and the 2015 Bonin deepest earthquake (Mw
                      7.9). Scientific Reports, 7, 44487. https://doi.org/10.1038/srep44487
LU ET AL. 11,567