Abstract
Large Low Shear Velocity Provinces (LLSVPs) near the core-mantle boundary (CMB) are key yet enigmatic structures. Their origin is often linked to the accumulation of subducted mid-ocean ridge basalt (MORB), but computational models question MORB as the sole source due to its predicted high shear wave velocity compared to normal mantle. This uncertainty is compounded by the lack of direct sound velocity measurements at CMB pressures. Here we address this gap through ultrahigh-pressure shear wave velocity measurements on CaCl2- and α-PbO2-type SiO2, major phases in MORB, at pressures exceeding those of the CMB. Our results show shear velocities in dense SiO2 phases are ~ 7–14% lower than previous predictions under these conditions. Incorporating these values into MORB models suggests that the typical seismic anomaly of −1.5% (δlnVS) observed in LLSVPs can be explained by ~ 23–33 vol.% oceanic crust along a cold slab geotherm, without invoking extreme thermal anomalies (+1500 K). Considering a subduction history exceeding 2 billion years, this scenario supports long-term MORB accumulation at the lowermost mantle. These findings provide new constraints on LLSVP composition and offer critical insights into deep mantle dynamics and the evolution of Earth’s interior.
Subject terms: Mineralogy, Seismology
Dense silicon dioxide phases have lower shear velocities than predicted, making subducted oceanic crust a viable Large Low Shear Velocity Provinces component, according to an analysis that uses ultrahigh-pressure Brillouin spectroscopy and thermodynamic modeling.
Main Text
The Large Low Shear Velocity Provinces (LLSVPs) are recognized as the most extensive seismic anomalies in the lower mantle, exhibiting negative shear wave velocity anomalies (δlnVS = δVS/VS*100%) of approximately 0.5% to 3.0% relative to the surrounding averaged seismic velocities1–9. These seismic anomalies are observed beneath Africa and the Pacific and extend thousands of kilometers horizontally and hundreds of kilometers vertically above the CMB region10,11. Clarifying the origin of the LLSVPs, which give rise to one of the largest seismic heterogeneities in the Earth’s mantle, is thus considered essential for gaining insights into the dynamics and evolution of Earth’s interior throughout its history. A number of possible origins for LLSVPs have been proposed to satisfy physical and chemical properties as inferred from seismic features, that are manifested as thermal, chemical, or thermochemical variations5,12. Waveform- and travel-time-based seismological studies that indicate a sharp decrease in shear wave velocity at the edges of LLSVPs12–14 suggest that the seismic characteristics of the LLSVPs are more plausibly explained by the presence of chemically distinct materials compared to the surrounding mantle3,15–17, rather than solely being caused by a high-temperature anomaly12. However, the possibility of a combination of thermal and chemical factors still cannot be fully ruled out.
Hypothesis: Accumulation of ancient subducted oceanic crust in the lowermost mantle
One of the most widely supported hypotheses among the models proposing that LLSVPs are composed of chemically distinct entities suggests that their origin is attributed to the long-term accumulation of ancient subducted oceanic crust in the lowermost mantle3,18,19. This hypothesis is supported by geodynamical simulations that predict accumulation of subducted oceanic crust to accumulate in LLSVPs regions2,8,20. The density of oceanic crust determined from high-temperature and high-pressure experiments was shown to be large enough to ensure gravitational stability in the lowermost mantle3,8,18. This implies that LLSVPs inferred from the accumulation of subducted oceanic crust could potentially maintain seismic heterogeneity persistently.
The first crucial step in verifying the plausibility of this hypothesis is to directly compare seismic wave velocities of subducted oceanic crust in the lowermost mantle with the seismic structures of the LLSVPs. However, in-situ high-pressure velocity measurements of constituent minerals of mid-ocean ridge basalt (MORB), which are bridgmanite (Bd), SiO2, CaSiO3-perovskite (CaPv), and Ca-ferrite-type phase (CF)18, at the extreme conditions down to the lowermost mantle, are rather limited. This hypothesis has thus so far primarily relied on verification through computational calculations21–25 or mineral physics approaches based on the elastic properties determined under ambient or very low-pressure conditions26–29. Both mineral physics approach and recent ab initio calculations30,31 have demonstrated that the shear wave velocity of MORB becomes higher than that of the surrounding mantle, under lowermost mantle conditions. This implies that the subducted oceanic crust alone might not be a plausible source for LLSVPs. Therefore, it has been suggested that explaining LLSVPs through MORB incorporation would require unrealistically high-temperature anomalies of 1500–1600 K at most8,32. On the other hand, recent measurements of elastic wave velocities for cubic CaPv26,27, conducted up to ~15–20 GPa, have indicated that the shear wave velocity of CaPv is slower than previously predicted from the computational works. This observation raises the compelling possibility that the reduced shear wave velocity of MORB induced by the CaPv could explain LLSVPs. Nevertheless, the determined stability field of CaPv, starting above 20 GPa33–35 remains experimentally unverified. Following the most recent first-principles calculations of36, the stability field of cubic CaPv starts at 50 GPa, 1200 K and remains stable throughout lower-mantle conditions. Furthermore, the substantial discrepancies between theoretical data37,38 and experimental results obtained under similar experimental setups and pressure conditions26,27,39 undermine the reliability of a robust argument on the seismic structures of the LLSVPs.
Among the constituent minerals of MORB, the SiO2 phase is thought to be the fourth most abundant major mineral, comprising ~17–20 vol.% in MORB assemblages18,40,41. However, within primary minerals present in the lower mantle, the SiO2 phase has been believed to be the hardest (in bulk and shear moduli) mineral, far surpassing even bridgmanite42. In addition to this, due to reported slow silicon diffusion43, it can preserve its hardness even at high temperatures, which implies that its shear wave velocity is considered one of the fastest among minerals in the Earth’s mantle. Therefore, in order to assess the shear wave velocity nature of MORB under lowermost mantle conditions, obtaining shear wave velocity data for the SiO2 phase down to the CMB is arguably a crucial experimental challenge that needs to be addressed. However, elastic wave velocity measurements for the high-pressure polycrystalline SiO2 phase has only been determined up to ~70 GPa and never been explored down to the lowermost mantle pressures28,44. A predominant high-pressure SiO2 phase, which is believed to have a CaCl2-type structure, is stably present under most of the lower mantle conditions. This phase is known to further undergo a transition to an α-PbO2-type structure under lowermost mantle conditions above ~120 GPa and 2400 K45–49. According to previous computational studies22–24,50, the shear wave velocity of the CaCl2-type SiO2 phase is notably higher than that of other mantle mineral phases. This suggests that it significantly contributes to the overall increased shear wave velocity structure of MORB, compared to the surrounding mantle. On the other hand, there is also a computational prediction suggesting that the transition from the CaCl2-type phase to the α-PbO2-type phase could cause a sharp drop in shear wave velocity of approximately 1–2% at the transition pressure24. This could possibly lead to a reduction in MORB’s shear velocity at the very bottom of the lower mantle. Yet, this phenomenon has not been experimentally verified. Therefore, making a quantitative assessment whether the ancient subducted MORB, which sank to the lowermost mantle, can indeed explain the seismic characteristics of LLSVPs, remains challenging.
In-situ ultrahigh-pressure shear wave velocity measurements of SiO2 phases
To address this issue, we determined the shear wave velocities of both the CaCl2-type and α-PbO2-type SiO2 phases under pressure conditions up to 148 GPa, well surpassing those of the lowermost mantle, using in-situ ultrahigh-pressure Brillouin scattering technique with a diamond anvil cell (DAC). This was conducted in combination with synchrotron X-ray diffraction (XRD) for high-pressure structural and stress analysis, as well as Raman scattering measurements (see Fig. 1, Materials and Methods, Figure S1–S5, Table S1). Two different starting materials were used to synthesize the CaCl2-type and α-PbO2-type SiO2 phases, respectively. The samples were employed to investigate the elastic properties of the high-pressure phases of SiO2 in the pure system and to discuss its implications in a MORB composition. The CaCl2-type SiO2 phase was synthesized in a DAC by compressing stishovite – pre-synthesized in a large-volume press – above 55 GPa. The α-PbO2-type SiO2 phase was directly synthesized by compressing synthetic α-cristobalite starting material above around 40 GPa in a DAC, following previous experimental studies47,51–54. The high-pressure SiO2 phases were confirmed as the intended target high-pressure phases through in-situ high-pressure synchrotron X-ray diffraction measurements (Figure S3, S4). The obtained lattice constants and unit cell volumes were found to be in good agreement with those reported by Grocholski et al.48 and supplemented with Raman scattering spectroscopic measurements (Figure S5, Table S2).
Fig. 1. Brillouin spectra of high-pressure SiO2 phases loaded with NaCl B2 as pressure medium in the lower mantle.
a, b Polycrystalline CaCl2-type SiO2 phase at 57 and 126 GPa and 300 K. c, d Polycrystalline α-PbO2-type SiO2 phase at 43 and 148 GPa and 300 K. Atomic distributions in the orthorhombic lattices of CaCl2-type (upper, green) and α-PbO2-type (lower, red) structures, where dark green/grey atoms represent silicon and green/red atoms represent oxygen89.
Figure 1 shows the representative high-pressure raw Brillouin scattering spectra from the two synthesized high-pressure SiO2 phases. These spectra were acquired in two independent series of measurements (Table S1, Figure S6), using centrally positioned samples in the DAC chambers. Although the samples were not annealed, stress conditions were evaluated at the synchrotron for the CaCl2-type and α-PbO2-type phases at 92 and 99 GPa, respectively, indicating deviatoric stresses of ~5–6 GPa (Figure S1, S2). To ensure the reliability of Brillouin data under ultrahigh-pressure conditions, we collected sharp, high-quality peaks by measuring multiple angular orientations and extending acquisition times over several days (Fig. 1, Table S1). Since deviatoric stress typically induces peak broadening in Brillouin spectra, the absence of systematic broadening in our data suggests that stress did not progressively increase with pressure. Both the lowest- and highest- pressure data points in Fig. 1 are within ±0.1 of the average FWHM value, consistent with the high-pressure synchrotron data indicating ~5–6 GPa of stress (Fig. 1, Figure S1, S2, Supplementary Text 5). For each Brillouin pressure point (Table S1), pressure was determined using the Raman T2g mode, measured at several spots within the central ~20 μm region of the probed sample both before and after each Brillouin acquisition. The values were averaged, and the standard deviation is reported in Table S1. A cross-check between Raman T2g-derived pressures and those obtained from the equation of state (EoS) of the NaCl B2 phase using synchrotron XRD at 92 GPa and 99 GPa (for the CaCl2-type and α-PbO2-type SiO2 phases, respectively) showed that Raman-based pressures were up to 2 GPa lower at these high-pressure points. The shear wave velocity profiles from the two phases are shown in Fig. 2 along with previous experimental and theoretical results24,28,50. All acquired spectra were subject to background subtraction, which resulted in a consistent VS reduction within 0.02 km/s in α-PbO2-type in all collected data, and within a maximum of 0.2 km/s in the CaCl2-type. The 3rd-order finite strain fits of the shear wave velocity data, shown by continuous lines in Fig. 2 (see Table S2), result in the shear moduli (G0) and its pressure derivatives (G0’) as follows: G0 = 180 ( ± 2) GPa, G0’ = 1.56 ( ± 0.02) for CaCl2-type SiO2 phase and G0 = 148 ( ± 2) GPa, G0’ = 1.67 ( ± 0.01) for α-PbO2-type SiO2 phase. Although the slope of the velocity profile of the α-PbO2-type SiO2 phase is far steeper than that of the CaCl2-type SiO2 phase, the α-PbO2-type SiO2 phase exhibits lower shear velocities than those of the CaCl2-type SiO2 phase throughout the pressure range that we explored (Fig. 2). This could possibly result in ~1.5% negative shear velocity contrast (ΔVS) under ambient conditions at the expected SiO2 phase transition pressure (120–125 GPa)45–48 (Figs. 2, 3). This could further potentially increase to ΔVS of ~3.0–3.2% under cold slab-55 or lower mantle geotherm56 (Fig. 4, Table S6) (see Modeling of VS profile of MORB for more details).
Fig. 2. Shear wave velocity profiles of high-pressure SiO2 phases under high-pressures.
Bold lines represent curves fitted using the third-order finite strain equation to experimental data from48 and VS data obtained in this study, averaged at each pressure. Error bars indicate uncertainties in pressure and VS. Dotted lines represent the shear wave velocity profiles from the computational studies24,50. Dashed line represents the fitted curve of shear wave velocity profile from the experimental study of28. The grey shaded area indicates the possible phase transition pressure range from CaCl2-type to α-PbO2-type SiO2 estimated from previous experiments45–48. The blue shaded area indicates the expected pressure range where the LLSVPs are primarily observed5.
Fig. 3. Shear wave velocity profiles of SiO2 and MORB assemblage at ambient temperature.
A Shear wave velocity profiles of SiO2 phase as a function of pressure, assuming a temperature variation of ±100 K. B Shear wave velocity profiles of MORB assemblage as a function of pressure, assuming a temperature variation of ±100 K. The adopted MORB composition and its molar fractions can be found in (Table S3). For CaPv phase, a recent experimentally-derived cubic phase26 was adopted. The blue shaded area indicates the expected pressure range where the LLSVPs are primarily observed5. The orange area represents the stability field range of α-PbO2-type SiO2 phase45–49.
Fig. 4. Shear wave velocity profiles of SiO2 and MORB assemblage under different mantle geotherms.
A Shear wave velocity profiles of SiO2 phase as a function of pressure along with the cold slab geotherm55 and lower mantle geotherm56. B Shear wave velocity profiles of MORB assemblage as a function of pressure along with the cold slab geotherm55 and lower mantle geotherm56. In the sequential order, the geotherms consider consistently averaged temperature variations of ±100 K and ±200 K, in comparison to the adopted geotherms used in the theoretical prediction of MORB VS profiles31 (Fig. S7). The adopted MORB composition and its molar fractions can be found in (Table S3). For CaPv phase, a recent experimentally-derived cubic phase26 was adopted. The comparison of all adopted geotherms and resultant negative shear wave velocity contrast (ΔVS) (with the expected SiO2 phase transitions) in both systems can be found in (Tables S4-S8). The blue shaded area indicates the expected pressure range where the LLSVPs are primarily observed5. The orange area represents the stability field range of α-PbO2-type SiO2 phase45–49.
As shown in Fig. 2, the most significant difference between our results and the prior theoretical studies24,50 is that our shear wave velocity VS values are on average lower by 7–14%. It should be noted that the ferroelastic transition along with the stishovite to the CaCl2-type phase transformation was not observed under the pressure conditions that we explored. This is anticipated from theory to induce a sharp shear softening24, and was also shown in an X-ray diffraction and Brillouin study of a single-crystal stishovite at 55 GPa by Zhang et al., 202144 or polycrystalline stishovite from Brillouin study of Asahara et al. 201328 at the pressure range of 25–35 GPa. Starting this study at a relatively high pressure of 55 GPa in polycrystalline SiO2 (CaCl2-type), similar to the material examined by Asahara et al. 201328, may explain why the ferroelastic transition - previously observed at 25–35 GPa - was not detected in our work. Comparison with previous experimental results on CaCl2-type SiO2 up to 60 GPa28, which determined G0 as 179 ( ± 3) GPa and G0’ as 1.80 ( ± 0.06), shows that the value of G0’ in our study is approximately 15% lower. In contrast G0 closely aligns with the previous study28. However, extrapolating data from28 would lead to a remarkable difference in shear wave velocity at the lowermost mantle pressure condition, reaching up to 5% (Fig. 2). The notable difference in G0’ may be attributed to the lower pressures examined in the previous study28, compared to those prevailing at the CMB. Asahara et al. 201328 performed only a single velocity measurement at each pressure point, keeping the cell orientation fixed throughout. In contrast, we conducted multiple velocity measurements at each pressure point using different cell orientations. When velocity measurements are conducted at a fixed angle throughout a single series of measurements, the fundamental assumption in polycrystalline measurements - that the sample is a randomly oriented, fine-grained aggregate - may no longer hold. This is especially true if the grain size or crystallographic texture has developed significantly within the sample. In such cases, there is a considerable risk that the measured velocities will deviate substantially from the true average velocity representative of a polycrystalline aggregate. In contrast, to minimize such concerns, we performed measurements at multiple orientations under the same pressure condition. This practice has been routinely adopted in previous studies using Brillouin scattering on polycrystalline samples, and we believe this approach improves the reliability of our measurements compared to those of Asahara et al. 201328, whose methodology did not include this standard step. Therefore, the results of Asahara et al. 201328 may have been influenced by the development of a preferred crystallographic orientation or texturing, potentially explaining the observed ~0.3 km/s higher average VS profile over ~30–130 GPa (Fig. 2). The differences are less discernible at lower pressures but become more pronounced in the extended range of ~60–130 GPa. The precise factors responsible for the significant inconsistency with theoretical results remain uncertain. Nonetheless, an important aspect to emphasize is that theoretical predictions for the G0 of SiO2 phase (either stishovite or CaCl2-type) fail to adequately replicate the established value determined through experimental studies conducted under ambient conditions, diverging from it by approximately 18%24,25. A similar discrepancy on CaPv between computational37,57 and recent experimental results was observed26,27. Furthermore, a noteworthy disparity arises in estimating the phase transition pressure from CaCl2-type to α-PbO2-type, with a 10–20 GPa variation established through X-ray experimental results45,47,48.
Modeling of VS profile of MORB
By combining previously reported thermodynamic parameters25 (Table S4) with the present shear wave velocity data (Fig. 2, Table S1), we modeled the shear wave velocity profiles in the lowermost mantle. This model considered the potential presence of our experimentally determined dense high-pressure SiO2 phases within the ancient subducted oceanic crust under relevant high-pressure and ambient temperature (Fig. 3). The assessment of potential high-temperature conditions was modeled in Fig. 4, and compared to the 1-dimensional seismic model (PREM)58 and the theoretically predicted shear wave velocity profiles of MORB31, which were adopted as reference points to assess the potential contribution of SiO2 phases (expressed as negative shear wave velocity contrast across discontinuities (ΔVS)) to the observed negative anomalies (δlnVS) within LLSVPs (see Fig. 4, Figure S7).
For this purpose, we applied two anticipated temperature profiles for the subducting slab and the surrounding lower mantle, respectively55,56 (Figure S8). The mineral assemblage, molar fractions, and chemical composition in the MORB system under lower mantle conditions were adopted from the previous experimental results18 (Table S3). All thermo-elastic parameters for the constituent minerals of the MORB used in this modeling are presented in (Tables S4, S5), and are exclusively derived from previous high-pressure experiments18. The resultant shear wave velocity profiles of the CaCl2-type and α-PbO2-type SiO2 phases, as well as the MORB, as a function of pressure up to the CMB condition, considering the two distinct temperature conditions, are shown in Fig. 4. We found that the phase transition from CaCl2-type to α-PbO2-type in the MORB and pure SiO2 system could lead to discontinuous shear wave velocity reductions, reaching a maximum of ΔVS of 0.6% and 3.2%, respectively (Fig. 4, Table S6). If we assume the modeled MORB under ambient temperature (Fig. 3), the contribution of SiO2 phase transition, as the negative shear velocity contrast across a discontinuity feature in MORB, would decrease to ΔVS of ~0.33%. Our experimental results reveal that the CaCl2-α-PbO2 phase transition in SiO2 produces a negative velocity discontinuity. At the transition pressure in pure SiO2, along the cold slab geotherm, this discontinuity corresponds to a decrease of approximately 3% in shear wave velocity. However, when considered in the context of MORB compositions, this reduction is mitigated to about 0.6%, which is relatively minor compared to the overall discrepancy of ~7–14% between the experimentally determined SiO2 velocity profile and theoretical predictions. This highlights the significant role of the overall velocity reduction in SiO2, which has a greater impact on the seismic velocity profile than the phase transition alone. The magnitude of the negative shear wave velocity contrast across a discontinuity (ΔVS), as determined in this study, closely aligns with the results of previous theoretical calculations25 (Tables S4, S6). It is known that there is some discrepancy in the elastic properties of cubic-CaPv, as determined in two previous high-pressure experimental studies26,27. However, the results from the study26, characterized by relatively larger G0 and smaller G0’ values, than those from the other research27, effectively counterbalance the velocity differences under high-pressure conditions (Tables S4, S6, Figure S9). As a result, the choice between the two options did not change the final MORB seismic structures, as shown in (Table S6, Figure S9). We also modeled that the expected variations in the bulk moduli (K0) and its pressure derivatives (K0’), as reported in previous experimental studies on SiO2 high-pressure phases46,48,49,51,54, might not result in the final negative shear velocity contrast (ΔVS) variations exceeding 2.8–3.6% in the pure SiO2 system and 0.5-0.7% in MORB, respectively (see Tables S3, S6, S7). Assuming that the α-PbO2 phase in MORB reaches an abundance of up to 23 wt.% in the lowermost mantle18, a more pronounced shear wave velocity reduction of 1.1–1.3% (ΔVS) in MORB system might be expected (Tables S3, S6, S8).
Comparison with PREM
Comparative analysis with PREM eventually reveals that the pure SiO2 and the MORB systems might lead to a shear velocity decrease on average of approximately 4% and 7%, respectively, following a cold slab geotherm (Fig. 4, Figure S7). On the contrary, as shown in the supplementary materials (Figure S7), considering the results employing the elastic properties of SiO2 phases derived from the theoretical studies25, both the pure SiO2 system and the MORB system (only in case of following the cold slab geotherm) could exhibit shear velocities higher than those projected by PREM. Given these results, the newly acquired elasticity data of SiO2 high-pressure phases under extreme pressure conditions can offer vital insights into a more comprehensive understanding of the seismic structure of LLSVPs.
Comparison with theoretical models and experimental data
Theoretical G0 and G0’ values tend to be overestimated compared to experimental values, such as in case of CaPv26,27 or our experimentally determined SiO2 phases in this study. Consequently, using these theoretical parameters in modeled MORB compositions, particularly for dominant phases, can lead to an elevated VS profile and an overestimated MORB volume fraction required to explain the observed −1.5% anomaly (δlnVS). Additionally, discrepancies between studies, such as our work and those by Thomson et al. 201927 and Wang et al. 202031, may arise from differences in the partitioning behavior and proportions of mineral phases used in MORB modeling. In this study, we assessed the overall velocity profile of MORB based on experimentally determined mineral phases and incorporated updated constraints on G0 and G0’ for all bridgmanite endmembers (Table S4). The lack of high-temperature experimental data on MORB’s mineral phases under lower mantle conditions presents challenges in reconciling discrepancies with previous studies. While Mattern et al. 200559 emphasized the primary influence of pressure on the VS profile, the role of G0’ remains critical for interpreting seismic observations. Furthermore, previous studies25,60 suggest that temperature sensitivity in ηS0 can introduce experimental uncertainties of ∼10%. In our study, such uncertainties correspond to a ± 1 vol.% variation in MORB content required to reproduce the seismically observed anomalies (δlnVS) of −1.5% to −3%, via modeled negative shear wave velocity contrast (ΔVS) of a similar magnitude.
If the CaPv content in MORB increases by 7 wt.% (from 23 wt.% in Table S3 to 30 wt.%), as suggested by Ricolleau et al. 201041, while the SiO2 phase decreases from 17 wt.% to 10 wt.% - a value significantly lower than those reported in other studies e.g., Hirose et al. 2005; Perrillat et al., 2006; Ricolleau et al. 2010; Ishii et al., 202218,41,61,62 - then the same MORB volume fraction could still account for the observed seismic anomalies (δlnVS). This adjustment would slightly reduce the SiO2 contribution, as the negative shear velocity contrast in MORB, from ΔVS of ∼0.6% (Fig. 4, Table S6) to ΔVS of ∼0.5%.
Our study presents a comprehensive analysis of the potential discontinuous feature associated with MORB in the lower mantle (Table S9). The MORB volume fractions required to explain the observed seismic velocity anomalies are significantly lower than those proposed by previous studies - for example, Thomson et al., 201927 estimated up to 64% MORB at 100 GPa and 48% at 125 GPa to explain a δlnVS of −1.5%. In contrast, our results indicate that 23–33 vol.% MORB may already account for a δlnVS of −1.5% (Table S9, Fig. 4). A more pronounced anomaly δlnVS of -3% would require a higher fraction of ~47–66 vol.%. Importantly, these values decrease under elevated temperatures exceeding 3000 K (Table S9). For instance, if lower mantle temperatures range from 2600 K at the top of the D″ layer to 4000 K at the core-mantle boundary, as suggested by Manthilake et al. 201163, the modeled shear wave velocity contrast (ΔVS) could increase even by a factor of two (e.g., from ~-0.6% at <2600 K to ~−0.7% at 3000 K and ~−1% at 4000 K). This increase in ΔVS would in turn reduce the required MORB volume fraction - for example, from 33 vol.% under a cold slab geotherm (Case 2 A) to 22 vol.% at 3000 K and as low as 14 vol.% at 4000 K (Table S9).
These results highlight the importance of experimental constraints in refining theoretical models and interpreting seismic profiles in the lower mantle. Although the effect of the negative shear velocity contrast in MORB, with the SiO2 phase transition (CaCl2-type to α-PbO2-type) is relatively small (ΔVS of ∼0.6–1%, Tables S6– S8) under assumed cold slab or lower mantle geotherms - and further diminishes to ΔVS of ∼0.33% under assumed ambient conditions (Fig. 3) - the collective contribution of MORB phases with experimentally refined G0’ values provide valuable insight into how a realistic VS profile of MORB can decrease relative to the PREM model. This study contributes to understanding the potential role of MORB in explaining the negative anomalies (δlnVS) in the range of -(1.5–3)%.
Origin of LLSVPs
It is generally accepted from previous shear wave tomographic observations that the negative shear velocity anomalies in LLSVPs exhibit depth-dependent variations with anomalies ranging from approximately −0.5% to −1% in shallow regions3,5,7,8 to around −3% at depths ranging from 100 to 200 km from the CMB6,12,14 (Fig. 5). If such depth-dependent shear velocity anomalies are simply attributed to the contamination of MORB lithology into the surrounding mantle, near the bottom of the lowermost mantle, and without relying on temperature anomalies, a −1.5% anomaly (δlnVS) within the LLSVP’s stability field, where SiO2 phase undergoes a transition from CaCl2-type to α-PbO2-type, may be explained by the presence of approximately ~23–33 vol.% MORB (Fig.4, Tables S6–S9). Assuming that the CaCl2-α-PbO2-type phase transition leads to silica enrichment in the lower mantle18 (Tables S8, S9), 22 vol.% of MORB could be sufficient to account for a −1.5% (δlnVS) anomaly in a cold slab. Instead, the distinctive δlnVS of −3% observed specifically at the bottom of the mantle could be explained by the combined effect of temperature effect from cold slab or lower mantle geotherms, phase transition of CaCl2- to α-PbO2-type phase of SiO2 in the accumulated MORB pile, and/or variation of MORB pile volume fraction (Tables S7–S9). Since the exact temperature around the CMB has not been conclusively established, it is suggested that the temperature near the base of the mantle deviates significantly from the adiabatic temperature gradient, possibly reaching around 3000 K to 4000 K64–66. For instance, if we assume a temperature of ~3890–3900 K at the very bottom of the lower mantle, the observed shear wave velocity anomaly (δlnVS) of −3% can be reasonably attributed to the presence of ~21–28 vol.% MORB (Table S9).
Fig. 5. Conceptual schematic of Pacific LLSVPs and related features.
The negative shear velocity anomalies (δInVS) in the LLSVPs exhibit depth-dependent variations with anomalies ranging from approximately −0.5% to −1% in shallow regions3,5,7,8 and around -3% at the base (bottom 100–200 km)6,12,14. The present study demonstrates that averaged δInVS typically observed at LLSVPs (−1.5%) can be explained by the presence of ~23−33 vol.% of MORB, with a discontinuous feature across SiO2 phase transition, and without relying on thermal anomalies ( + 1500 K). The distinctive δInVS of −3% observed specifically at the bottom of the mantle could be explained by the combined effect of high-temperature, phase transition of CaCl2- to α-PbO2-type phase of SiO2 in the accumulated MORB pile, and/or variation of MORB pile volume fraction. This conceptual representation is based on interpretations discussed in He & Wen (2009, 2012)11,90, including features developed in this work.
Depending on the temperature conditions at the CMB, it is possible that just above the CMB, temperature surpasses the solidus of MORB64,67, and contributes to the potential formation of partial melt. These factors might be significant contributors to the formation of the Ultra-Low Velocity Zones (ULVZs) at the CMB65, but have not been investigated in more detail in this study. Instead, this study focused on the combination of factors such as temperature, an abundance of α-PbO2 phase in MORB (up to 23 wt.% in the lowermost mantle), and the MORB volume fractions variations on the seismic structure of LLSVPs. The results of this analysis, presented in (Table S9), suggest that the temperature effect of ~3000–4000 K gains significance in generating a larger anomaly, such as −3% (δlnVS), and thus lowering the vol.% of MORB below 44%. This could be attributed to the unique high-temperature conditions near the CMB. Having considered all factors, it is also noteworthy that in the mineral assemblage of both MORB and the ambient mantle, the post-perovskite phase transition could occur at around 113–120 GPa3,18,68. Despite the fact, that post-perovskite can potentially form within a close range to the examined SiO2 phase transition in this study, it is also well known that the shear velocity jump across the post-perovskite phase transition strongly depends on its texture development69,70. Thus, in case we assume an elastically isotropic condition, no significant positive jump in shear velocity VS should be expected68. If detected, however, the effect of the post-perovskite phase transition might not be that significant. Therefore, the present study demonstrates that the lower magnitude of LLSVPs’ anomaly (δlnVS) (−1.5%) can be explained by the presence of only around ~23–33 vol.% of MORB, without relying on thermal anomalies (Fig. 5, Table S9). In addition to this, the discrepancy between theoretically-predicted24,50 and experimentally-derived shear velocities in SiO2 in this study was revealed.
Among the subduction process hypothesis, proposing that chemical heterogeneity is a primary cause of LLSVPs3,15–17, there are alternative explanations for such a cause. One of them suggests that they could be attributed to primordial residues inherited from early differentiation processes in the Earth. This hypothesis proposes that the low shear velocity anomalies within LLSVPs are caused by high-density phases enriched in iron that selectively crystallized from the basal magma ocean in early Earth71,72. This hypothesis can also account for the low shear velocity and high-density features of LLSVPs. While we cannot disregard the importance of the primordial heterogeneity hypothesis proposed by several geodynamic studies2,8,20,73,74, it remains unclear whether a sufficient amount of primordial residues near the CMB can be created or preserved to account for the size of LLSVPs. Another, more recent hypothesis suggests that LLSVPs formed through a giant-impact scenario involving Theia colliding with proto-Earth75. This scenario involves entraining both molten and solidified fractions, which were subsequently sunk into the LLSVPs regions. These two hypotheses, however make it challenging to provide a strong inherent explanation for why LLSVPs are specifically located beneath Africa and the Pacific. On the other hand, considering at least 2 billion years of the history of subduction processes76 into the Earth’s interior, the presence of ~23–33 vol% of oceanic crust within LLSVPs appears to be a feasible accumulation (Table S9)77,78.
The findings of this study, which demonstrate that relatively small amounts of oceanic crust can explain seismological characteristics of LLSVPs, it is worth noting that previous research has also highlighted the greater density of oceanic crust compared to the surrounding mantle, which could maintain seismic heterogeneity persistently18,40. Considering these factors collectively, this hypothesis seems to be a suitable explanation for the origin of LLSVPs.
Furthermore, the presence of a large-scale, compositionally distinct basaltic pile with a laterally skewed distribution directly above the core-mantle boundary may introduce lateral heterogeneity to the heat transport mechanism from the core2,74,79. This may imply the possibility of receiving an enhanced thermal flux in certain regions, although without creating abrupt thermal anomalies of +1500 K, and potentially lead to significant mantle upwelling flows. The heterogeneity in heat transport mechanism is expected to partially contribute, along with the process of subducted MORB material, to the shear velocity reduction in LLSVPs at the very bottom region.
The long-term accumulation of the chemically distinct subducted oceanic crust in the lowermost mantle is thus anticipated to alter the bulk chemistry of the lower mantle towards more SiO2- and Al2O3-rich composition over the subduction history. The effect of Al2O3 across stishovite to CaCl2-type SiO2 phase transition was shown to reduce the transition pressure (e.g., refs. 80–83) and to decrease the bulk modulus (K) of SiO2 phases (e.g., see refs. 48,84–87). However, the shear properties of Al-bearing CaCl2-type or α-PbO2-type SiO2 phases remain experimentally unconstrained. For example, Lakshtanov et al. 200785 investigated Al- and H-bearing stishovite at room pressure up to ~25 GPa and reported a decrease in shear modulus (G) compared to pure room pressure SiO2 (e.g., Jiang et al. 200988). These results, however, are limited to the stishovite stability field and involve a minor presence of water. The current absence of experimental data on the shear modulus (G) and its pressure derivative (G’) for Al-bearing CaCl2-type and α-PbO2-type SiO2 phases, particularly under anhydrous conditions, highlights the importance of further high-pressure elasticity studies to quantify the effect of Al2O3 on the seismic properties of these high-pressure SiO2 phases.
Reporting summary
Further information on research design is available in the Nature Portfolio Reporting Summary linked to this article.
Supplementary information
Acknowledgements
This research was supported by ETH Zurich start-up funding (PSP1-001828-000) and the Swiss National Science Foundation (Grant No. 200021_197187) awarded to Motohiko Murakami. Synchrotron X-ray diffraction experiments were performed at the European Synchrotron. Radiation Facility (ESRF), beamline ID27, Grenoble, France. We thank E. Ito (sample synthesis using large volume press), T. de Selva-Dewint (Brillouin scattering measurements), R. Popa, M. Lüder (FT-IR measurements), N. Ma, S. Merkel, T. Poreba, M. Mezouar (synchrotron XRD measurements) for their experimental assistance. Y. Mori is acknowledged for his valuable comments and suggestions.
Author contributions
Conceptualization: M.M.; Methodology: E.K., M.M.; Investigation: E.K., M.M., P.S., C.L.; Visualization: E.K., M.M.; Funding acquisition: M.M.; Supervision: M.M., P.S., C.L.; Writing – original draft: E.K., M.M.; Writing – review & editing: E.K., M.M., P.S. and C.L.
Peer review
Peer review information
Communications Earth & Environment thanks the anonymous reviewers for their contribution to the peer review of this work. Primary Handling Editors:João Duarte, Martina Grecequet, Carolina Ortiz Guerrero. A peer review file is available.
Data availability
All data supporting the findings of this study are available in the main text and Supplementary Information.
Code availability
No custom code was developed for this study. Data analyses were carried out using OriginPro. 2023, Python 3.10 (including NumPy, SciPy, and Matplotlib), and the open-source BurnMan. package (version 1.1.0). Dioptas and PDIndexer were used for X-ray diffraction data integration, and indexing. All input parameters necessary to reproduce the results are provided in the main text and Supplementary Information.
Competing interests
The authors declare no competing interests.
Footnotes
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Contributor Information
Ewa Krymarys, Email: ewa.krymarys@eaps.ethz.ch.
Motohiko Murakami, Email: motohiko.murakami@eaps.ethz.ch.
Supplementary information
The online version contains supplementary material available at 10.1038/s43247-025-02700-8.
References
- 1.Lay, T., Hernlund, J., Garnero, E. J. & Thorne, M. S. A post-perovskite lens and D″ heat flux beneath the central Pacific. Science314, 1272–1276 (2006). [DOI] [PubMed] [Google Scholar]
- 2.Garnero, E. J. & McNamara, A. K. Structure and dynamics of earth’s lower mantle. Science320, 626–628 (2008). [DOI] [PubMed] [Google Scholar]
- 3.Ohta, K., Hirose, K., Shimizu, K., Lay, T. & Cruz, S. Phase transitions in pyrolitic mantle and MORB materials and their electrical conductivities at lowermost mantle conditions. Earth Planet. Sci. Lett.267, 107–117 (2008). [Google Scholar]
- 4.Kawai, K. & Geller, R. J. The vertical flow in the lowermost mantle beneath the Pacific from inversion of seismic waveforms for anisotropic structure. Earth Planet. Sci. Lett.297, 190–198 (2010). [Google Scholar]
- 5.Ballmer et al. Compositional layering within the large low shear-wave velocity provinces in the lower mantle. Geochem. Geophys. Geosystems17, 1312–1338 (2016). [Google Scholar]
- 6.Shephard, G. E., Matthews, K. J., Hosseini, K. & Domeier, M. On the consistency of seismically imaged lower mantle slabs. Sci. Rep.7, 1–17 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Thorne, M. S., Zhang, Y. & Ritsema, J. Evaluation of 1-D and 3-D seismic models of the Pacific lower mantle with S, SKS, and SKKS traveltimes and amplitudes. J. Geophys. Res. Solid Earth118, 985–995 (2013). [Google Scholar]
- 8.Jones, T. D., Maguire, R. R., van Keken, P. E., Ritsema, J. & Koelemeijer, P. Subducted oceanic crust as the origin of seismically slow lower-mantle structures. Prog. Earth Planet. Sci.7, 1–16 (2020). [Google Scholar]
- 9.Wang et al. Formation of large low shear velocity provinces through the decomposition of oxidized mantle. Nat. Commun.12, 1–10 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Konishi, K., Kawai, K., Geller, R. J. & Fuji, N. MORB in the lowermost mantle beneath the western Pacific: Evidence from waveform inversion. Earth Planet. Sci. Lett.278, 219–225 (2009). [Google Scholar]
- 11.He, Y. & Wen, L. Geographic boundary of the ‘Pacific Anomaly’ and its geometry and transitional structure in the north. J. Geophys. Res. Solid Earth117, 1–16 (2012). [Google Scholar]
- 12.Wen, L. An SH hybrid method and shear velocity structures in the lowermost mantle beneath the central Pacific and South Atlantic Oceans. J. Geophys. Res. 107, 2055 (2002).
- 13.Konishi, K., Kawai, K., Geller, R. J. & Fuji, N. Waveform inversion of broad-band body wave data for the S-velocity structure in the lowermost mantle beneath the Indian subcontinent and Tibetan Plateau. Geophys. J. Int.191, 305–316 (2012). [Google Scholar]
- 14.Fan, A., Sun, X., Zhang, Z., Zhang, P. & Zong, J. From Subduction to LLSVP: The Core-Mantle Boundary Heterogeneities Across North Atlantic. Geochem., Geophys. Geosystems23, 1–12 (2021). [Google Scholar]
- 15.McNamara, A. K. & Zhong, S. Thermochemical structures within a spherical mantle: Superplumes or piles?. J. Geophys. Res. Solid Earth109, 1–14 (2004). [Google Scholar]
- 16.Zhang, N., Zhong, S., Leng, W. & Li, Z. X. A model for the evolution of the Earth’s mantle structure since the Early Paleozoic. J. Geophys. Res. Solid Earth115, 1–22 (2010). [Google Scholar]
- 17.Bower, D. J., Gurnis, M. & Seton, M. Lower mantle structure from paleogeographically constrained dynamic Earth models. Geochem., Geophys. Geosystems14, 44–63 (2013). [Google Scholar]
- 18.Hirose, K., Takafuji, N., Sata, N. & Ohishi, Y. Phase transition and density of subducted MORB crust in the lower mantle. Earth Planet. Sci. Lett.237, 239–251 (2005). [Google Scholar]
- 19.Walter, M. J. et al. Deep mantle cycling of oceanic crust: Evidence from diamonds and their mineral inclusions. Science334, 54–57 (2011). [DOI] [PubMed] [Google Scholar]
- 20.Garnero, E. J., McNamara, A. K. & Shim, S. H. Continent-sized anomalous zones with low seismic velocity at the base of Earth’s mantle. Nat. Geosci.9, 481–489 (2016). [Google Scholar]
- 21.Karki B. B., Warren M. C., Stixrude L. & Ackland G. J. Ab initio studies of high-pressure structural transformations in silica. Phys. Rev. B - Condens. Matter Mater. Phys. 55, 3465–3471 (1997).
- 22.Karki B. B., Warren M. C., Stixrude L., Ackland G. J., Crain J. Erratum: Ab initio studies of high-pressure structural transformations in silica. Phys. Rev. B - Condens. Matter Mater. Phys. 56, 2883 (1997).
- 23.Karki, B. B., Stixrude, L. & Vrain, J. Ab initio elasticity of three high-pressure polymorphs of silica. Geophys. Res. Lett.24, 3269–3272 (1997). [Google Scholar]
- 24.Karki, B. B., Stixrude, L. & Wentzcovitch, R. M. High-pressure elastic properties of major materials of earth’s mantle from first principles. Rev. Geophys.39, 507–534 (2001). [Google Scholar]
- 25.Stixrude, L. & Lithgow-Bertelloni, C. Thermodynamics of mantle minerals - II. Phase equilibria. Geophys. J. Int.184, 1180–1213 (2011). [Google Scholar]
- 26.Gréaux, S. et al. Sound velocity of CaSiO3 perovskite suggests the presence of basaltic crust in the Earth’s lower mantle. Nature565, 218–221 (2019). [DOI] [PubMed] [Google Scholar]
- 27.Thomson, A. R. et al. Seismic velocities of CaSiO3 perovskite can explain LLSVPs in Earth’s lower mantle. Nature572, 643–647 (2019). [DOI] [PubMed] [Google Scholar]
- 28.Asahara, Y. et al. Acoustic velocity measurements for stishovite across the post-stishovite phase transition under deviatoric stress: Implications for the seismic features of subducting slabs in the mid-mantle. Am. Mineral.98, 2053–2062 (2013). [Google Scholar]
- 29.Dai, L. et al. Sound velocities of Na0.4Mg0.6Al1.6Si0.4O4 NAL and CF phases to 73 GPa determined by Brillouin scattering method. Phys. Chem. Miner.40, 195–201 (2013). [Google Scholar]
- 30.Kono, Y., Irifune, T., Ohfuji, H., Higo, Y. & Funakoshi, K. I. Sound velocities of MORB and absence of a basaltic layer in the mantle transition region. Geophys. Res. Lett.39, 1–5 (2012). [Google Scholar]
- 31.Wang et al. Velocity and density characteristics of subducted oceanic crust and the origin of lower-mantle heterogeneities. Nat. Commun.11, 1–8 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Deschamps, F., Cobden, L. & Tackley, P. J. The primitive nature of large low shear-wave velocity provinces. Earth Planet. Sci. Lett.349–350, 198–208 (2012). [Google Scholar]
- 33.Tschauner, O. et al. Discovery of davemaoite, CaSiO3-perovskite, as a mineral from the lower mantle. Science374, 891–894 (2021). [DOI] [PubMed] [Google Scholar]
- 34.Saikia, A. The Formation of Calcium Perovskite From Majorite Garnet: Implications for Splitting of the 520 km Seismic Discontinuity. Science319, 1515–1518 (2008). [DOI] [PubMed]
- 35.Hirose, K., Sinmyo, R. & Hernlund, J. Perovskite in Earth’s deep interior. Science738, 734–738 (2017). [DOI] [PubMed] [Google Scholar]
- 36.Wu, F., Sun, Y., Wan, T., Wu, S. & Wentzcovitch, R. M. Deep-Learning-Based Prediction of the Tetragonal → Cubic Transition in Davemaoite. Geophys. Res. Lett. 51, e2023GL108012 (2024).
- 37.Stixrude, L., Lithgow-Bertelloni, C., Kiefer, B. & Fumagalli, P. Phase stability and shear softening in CaSiO3 perovskite at high pressure. Phys. Rev. B - Condens. Matter Mater. Phys. 75, 024108 (2007).
- 38.Tsuchiya, T., Tsuchiya, J., Dekura, H. & Ritterbex, S. Ab Initio Study on the Lower Mantle Minerals. Annu. Rev. Earth Planet. Sci.48, 99–119 (2020). [Google Scholar]
- 39.Kudo, Y. et al. Sound velocity measurements of CaSiO3 perovskite to 133 GPa and implications for lowermost mantle seismic anomalies. Earth Planet. Sci. Lett.349–350, 1–7 (2012). [Google Scholar]
- 40.Irifune, T. & Ringwood, A. E. Phase transformations in subducted oceanic crust and buoyancy relationships at depths of 600-800 km in the mantle. Earth Planet. Sci. Lett.117, 101–110 (1993). [Google Scholar]
- 41.Ricolleau, A. et al. Phase relations and equation of state of a natural MORB: Implications for the density profile of subducted oceanic crust in the Earth’s lower mantle. J. Geophys. Res. Solid Earth115, B08202 (2010).
- 42.Leger, J. M. et al. Discovery of hardest known oxide. Nature383, 401 (1996). [Google Scholar]
- 43.Xu, F. et al. Silicon and oxygen self-diffusion in stishovite: Implications for stability of SiO2-rich seismic reflectors in the mid-mantle. Earth Planet. Sci. Lett.459, 332–339 (2017). [Google Scholar]
- 44.Zhang, Y., Fu, S., Wang, B. & Lin, J. F. Elasticity of a Pseudoproper Ferroelastic Transition from Stishovite to Post-Stishovite at High Pressure. Phys. Rev. Lett.126, 25701 (2021). [DOI] [PubMed] [Google Scholar]
- 45.Murakami, M., Hirose, K., Ono, S. & Ohishi, Y. Stability of CaCl2-type and α-PbO2-type SiO2 at high pressure and temperature determined by in-situ X-ray measurements. Geophys. Res. Lett.30, 1999–2002 (2003). [Google Scholar]
- 46.Andrault, D., Angel, R. J., Mosenfelder, J. L. & Le Bihan, T. Equation of state of stishovite to lower mantle pressures. Am. Mineral.88, 301–307 (2003). [Google Scholar]
- 47.Shieh, S. R., Duffy, T. S. & Shen, G. X-ray diffraction study of phase stability in SiO2 at deep mantle conditions. Earth Planet. Sci. Lett.235, 273–282 (2005). [Google Scholar]
- 48.Grocholski et al. Stability, metastability, and elastic properties of a dense silica polymorph, seifertite. J. Geophys. Res. E Planets118, 4745–4757 (2013). [Google Scholar]
- 49.Sun, N., Shi, W., Mao, Z., Zhou, C. & Prakapenka, V. B. High Pressure-Temperature Study on the Thermal Equations of State of Seifertite and CaCl2-Type SiO2. J. Geophys. Res. Solid Earth124, 12620–12630 (2019). [Google Scholar]
- 50.Yang, R. & Wu, Z. Elastic properties of stishovite and the CaCl2-type silica at the mantle temperature and pressure: An ab initio investigation. Earth Planet. Sci. Lett.404, 14–21 (2014). [Google Scholar]
- 51.Dubrovinsky et al. Pressure-induced transformations of cristobalite. Chem. Phys. Lett.333, 264–270 (2001). [Google Scholar]
- 52.Dubrovinskaia et al. Direct transition from cristobalite to post-stishovite α-PbO2-like silica phase. Eur. J. Mineral.13, 479–483 (2001). [Google Scholar]
- 53.Prokopenko, V. B., Dubrovinsky, L. S., Dmitriev, V. & Weber, H. P. In situ characterization of phase transitions in cristobalite under high pressure by Raman spectroscopy and X-ray diffraction. J. Alloy. Compd.327, 87–95 (2001). [Google Scholar]
- 54.Cernok, A. et al. Compressional pathways of α-cristobalite, structure of cristobalite X-I, and towards the understanding of seifertite formation. Nat. Commun.8, 15647 (2017). [DOI] [PMC free article] [PubMed]
- 55.Lin, Y. et al. Hydrous SiO2 in subducted oceanic crust and H2O transport to the core-mantle boundary. Earth Planet. Sci. Lett.594, 117708 (2022).
- 56.Katsura, T. A Revised Adiabatic Temperature Profile for the Mantle. J. Geophys. Res. Solid Earth127, 1–11 (2022). [Google Scholar]
- 57.Tsuchiya, T. Elasticity of subducted basaltic crust at the lower mantle pressures: Insights on the nature of deep mantle heterogeneity. Phys. Earth Planet. Inter.188, 142–149 (2011). [Google Scholar]
- 58.Dziewonski, A. M. & Anderson, D. L. Preliminary reference Earth model. Phys. Earth Planet. Inter.25, 297–356 (1981). [Google Scholar]
- 59.Mattern, E., Matas, J., Ricard, Y. & Bass, J. Lower mantle composition and temperature from mineral physics and thermodynamic modelling. Geophys. J. Int.160, 973–990 (2005). [Google Scholar]
- 60.Murakami, M., Ohishi, Y., Hirao, N. & Hirose, K. A perovskitic lower mantle inferred from high-pressure, high-temperature sound velocity data. Nature485, 90–94 (2012). [DOI] [PubMed] [Google Scholar]
- 61.Perrillat, J. P. et al. Phase transformations of subducted basaltic crust in the upmost lower mantle. Phys. Earth Planet. Inter.157, 139–149 (2006). [Google Scholar]
- 62.Ishii, T. et al. High pressure-temperature phase relations of basaltic crust up to mid-mantle conditions. Earth Planet. Sci. Lett.584, 117472 (2022). [Google Scholar]
- 63.Manthilake, G. M., De Koker, N., Frost, D. J. & McCammon, C. A. Lattice thermal conductivity of lower mantle minerals and heat flux from Earth’s core. Proc. Natl. Acad. Sci. Usa.108, 17901–17904 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Pradhan, G. K. et al. Melting of MORB at core-mantle boundary. Earth Planet. Sci. Lett.431, 247–255 (2015). [Google Scholar]
- 65.Hansen, S. E., Garnero, E. J., Li, M., Shim, S. H. & Rost, S. Globally distributed subducted materials along the Earth’s core-mantle boundary: Implications for ultralow velocity zones. Sci. Adv.9, 1–12 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Deschamps, F. & Cobden, L. Estimating core-mantle boundary temperature from seismic shear velocity and attenuation. Front. Earth Sci.10, 1–15 (2022).35300381 [Google Scholar]
- 67.Andrault, D. et al. Melting of subducted basalt at the core-mantle boundary. Science344, 892–895 (2014). [DOI] [PubMed] [Google Scholar]
- 68.Murakami, M. et al. Sound velocity of MgSiO3 post-perovskite phase: A constraint on the D″ discontinuity. Earth Planet. Sci. Lett.259, 18–23 (2007). [Google Scholar]
- 69.Oganov, A. R., Martonák, R., Laio, A., Raiteri, P. & Parrinello, M. Anisotropy of earth’s D″ layer and stacking faults in the MgSiO3 post-perovskite phase. Nature438, 1142–1144 (2005). [DOI] [PubMed] [Google Scholar]
- 70.Walte, N. P. et al. Transformation textures in post-perovskite: Understanding mantle flow in the D” layer of the Earth. Geophys. Res. Lett.36, 3–7 (2009). [Google Scholar]
- 71.Lee, C. T. et al. Upside-down differentiation and generation of a primordial lower mantle. Nature463, 930–933 (2010). [DOI] [PubMed] [Google Scholar]
- 72.Deng, X. et al. Compositional and thermal state of the lower mantle from joint 3D inversion with seismic tomography and mineral elasticity. PNAS120, e2220178120 (2023). [DOI] [PMC free article] [PubMed]
- 73.Li, Y., Frédéric, D. & Paul, J. T. The stability and structure of primordial reservoirs in the lower mantle: Insights from models of thermochemical convection in three-dimensional spherical geometry. Geophys. J. Int.199, 914–930 (2014). [Google Scholar]
- 74.Ballmer, M. D., Houser, C., Hernlund, J. W., Wentzcovitch, R. M. & Hirose, K. Persistence of strong silica-enriched domains in the Earth’s lower mantle. Nat. Geosci.10, 236–240 (2017). [Google Scholar]
- 75.Yuan, Q. et al. Moon-forming impactor as a source of Earth’s basal mantle anomalies. Nature623, 95–99, In revision (2023). [DOI] [PubMed]
- 76.Wan, B. et al. Seismological evidence for the earliest global subduction network at 2 Ga ago. Sci. Adv.6, 1–10 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Helffrich & Wood, B. J. The Earth’s mantle. Nature412, 501–507 (2001). [DOI] [PubMed]
- 78.Yang, A. Y. et al. A subduction influence on ocean ridge basalts outside the Pacific subduction shield. Nat. Commun.12, 1–10 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Ford, S. R., Garnero, E. J. & McNamara, A. K. A strong lateral shear velocity gradient and anisotropy heterogeneity in the lowermost mantle beneath the southern Pacific. J. Geophys. Res. Solid Earth111, 1–14 (2006). [Google Scholar]
- 80.Lakshtanov et al. The post-stishovite phase transition in hydrous alumina-bearing SiO2 in the lower mantle of the earth. Proc. Natl. Acad. Sci. Usa.104, 13588–13590 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Zhang et al. Elasticity of Hydrated Al-Bearing Stishovite and Post-Stishovite: Implications for Understanding Regional Seismic VS Anomalies Along Subducting Slabs in the Lower Mantle. J. Geophys. Res. Solid Earth127, 1–16 (2022). [Google Scholar]
- 82.Criniti, G., Ishii, T., Kurnosov, A., Glazyrin, K. & Boffa Ballaran, T. High-pressure phase transition and equation of state of hydrous Al-bearing silica. Am. Mineral.108, 1558–1568 (2023). [Google Scholar]
- 83.Bolfan-Casanova, N., Andrault, D., Amiguet, E. & Guignot, N. Equation of state and post-stishovite transformation of Al-bearing silica up to 100 GPa and 3000. K. Phys. Earth Planet. Inter.174, 70–77 (2009). [Google Scholar]
- 84.Andrault, D. et al. Phase diagram and P-V-T equation of state of Al-bearing seifertite at lowermost mantle conditions. Am. Mineral.99, 2035–2042 (2014). [Google Scholar]
- 85.Lakshtanov, D. L. et al. Effect of Al3+ and H+ on the elastic properties of stishovite. Am. Mineral.92, 1026–1030 (2007). [Google Scholar]
- 86.Ono, S. et al. Equation of state of Al-bearing stishovite to 40 GPa at 300 K. Am. Mineral.87, 1486–1489 (2002). [Google Scholar]
- 87.Panero, W. R., Benedetti, L. R. & Jeanloz, R. Transport of water into the lower mantle: Role of stishovite. J. Geophys. Res. Solid Earth108, 2039 (2003).
- 88.Jiang, F., Gwanmesia, G. D., Dyuzheva, T. I. & Duffy, T. S. Elasticity of stishovite and acoustic mode softening under high pressure by Brillouin scattering. Phys. Earth Planet. Inter.172, 235–240 (2009). [Google Scholar]
- 89.https://jp-minerals.org/vesta/en/.
- 90.He, Y. & Wen, L. Structural features and shear-velocity structure of the ‘Pacific Anomaly’. J. Geophys. Res. Solid Earth114, 1–17 (2009). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
All data supporting the findings of this study are available in the main text and Supplementary Information.
No custom code was developed for this study. Data analyses were carried out using OriginPro. 2023, Python 3.10 (including NumPy, SciPy, and Matplotlib), and the open-source BurnMan. package (version 1.1.0). Dioptas and PDIndexer were used for X-ray diffraction data integration, and indexing. All input parameters necessary to reproduce the results are provided in the main text and Supplementary Information.