Summary

Determining detailed descriptions of the space–time distribution of an earthquake's rupture using only teleseismic data is often a difficult and non unique process. However, a description of an earthquake that goes beyond a point source is often needed for many seismotectonic problems. We have developed a method that inverts measurements of the shifts in traveltime and amplitude of various seismic phases at global stations for the second degree polynomial moments of an event's space–time distribution. Our values for the second moments of the 1995 Mw 8.0 Jalisco, Mexico, thrust event correspond to a characteristic rupture length of 121±10 km along strike, a characteristic rupture width of 76±11 km downdip, a characteristic duration of 29±2 s, and a direction of rupture propagation along the strike of the subduction zone to the NW. For the 1995 Mw 7.2 Gulf of Aqaba earthquake, our estimates of the second moments correspond to a characteristic rupture length of 53±2 km, a characteristic duration of 12±1 s, and rupture propagation to the NE along the strike of the Dead Sea Transform. Our estimates of the first and second moments agree well with values from local seismic and geodetic networks. For both the Jalisco and Aqaba events, we are able to resolve the fault plane ambiguity associated with the events' moment tensors using our estimates of the second spatial moment.

Introduction

The spatial and temporal distributions of earthquake moment release have a large effect on the displacement field observed at locations near the fault. This strong dependence has enabled seismologists to image the space–time evolution of the rupture of earthquakes recorded by nearby instruments in remarkable detail (Wald & Heaton 1994; Ide et al. 1996). However, abundant near field data have generally been lacking except in regions of dense regional seismic networks, such as California and Japan. This has motivated a large amount of work on imaging earthquake ruptures from teleseismic data. Such far away observations have limited resolution; however, Ihmlé (1998) recently showed that P wave inversions for the Mw 8.2 Bolivia deep earthquake, which had good teleseismic coverage and relatively simple P wave propagation, were unable to differentiate between a slip distribution composed of compact subevents and one composed of widely distributed slip. This inherent non uniqueness is often associated with a choice of the spatial smoothing parameter that is used to regularize these inversions (Ihmlé 1998). Imaging the slip distribution of large earthquakes is further complicated because P waves provide only weak constraints on the total moment of an event (Ekström 1989). Thus, attempts to determine the details of slip distributions for earthquakes that are only recorded teleseismically are at best extremely non unique. There are several important properties of an earthquake's space–time behaviour that do not require this type of detailed finite fault inversion, such as resolution of the fault plane ambiguity, approximate extent of the rupture area, duration and direction of rupture propagation. In this paper, we develop a method for robustly estimating these large scale features of the slip distribution from teleseismic data.

General description of seismic sources

In the normal mode formulation, the spatial and temporal extent of an earthquake's rupture enters the equation for the displacement u at a position x through the stress glut rate tensor Γ˙(r, t) (Gilbert 1970; Dahlen & Tromp 1998):

1

where

2

and sk is the eigenvector of the kth mode with eigenfrequency ωk, attenuation parameter αk and associated strain tensor Ek. Γ˙(r, t), introduced and described in detail by Backus & Mulcahy (1976a,b), is non zero only in the region of non elastic deformation, i.e. the source volume. In this paper, we use a dot to represent the time derivative and assume that Γ˙(r, t) has the form

3

where is a unit norm moment tensor and (r, t) is the scalar source space–time function. For slip on a simple fault, this assumption is equivalent to requiring a constant orientation of the fault plane and slip vector, but it also allows slip on parallel faults. In contrast to near field inversions that seek to determine a detailed parametrization of (r, t), we will invert for the polynomial moments of (r, t), μ(i,j), which are tensors of order i and have spatial degree i, temporal degree j and total degree i+j. The zeroth degree moment

4

is the standard seismic moment, M0, while the normalized first degree moments μ(1,0)/M0 and μ(0,1)/M0 give the spatial and temporal centroids, respectively:

5

The degree 2 moments about a point r0 and time t0 are called central moments when r0=μ(1,0)/M0 and t0=μ(0,1)/M0:

6

graphic

7

The coordinate system axes (r, θ, φ) correspond to radius, colatitude and longitude respectively. If the degree 2 moments are calculated about a point that is slightly offset from the true centroid location, there is a small difference between the second moment and the second central moment:

8

The second central moments describe the spatial and temporal extent of the source. Following Backus (1977a) and Silver & Jordan (1983), we can define the characteristic dimension, xc(), of the source region in a direction , the characteristic duration τc, the characteristic (or apparent) rupture velocity vc and the average velocity of the instantaneous spatial centroid, v0:

graphic

9

where Lc is the maximum value of xc, that is, the characteristic rupture length along the direction of the eigenvector associated with the largest eigenvalue of μ̂(2,0).

A number of theoretical papers on second moments have examined the non uniqueness in source inversions and the relationships between second moments and standard source models (Backus 1977b; Okal 1982; Doornbos 1982a,b; Pavlov 1994; Das & Kostrov 1997), but prior observational work on determining the polynomial moments of earthquakes has been limited to the inversion of body waveform data (Gusev & Pavlov 1988) or surface wave amplitude spectra assuming a fault plane a priori (Bukchin 1995). In this paper, we develop a formalism for inverting frequency dependent traveltime and amplitude measurements for the source moments of polynomial degrees 0, 1 and 2. Data are obtained for arbitrary wave groups, such as surface waves and body waves, as differential measurements relative to synthetics computed for a point source in a 3 D earth model.

2010 Frequency-dependent phase and amplitude data

We measure the phase and amplitude of a target wave group relative to a synthetic seismogram using the GSDF technique of Gee & Jordan (1992). The synthetic seismogram, , is represented as a sum over travelling wave groups, indexed by mode number n. In the frequency domain, it is assumed that the data seismogram, s, can be related to the synthetic by small differences in phase delay, δτpn, and amplitude reduction time, δτqn:

10

GSDF measurements of the generalized delay time, δτpn(ω), have been utilized in a variety of structural inverse problems (Gaherty et al. 1996; Katzman et al. 1998) because the relationships between δτpn(ω) and structural model parameters are easily linearized. In those studies, synthetic seismograms were calculated for short duration earthquakes (Mw∼6) using a 1 D earth model, and the phase delays of the data were attributed to structural variations along the path between the source and station. Here, we study large earthquakes (Mw≥7.0) that cannot be represented adequately as point sources, even at low frequencies. We make approximate corrections for Earth structure using a 3 D earth model, so that our measurements of δτpn and δτqn can be attributed primarily to properties of the seismic source.

The synthetic seismograms in this study were calculated for a point source in both space and time through normal mode summation for the PREM earth model (Dziewonski & Anderson 1981), corrected for 3 D elastic structure using the degree 12 aspherical model of Su et al. (1994) and the asymptotic approximations of Woodhouse & Dziewonski (1984). We also corrected fundamental modes above 7 mHz for smaller scale heterogeneity using the degree 40 phase velocity maps of Ekström et al. (1997). The 3 D corrections only offset the phase of the surface wave arrivals; no corrections are made to the amplitudes for the effects of 3 D anelastic structure or focusing and defocusing. The source is specified by a centroid location, centroid time and moment tensor from the Harvard CMT catalogue (Dziewonski et al. 1997). To limit any contributions to our measurements by lateral heterogeneity unaccounted for by the 3 D models, we restricted our measurements of surface waves to low frequencies (≤20 mHz), and we omitted any stations for which our synthetic seismograms did not closely resemble the observed seismograms.

The first step in the measurement process was the creation of an isolation filter by a weighted summation of normal modes. The isolation filter was designed to approximate the complete synthetic in the time interval containing a target wave group and decay to zero outside that interval (Fig. 1). The isolation filter was then cross correlated with both the complete synthetic and the data seismograms. The resulting cross correlagrams were windowed to localize the signals in the time domain and then put through a series of narrow band filters (with centre frequencies ωi) to localize them in the frequency domain. The windowed and filtered cross correlagrams were then fitted with two five parameter Gaussian wavelets from which the phase and amplitude parameters, δτp(ωi) and δτq(ωi), were calculated (Gee & Jordan 1992). This procedure corrects for the interference between the phase targeted by the isolation filter and other arrivals on the seismogram, as well as the biases introduced by the windowing and filtering operations. For shallow earthquakes, we typically made measurements on vertical component seismograms for fundamental mode Rayleigh waves at frequencies between 3 and 20 mHz and for P waves between 10 and 50 mHz.

(a) Examples of isolation filters. The R1 isolation filter in the top panel was made by summing only the modes shown in black in the PREM dispersion diagram in the bottom panel. The full synthetic was calculated by summing all the spheroidal modes of the PREM earth model (i.e. the entire bottom panel). All three traces in the top panel have been narrow band filtered around 10 mHz. (b) Examples of isolation filters. The P wave isolation filter in the top panel was made by summing the modes shown in black in the bottom panel. Modes shown in grey were not included in the summation for the isolation filter, but were included in the full synthetic. All three traces were narrow band filtered around 35 mHz.
Figure 1

(a) Examples of isolation filters. The R1 isolation filter in the top panel was made by summing only the modes shown in black in the PREM dispersion diagram in the bottom panel. The full synthetic was calculated by summing all the spheroidal modes of the PREM earth model (i.e. the entire bottom panel). All three traces in the top panel have been narrow band filtered around 10 mHz. (b) Examples of isolation filters. The P wave isolation filter in the top panel was made by summing the modes shown in black in the bottom panel. Modes shown in grey were not included in the summation for the isolation filter, but were included in the full synthetic. All three traces were narrow band filtered around 35 mHz.

Effects of an extended source on GSDF measurements

To relate the differences between the point source synthetic and data seismograms to our source parametrization, we determined the partial derivatives of δτpn and δτqn with respect to the polynomial moments of (r, t). A Taylor series expansion truncated to second order gives the expression relating a point source seismogram, , to the seismogram resulting from an extended source, s (e.g. Dahlen & Tromp 1998):

11

where rR and rS are the receiver and point source locations respectively, tS is the point source time, and μ(1,0) and μ(0,1) represent any deviation of the centroid (r0, t0) from the point source location. The operator S acts on the source spatial coordinates (rS, θS, φS) only. In the GSDF formalism, and s are represented as sums over travelling modes:

12
13

where we have dropped the argument rR for convenience. To obtain the explicit expression for , we converted the expression for the synthetic seismogram from a sum over normal modes to a sum over travelling modes using the Poisson sum formula (Appendix A):

14

where Δ is the epicentral distance and An(ω) and Bn(ω) depend on the eigenfunctions of the earth model, the assumed centroid moment tensor and the epicentral distance. The perturbations to the continuous wavenumber, δλn, account for the corrections made to the synthetics for 3 D Earth structure (Appendix A). Eqs (12) and (14) yield expressions for τpn(ω) and τqn(ω):

15
16

Substituting eqs (12) and (13) into eq. (11) and making the low frequency approximation,

17

gives the following formulae for δτpn(ω) and δτqn(ω):

18
19

Expressions for the terms involving τpn and τqn are calculated from eqs (15) and (16). The partial derivatives of δτpn and δτqn with respect to the 14 independent elements of μ(0,1), μ(1,0), μ(0,2), μ(1,1) and μ(2,0) can be simply read off eqs (18) and (19). We have also calculated the partial derivatives of the GSDF measurements with respect to changes in the moment tensor elements, Mij, using the expressions

20
21

In practice, isolation filters for most phases involve contributions from a large number of normal modes ranging over multiple values of the overtone number n. Thus the measured values of δτp and δτq will be a weighted sum over the δτpn and δτqn for the individual branches that contribute to the isolation filter. The partial derivative of a GSDF measurement can also be calculated by a weighted sum of the partial derivatives of the individual branches (Gee & Jordan 1992, eqs 94–95). Fig. 1 shows examples of the isolation filters for the P wave and the Rayleigh wave from a shallow source along with the normal mode weights used to calculate the filters. Only the fundamental mode branch is needed to match the Rayleigh wave, while contributions from several branches are needed to match the P wave. While all of the GSDF measurements in this paper were made on vertical component records of either fundamental mode Rayleigh waves or P waves, the technique is completely generalizable to an arbitrary waveform. Thus, intermediate and deep earthquakes can be studied by the inclusion of higher mode surface waves. Additionally, second orbit surface waves can be used to increase azimuthal coverage of the source (Appendix A). In the future, we will incorporate Love and other SH waves to improve our resolution of the higher moments.

Examples of partial derivatives

Several intuitive properties can be seen in the partial derivatives (Fig. 2) for the Rayleigh wave isolation filter shown in Fig. 1. The effect of a shift in centroid time, μ(0,1), on the frequency dependent arrival times δτp(ω) is independent of azimuth, and this type of shift has no effect on the frequency dependent amplitude parameter δτq(ω). ∂δτp/∂μ(1,0) shows a cosine dependence on azimuth with a maximum at 270°. This pattern corresponds to late arrivals at stations to the west and early arrivals at stations to the east owing to an eastward shift in the centroid location. A similar pattern, shifted by 90°, is found for ∂δτp/∂μ(1,0). These traveltime shifts due to a mislocated rS are independent of frequency. The centroid depth, μr(1,0), is the only component of the first degree moment that significantly perturbs the amplitude reduction time, δτq. In contrast, the degree 2 moments primarily perturb the δτq measurements, having almost no influence on δτp. ∂δτq/∂μ̂(1,1) approximates a cosine in azimuth and increases in amplitude with frequency. This behaviour is just the directivity pattern of Ben Menahem (1961), which has been used in many studies to infer directivity from surface waves. The φφ component of μ̂(2,0) is non zero for sources with finite extent in the east–west direction. ∂δτq/∂μ̂(2,0) has positive values (i.e. decreased data amplitudes) at stations to both the east and west due to destructive interference (Fig. 3). ∂δτq/∂μ̂(0,2) is independent of azimuth, but increases in amplitude with frequency, corresponding to the spectral roll off expected for a source of finite duration.

Examples of the partial derivatives of the GSDF measurements for the fundamental mode Rayleigh wave. The partials are for δτp(ω) (a, c, e, g) and δτq(ω) (b, d, f, h) at frequencies of 3.5 mHz (a, b, e, f) and 5.0 mHz (c, d, g, h) mHz. (a)–(d) Partials with respect to the first moments, centroid time (black circles), longitude (grey triangles), colatitude (black triangles) and depth (grey circles). Y axis tick marks are every 0.0026 s (1020 Nm s)−1, 0.001 s (1020Nm km)−1 , 0.0008 s (1020Nm km)−1 and 0.0008 s (1020Nm km)−1 for μ(0,1), μr(1,0), μ(1,0) and μ(1,0) respectively. (e)–(h) Partials with respect to the second moments μ●(2,0) (black triangles), μ●rr(2,0) (grey circles), μ●0,2 (black circles) and μ●(1,1) (grey triangles). Y axis tick marks are every 0.000053 s (1020Nm s2)−1, 0.000011 s (1020Nm km s)−1 and 0.000011 s (1020Nm km2)−1 for μ0,2, μ(1,1) and μ(2,0) respectively.
Figure 2

Examples of the partial derivatives of the GSDF measurements for the fundamental mode Rayleigh wave. The partials are for δτp(ω) (a, c, e, g) and δτq(ω) (b, d, f, h) at frequencies of 3.5 mHz (a, b, e, f) and 5.0 mHz (c, d, g, h) mHz. (a)–(d) Partials with respect to the first moments, centroid time (black circles), longitude (grey triangles), colatitude (black triangles) and depth (grey circles). Y axis tick marks are every 0.0026 s (1020 Nm s)−1, 0.001 s (1020Nm km)−1 , 0.0008 s (1020Nm km)−1 and 0.0008 s (1020Nm km)−1 for μ(0,1), μr(1,0), μ(1,0) and μ(1,0) respectively. (e)–(h) Partials with respect to the second moments μ●(2,0) (black triangles), μ●rr(2,0) (grey circles), μ●0,2 (black circles) and μ●(1,1) (grey triangles). Y axis tick marks are every 0.000053 s (1020Nm s2)−1, 0.000011 s (1020Nm km s)−1 and 0.000011 s (1020Nm km2)−1 for μ0,2, μ(1,1) and μ(2,0) respectively.

(a) GSDF measurements of the fundamental mode Rayleigh wave from ‘synthetic data’ constructed for a unilateral rupture model. (b) GSDF measurements of the fundamental mode Rayleigh wave from ‘synthetic data’ constructed for a bilateral rupture model.
Figure 3

(a) GSDF measurements of the fundamental mode Rayleigh wave from ‘synthetic data’ constructed for a unilateral rupture model. (b) GSDF measurements of the fundamental mode Rayleigh wave from ‘synthetic data’ constructed for a bilateral rupture model.

Fig. 3 shows the GSDF measurements that result from two simple models of an extended earthquake source, unilateral and bilateral rupture. In these calculations, the centroid and moment tensor are known exactly, and (r, t) is symmetric about the centroid, so the dominant contribution comes from the second moments. In both cases, almost no deviation from the point source synthetic is observed at low frequencies, but a characteristic pattern emerges in the δτq measurements as frequency increases. For unilateral, westward rupture (Fig. 3a), the effects of μ̂(1,1) and μ̂(2,0) approximately cancel in the direction of rupture propagation, while they add constructively in the opposite direction. Thus a ‘half cosine’ pattern emerges in the δτq(ω) measurements, with the smallest data amplitudes occurring 180° away from the direction of rupture propagation. We observe this azimuthal pattern, which is diagnostic of a predominately unilateral rupture, for both the 1995 Jalisco and Gulf of Aqaba earthquakes, which are discussed in the Results section. In the bilateral case (Fig. 3b), μ̂(1,1) is zero, so the μ̂(2,0) term dominates, producing small amplitudes in both directions along the strike of the fault. For both of these examples, the finite duration of the source shifts the δτq measurements to positive values with increasing frequency, owing to spectral roll off.

Directivity effects are often assumed to average to zero in source time function and source spectrum estimation techniques if sufficient azimuthal coverage is available. The amplitude perturbations owing to finite source effects shown in Fig. 3 have only positive values of δτq, indicating a reduction in the data amplitude relative to the point source synthetic. While the perturbations due to the propagation of the centroid, ∂δτq/∂μ(1,1) (Figs 2f and g), average to zero as a function of azimuth, the perturbations owing to μ̂(2,0) do not (Figs 2f and g). The amplitude increases in the direction of rupture propagation resulting from μ̂(1,1) are cancelled in the unilateral case by the decreases resulting from μ̂(2,0) (Fig. 3a). In the perfectly bilateral case (Fig. 3b), μ̂(1,1)=0, so only the amplitude reductions resulting from μ̂(2,0) occur. Thus techniques that assume that amplitude effects resulting from directivity average to zero will underestimate the earthquake's seismic moment.

Inversion method

Optimization technique

We can formulate a linear inverse problem to minimize ||Amb|| for a vector m, which contains the 14 independent elements of μ(0,1), μ(1,0), μ(0,2), μ(1,1) and μ(2,0) as well as the six independent elements of M. The data vector, b, contains the GSDF measurements and A contains the partial derivatives calculated using eqs (18)–(21). For sources that obey the assumption in eq. (3), the moments of (r, t) are uniquely determined by the displacement field (Bukchin 1995), making the linear inverse problem theoretically solvable. In practice, however, owing to errors in the assumed moment tensor, the effects of unmodelled lateral heterogeneity and the small relative size of the minimum eigenvalue of μ(2,0), a linear least squares inversion does not result in a robust solution. Das & Kostrov (1997) incorporated linear inequality constraints among the moments to stabilize their inversion; however, their constraints, which are based on the Hausdorff inequalities, required a priori information about the maximum size of the source region. The power of these constraints depended strongly on the a priori limits on the source region size (Das & Kostrov 1997). To avoid setting a priori limits on the extent of the source region, we stabilized our inversion by incorporating the physical constraint that the 4 D source region have non negative volume. This constraint, which is an inherent property of all earthquakes, is equivalent to requiring the 4×4 matrix

22

to be positive semi definite. This constraint is non linear in the elements of m and encompasses the constraints placed on the second degree moments of (r, t) by other authors through the use of Bessel inequalities (Bukchin 1995, see also Appendix B). Methods for solving linear problems subject to matrix inequality constraints such as (22) have recently been developed in the optimization literature. We use the semi definite programming approach of Vandenberghe & Boyd (1996). To put our problem into this form, it is necessary to restate the non linear (least squares) objective function as a linear objective function subject to a non linear (quadratic) constraint. We introduce a dummy variable, c, such that ||Amb||≤c. Then our problem can be restated as

graphic

23

graphic

where IN is the identity matrix with dimension equal to the number of GSDF measurements, N, and ≥0 indicates that the matrix is positive semi definite. The equivalence between the linear least squares problem and (22) can be seen by calculating the eigenvalues of the N+1×N+1 matrix in (23), which are non negative when the matrix is positive semi definite. This restatement of the problem is known as using Schur complements to represent a non linear constraint as a linear matrix inequality (Vandenberghe & Boyd 1996). The semi definite programming algorithm also allows us to incorporate a variety of linear inequality constraints, such as requiring the source to lie on one of the fault planes of the moment tensor or constraining the source to occur below the free surface, when appropriate for a particular event. The physical constraint (22) greatly reduces the non uniqueness in the minimization problem without requiring any a priori information about the extent of the source.

Any difference between the assumed point source location (rS, tS) and the true centroid location(r0, t0) will result in a slight difference between the estimated value of the second moments and the values of the second central moments (see eq. 8). Correcting the estimates of the second moments for a change in centroid location can violate the positive definite constraint (22). Thus we prefer to iterate the inversion (by updating the CMT parameters) until the correction term is negligible. Usually only one additional iteration is necessary for the elements of m corresponding to the changes in the CMT parameters to be negligible.

For events of moderate size, Mw≈7, we usually need to include high frequency (20–50 mHz) P wave measurements to resolve the second moments accurately. These inversions also require the inclusion of low frequency Rayleigh wave measurements to constrain the moment tensor. We have noticed that combining these two types of measurements often leads to smaller than expected values of μ(0,2). This bias results from a small discrepancy (≤20 per cent) in the relative amplitudes of P and R1 waves in our synthetics as compared to the data. Iterating the centroid depth often removes most of this anomaly, but there appears to be a residual bias even after this process. To remove this bias from our inversions, we include a denuisance parameter that allows for a small, frequency independent excitation anomaly in the P wave measurements. The estimated value of this parameter usually corresponds to δτq values of 1–2 s, and hence is a smaller signal than the observed directivity.

Error analysis

We have also incorporated the non negative volume constraint (22) into our estimates of the errors in our solution. This is important because the optimal solution to a semi definite programming problem will often lie on the boundary of the feasible region defined by the matrix inequality constraints (Vandenberghe & Boyd 1996). Standard linear estimates do not account for the existence of an infeasible zone, but one method that can is the grouped jackknife (Tukey 1984). The jackknife technique relies on the assumption that different elements of a data set are independent. In our data sets of GSDF measurements, errors in the measurements at a given station are likely to be correlated from one frequency to the next. For instance, all the arrivals at one station may be early owing to fast velocity anomalies along the path. Thus, to provide a more conservative (larger) estimate of the variance of our model, we created M subsets of data by deleting all the measurements in each of M azimuthal bins (usually 12 bands of 30°) to create the ith subset. The jackknife method then gives an estimate of the covariance matrix, Cmm, of our model parameters:

24

where i is the estimate produced by the semi definite program technique using the ith subset of the data, and (.) is the average of the M values of i. The diagonal elements of Cmm give the variances of the individual model parameters.

The jackknife estimate of Cmm can also be used to determine the variances of derived quantities such as the characteristic rupture dimension, xc, in a direction . The equation for our estimate of xc, eq. (9), can be rewritten in vector form as

25

where z is a function of , and contains the elements of our estimate of μ̂(2,0). Using standard error propagation equations (Bevington & Robinson 1992, p. 41), we can also estimate the variance in our estimate of xc:

26

Results

1995 Kobe earthquake

The 1995 Mw 6.9 Kobe earthquake was recorded extremely well by the Japanese strong motion network, allowing for a good comparison of our teleseismic results with more accurate estimates determined from the local data. In particular, we use it to check for biases in the zeroth and first degree moments of (r, t). For this event we initially calculated synthetics using the Harvard CMT (Dziewonski et al. 1996) (20:46:59.4, 34.78°N, 134.99°E, 20 km depth). We then made measurements, shown in Fig. 4, of the fundamental mode first orbit Rayleigh wave at 46 global stations for which our synthetics closely resembled the observed seismograms. The δτp(ω) measurements showed a pattern of late arrivals on the data seismograms at stations to the north and early data arrivals at stations to the south, while the δτq measurements were essentially a flat function of azimuth, indicating that the data amplitudes were slightly larger than the synthetics. The lack of a clear directivity signal in the amplitude measurements is because the measurements were made at wavelengths for which an event of this size is well represented by a point source.

GSDF measurements (open squares) for the Harvard CMT as a function of station azimuth for the fundamental mode, first orbit Rayleigh wave for the 1995 Kobe event at global stations. Black circles denote the fit to the measurements from our inversion procedure. A value of zero indicates that no perturbation of the data relative to the synthetic was observed.
Figure 4

GSDF measurements (open squares) for the Harvard CMT as a function of station azimuth for the fundamental mode, first orbit Rayleigh wave for the 1995 Kobe event at global stations. Black circles denote the fit to the measurements from our inversion procedure. A value of zero indicates that no perturbation of the data relative to the synthetic was observed.

Owing to the lack of a clear directivity signal in the amplitude measurements, we only inverted for changes to the first moments and moment tensor. The results of the inversion are given in Table 1 and Fig. 5, and the fit to the data is displayed in Fig. 4. The centroid location moved 24 km to the south, which fits the δτp pattern, and 5 km shallower. The changes in the moment tensor slightly increased in the seismic moment while retaining approximately the same focal mechanism. However, the changes to the M and M components are not well constrained. The new centroid location is about 10 km from the value determined by (Ide & Takeo 1997) (34.61°N, 135.075°E, 9.2 km) using local data and provides a better estimate than the starting Harvard CMT values (Fig. 5). We have calculated synthetic seismograms for the revised CMT and remeasured the data (Fig. 6). The results show essentially flat functions of azimuth at a value of zero for both δτp and δτq. Inversion of these remeasured data produces extremely small (≤∼1 km) shifts in the centroid location and correspondingly small shifts in the moment tensor, which are below our resolution limit. The residual scatter in Fig. 6, which has an average rms amplitude of 7 s depending on frequency band, is indicative of the amount of error in our measurements due to unmodelled propagation effects. This level of scatter is significantly smaller than the signal produced by directivity in large earthquakes, as will be shown in the next section.

First moments inversion results.
Table 1

First moments inversion results.

Map of the 1995 Kobe earthquake. Small dots denote the locations of aftershocks from 1995 January 25 to 1995 February 9 as determined by the Japanese University Group for Urgent Joint Observation of the 1995 Hyogo ken Nanbu Earthquake. Centroid locations determined by the Harvard CMT, Ide et al. (1996) and this study are shown as the large square, circle and triangle respectively.
Figure 5

Map of the 1995 Kobe earthquake. Small dots denote the locations of aftershocks from 1995 January 25 to 1995 February 9 as determined by the Japanese University Group for Urgent Joint Observation of the 1995 Hyogo ken Nanbu Earthquake. Centroid locations determined by the Harvard CMT, Ide et al. (1996) and this study are shown as the large square, circle and triangle respectively.

GSDF measurements (open squares) for the Kobe event using the revised centroid given in Table 1, and the fit to them (black circles) after inversion. A value of zero indicates that no perturbation of the data relative to the synthetic was observed.
Figure 6

GSDF measurements (open squares) for the Kobe event using the revised centroid given in Table 1, and the fit to them (black circles) after inversion. A value of zero indicates that no perturbation of the data relative to the synthetic was observed.

1995 Jalisco earthquake

The 1995 Mw 8.0 Jalisco, Mexico, earthquake occurred on a shallow thrust plane from the subduction of the Rivera Plate under the North American Plate (Fig. 7). A good test of our method comes from comparing our estimate of μ̂(2,0) with the slip distribution determined from a local GPS study by Melbourne et al. (1997). We performed an initial inversion for changes in the centroid and moment tensor. The new centroid location (15:36:28.80, 19.43°N, 104.92°W, 8.0 km) corrected the early R1 arrivals at stations to the southeast found for the Harvard CMT (Dziewonski et al. 1997). For most shallow sources, the M and M components were not well constrained, so we settled on a moment tensor (Mrr=4.6, M=−3.2, M=−1.3, M=5.0, M=−3.0, M=1.9×1020 N m) through a process of forward modelling. Measurements for synthetics calculated from this CMT are shown in Fig. 8. The primary signals visible in the data are low R1 amplitudes at stations to the SE. The magnitudes of these signals increase with frequency, as expected for a finite source. The baseline level of the amplitude reduction times, δτq, also shifts upwards from about 8 s at 5 mHz to about 12 s at 10 mHz. This overall decrease in amplitude with frequency is indicative of the spectral roll off caused by the finite duration of the source time function. We inverted the data in Fig. 8 under the non negative volume constraint (22) and under the constraint that the characteristic length in the radial direction, xc(), be ≤10 km. This added constraint ensured that the source volume did not extend above the free surface.

Map of the region around the 1995 Jalisco, Mexico, earthquake. The focal mechanism shown is for the Harvard CMT.
Figure 7

Map of the region around the 1995 Jalisco, Mexico, earthquake. The focal mechanism shown is for the Harvard CMT.

GSDF measurements (open squares) for the Jalisco event using the revised centroid given in Table 1, and the fit to them (black circles) after inversion.
Figure 8

GSDF measurements (open squares) for the Jalisco event using the revised centroid given in Table 1, and the fit to them (black circles) after inversion.

Our inversion (Tables 1, 2 and 3) yielded a characteristic duration τc of 29±2 s, a centroid velocity v0 of 3.7±0.2 km s−1 at an azimuth of 307°, a characteristic rupture length Lc of 121±10 km along the strike of the subduction zone, a characteristic width wc of 76±11 km downdip, and a characteristic rupture velocity vc of 4.1±0.1 km s−1. The three eigenvalues of μ̂(2,0) specify the semi axis lengths of an ellipsoid whose axes are oriented in the direction of the associated eigenvectors. This characteristic source volume, shown for the Jalisco event in Fig. 9, is an estimate of the region that contributed significantly to the moment release. The source ellipsoid is nearly planar, with its longest dimension oriented NW–SE along the strike of the subduction zone. The ellipsoid dips slightly to the NE as expected for a shallow thrust plane. Fig. 10 compares our estimate of the characteristic source dimensions (solid ellipse) to the GPS slip inversion of Melbourne et al. (1997). Our centroid location agrees well with the value calculated from the slip distribution and the Harvard CMT. The NEIC epicentre is displaced about 80–100 km to the SE. The position of the centroid relative to the epicentre agrees with the direction of the characteristic velocity. We calculated the characteristic dimensions for Melbourne et al.'s distribution (dashed ellipse in Fig. 10), which was confined to a single plane (strike 302°, dip 16°). Its larger dimension is along strike, Lc=91 km, and its smaller dimension is oriented downdip, wc=50 km. Thus, our estimate of μ̂(2,0) is consistent with the local inversion.

Plots of the characteristic source volume of the Jalisco event defined by μ(2,0). Shading indicates deeper depths. (a) Cross sectional view. (b) Map view.
Figure 9

Plots of the characteristic source volume of the Jalisco event defined by μ(2,0). Shading indicates deeper depths. (a) Cross sectional view. (b) Map view.

Contours of slip (m) for the Jalisco event taken from Melbourne et al. (1997). Centroid locations are shown for the Harvard CMT (small circle), Melbourne et al. (black square) and this study (black triangle). Characteristic source ellipses are shown for this study (solid black) and for the Melbourne et al. slip distribution (dashed black). The large circle is the PDE epicentre and the arrow represents the direction of the characteristic velocity.
Figure 10

Contours of slip (m) for the Jalisco event taken from Melbourne et al. (1997). Centroid locations are shown for the Harvard CMT (small circle), Melbourne et al. (black square) and this study (black triangle). Characteristic source ellipses are shown for this study (solid black) and for the Melbourne et al. slip distribution (dashed black). The large circle is the PDE epicentre and the arrow represents the direction of the characteristic velocity.

Second moments inversion results.
Table 2

Second moments inversion results.

Second moments inversion results.
Table 3

Second moments inversion results.

To resolve the fault plane ambiguity for the Jalisco focal mechanism (Fig. 7), we calculated the characteristic rupture length in the direction normal to each of the candidate fault planes. For the shallowly dipping plane (strike 302°, dip 19°), xc()=21±2 km and for the steeply dipping plane xc() = 76±10 km. The shallowly dipping plane has a much smaller value of xc(), and we therefore infer that it is the true rupture plane. This agrees with Melbourne et al.'s study using local data. The small value of xc() is due to the dip of the characteristic source volume to the northeast. This dip is a robust feature of the inversion. Fig. 11 shows histograms for each element of the first and second moments for the 12 different inversions carried out for the jackknife calculation (corresponding to deleting each of the 12 30° azimuth bins from the full data set). In general, the distributions are peaked near the value for the full data set with a relatively small range of values. The positive values of μ(2,0) (most trials) and the negative values of μ(2,0) (11 of 12 trials) both correspond to a source volume that dips to the NE. The characteristic volume has a dip of about 7° to the NE, while the moment tensor (19°) and the fault plane determined from the local network (17°) have slightly steeper values. Thus, even for this shallow thrust event, it is possible to recover the approximate dip direction from the second moments.

Histograms of the values of the Jalisco event's first and second moments determined from the Jackknife error test. 12 inversions were run, with each inversion corresponding to a subset of data where all the measurements for one of the 12 azimuth bands had been deleted.
Figure 11

Histograms of the values of the Jalisco event's first and second moments determined from the Jackknife error test. 12 inversions were run, with each inversion corresponding to a subset of data where all the measurements for one of the 12 azimuth bands had been deleted.

1995 Gulf of Aqaba earthquake

The 1995 Mw7.2 Gulf of Aqaba earthquake ruptured the southern extension of the Dead Sea Transform (Fig. 12a). First orbit Rayleigh waves at low frequencies (≤10 mHz), measured using synthetics computed for the Harvard CMT, showed no systematic signal in either δτp or δτq. However, higher frequency measurements of teleseismic P waves did display systematically low amplitudes at stations to the south of the event (Fig. 13). We inverted measurements of δτp and δτq for R1 at 5 and 10 mHz and P waves at 25, 35 and 45 mHz for the first and second moments and the moment tensor elements. The characteristic dimension in the radial direction was constrained to be less than the Harvard CMT centroid depth of 18 km. The results of our inversion (Tables 1, 2 and 3) correspond to a characteristic duration, τc, of 12±1 s, a centroid velocity, v0, of 4.3±0.3 km s−1 in the direction N05°E and a characteristic rupture length, Lc, of 53±2 km in the direction N06°E. The centroid location (Fig. 12) is shifted slightly to the west of the gulf of Aqaba, presumably due to incorrectly modelled 3 D elastic structure.

(Top) Map of the Eastern Mediterranean region showing the location and mechanism of the 1995 Aqaba earthquake. (Bottom) Map view of the source region of the Aqaba earthquake. The triangle shows the centroid location given in Table 1. The arrow represents the characteristic velocity vector, scaled to a length equal to the product of τc and vc. The ellipse is the map view projection of the characteristic source volume corresponding to the values of μ●(2,0) in Table 1.
Figure 12

(Top) Map of the Eastern Mediterranean region showing the location and mechanism of the 1995 Aqaba earthquake. (Bottom) Map view of the source region of the Aqaba earthquake. The triangle shows the centroid location given in Table 1. The arrow represents the characteristic velocity vector, scaled to a length equal to the product of τc and vc. The ellipse is the map view projection of the characteristic source volume corresponding to the values of μ●(2,0) in Table 1.

GSDF measurements (open squares) of the P wave at global stations for the Aqaba event. Black circles show the fit to the measurements after inversion.
Figure 13

GSDF measurements (open squares) of the P wave at global stations for the Aqaba event. Black circles show the fit to the measurements after inversion.

The characteristic rupture volume corresponding to our values of μ̂(2,0) is shown in map view in Fig. 12(b). One fault plane of the moment tensor strikes NNE–SSW while the other strikes WNW–ESE. The NNE striking plane has a value of xc()=12±1 km while the ESE striking plane has a value of xc()=50±2 km. Thus, we prefer the NNE striking plane as the true rupture plane. This fault plane is consistent with the orientation of the Dead Sea Transform in this region. The small difference in the azimuth of the characteristic rupture length, 006°, from the strike of the fault plane of the moment tensor, 018°, may be related to this event rupturing more than one fault. Pinar (1997) and Klinger et al. (1999) showed that this event's rupture jumped north across a series of faults in the Dead Sea Transform system. This jump would cause a difference between the azimuth of the fault plane in the focal mechanism and the azimuth of the characteristic rupture volume of the sense we determine, but the difference may also be a result of the limited station coverage in southern azimuths.

Discussion

We have developed an inversion procedure that can estimate the first and second degree moments of earthquake source distributions without any a priori constraints on the orientation or extent of the source. The second moments are a desirable quantity to determine because they are independent of the types of parametrizations and smoothing operations typically employed in teleseismic finite fault inversions, and they can be used as integral constraints in determining the space–time history of rupture. The technique presented in this paper can be applied to any seismic phase for which an isolation filter can be constructed. Future studies will take advantage of the different samplings of the source provided by using measurements from a variety of phases/components. Our estimates of the first and second moments agree well with values derived from slip inversions using local data for both the Kobe and Jalisco events. Additionally, the Jalisco and Aqaba events demonstrate the ability of the technique to resolve the fault plane ambiguity.

One potential source of error in our estimates comes from the affects of unmodelled lateral heterogeneity. We have limited our measurements to combinations of frequency bands and phases that minimize these affects. Also, our data selection process removes any waveforms with large, unmodelled anomalies. The clearest way to evaluate the amplitude of unmodelled heterogeneity is to make GSDF measurements for relatively small earthquakes that should show little to no directivity affects in the frequency bands we work in. Fig. 6 is an example of this using the Kobe earthquake. There are no systematic azimuthal anomalies for this event's Rayleigh waves at frequencies below 15 mHz, and the amplitude of the scatter, which presumably results from unmodelled 3 D heterogeneity, is much smaller than the directivity signal observed for the Jalisco event. We have made measurements for other relatively small earthquakes and find that the δτq measurements for P waves in the band between 20 and 50 mHz typically have rms scatter of 1–2 s, much smaller than the directivity signal observed for the Aqaba event. Future improvements in the accuracy of our synthetic seismograms, such as the inclusion of 3 D attenuation models, will expand the range of frequencies in which we can confidently measure the waveform anomalies resulting from the finite spatial extent of seismic sources.

While regions such as Mexico and Japan have extensive local networks that allow detailed study of their earthquakes, there are many interesting events that do not have sufficient local data sets. For instance, the 1998 Mw 8.2 Antarctic earthquake occurred over 1000 km from the nearest seismometer. We applied the technique described in this paper to measurements of the fundamental mode Rayleigh wave for this event to determine its first and second moments (McGuire et al. 2000). In addition to being able to resolve the fault plane ambiguity for this event, the estimates of the centroid and rupture length, lc, were used by Toda & Stein (2000) to calculate the change in Coulomb failure stress at the locations of this event's aftershocks. In the future, we hope to utilize second moments to help constrain the details of slip distributions from teleseismic data. The moments can be used as integral constraints to help regularize teleseismic finite fault inversions without requiring arbitrary choices of damping parameters. We expect future studies of second moments using the technique developed in this paper to be useful in a wide range of earthquake source, seismotectonic and plate motion studies.

2010 Appendix A:Travelling wave representation of normal-mode summation synthetic seismograms

Here we present an outline of the derivation of eq. (14), the travelling wave representation of our synthetic seismograms. The expression for a synthetic seismogram due to a point source computed by the summation of normal modes can be written as

A1

where n is the overtone number, l is the angular degree, Plm is the associated Legendre polynomial and nIlR and nElS are differential operators related to the excitation of the kth mode [k=(n, l)] at the receiver and by the source, respectively (Zhao & Jordan 1998, Appendix A). Using the Poisson sum formula (Dahlen & Tromp 1998, Chapter 11), the sum over angular degree l can be transformed into an integral over continuous wavenumber λ:

A2

Substituting in the asymptotic expression for P(−cos Δ),

A3

and performing the contour integral in the upper half plane gives

A4

where λn is the wavenumber at which ωn(λ)=ω, vn(λn) is the group velocity at λn, and the index j relates to the orbit number. The corrections made to our synthetics for lateral heterogeneity lead to a slight shift in the eigenfrequencies of the modes such that λn→λn+δλn.

For first orbit waves, j=0 and only the first exponential term (corresponding to waves travelling in the direction from the source to the receiver along the minor arc) is retained, leading to eq. (14):

A5

where the expressions for An and Bn are

A6
A7

with the source scalars

A8
A9
A10
A11
A12
A13

with

A14
A15

and the receiver scalars

A16
A17

where v=(vr, vθ, vφ) is the unit directional vector of the instrument and a dot over an eigenfunction (U, V, W) represents a radial derivative. The source and receiver locations are rS=(rS, θR, φR) and rR=(rR, θR, φR). Results for other odd orbit waves, such as R3, are obtained in the same way with appropriate values of j. The expressions for even orbit waves, such as R2, are found from eq. (A4) by taking the second exponential term corresponding to waves travelling along the major arc with the appropriate values of j.

2010 Appendix B:Non-linear constraints on the moments

Here we show that the matrix inequality constraint we incorporate in our inversion (22) encompasses the constraint derived from Bessel inequalities by Bukchin (1995). For the planar fault case treated by Bukchin, eq. (22) is equivalent to

B1

where x and y are the Cartesian spatial coordinates and t is the time coordinate. One property of a positive semi definite matrix is that its upper left determinates are non negative, i.e.

B2

and

B3

The determinant in eq. (B2) is

B4

Solving for the second temporal moment assuming eq. (B3) holds gives

B5

which can be written in terms of the second moments as

B6

where Adj denotes the adjoint of a matrix. Eq. (B6) is equivalent to the bessel inequality, eq. (24), of (Bukchin 1995).

Acknowledgments

We thank Satoshi Ide for providing us with the centroid of his slip distribution for the Kobe event, and Tim Melbourne for providing his slip distribution for the Jalisco event. Figures in this paper were generated using the GMT software freely distributed by Wessel & Smith (1991). This research was sponsored by the National Science Foundation under grant EAR 9805202.

References

Backus
G.
Mulcahy
M.
,
1976a
.
Moment tensors and other phenomenological descriptions of seismic sources—I. Continuous displacements
,
Geophys. J. R. astr. Soc
,
46
,
341
361
.

Backus
G.
Mulcahy
M.
,
1976b
.
Moment tensors and other phenomenological descriptions of seismic sources—II. Discontinuous displacements
,
Geophys. J. R. astr. Soc
,
47
,
301
329
.

Backus
G.E.
,
1977a
.
Interpreting the seismic glut moments of total degree two or less
,
Geophys. J. R. astr. Soc
,
51
,
1
25
.

Backus
G.E.
,
1977b
.
Seismic sources with observable glut moments of spatial degree two
,
Geophys. J. R. astr. Soc
,
51
,
27
45
.

Ben-Menahem
A.
,
1961
.
Radiation of seismic waves from finite moving sources
,
Bull. seism. Soc. Am
,
51
,
403
435
.

Bevington
P.R.
Robinson
D.K.
,
1992
.
Data Reduction and Error Analysis for the Physical Sciences
,
McGraw-Hill
, New York.

Bukchin
B.G.
,
1995
.
Determination of stress glut moments of total degree 2 from teleseismic surface wave amplitude spectra
,
Tectonophysics
,
248
,
185
191
.

Dahlen
F.A.
Tromp
J.
,
1998
.
Theoretical Global Seismology
,
Princeton University Press
, Princeton, NJ.

Das
S.
Kostrov
B.V.
,
1997
.
Determination of the polynomial moments of the seismic moment rate density distributions with positivity constraints
,
Geophys. J. Int
,
131
,
115
126
.

Doornbos
D.J.
,
1982a
.
Seismic moment tensors and kinematic source parameters
,
Geophys. J. R. astr. Soc
,
69
,
235
251
.

Doornbos
D.J.
,
1982b
.
Seismic source spectra and moment tensors
,
Phys. Earth planet. Inter
,
30
,
214
227
.

Dziewonski
A.M.
Anderson
D.L.
,
1981
.
Preliminary reference Earth model
,
Phys. Earth planet. Inter
,
25
,
297
356
.

Dziewonski
A.M.
Ekström
G.
Salganik
M.P.
,
1996
.
Centroid-moment tensor solutions for January-March 1995
,
Phys. Earth planet. Inter
,
93
,
147
157
.

Dziewonski
A.M.
Ekström
G.
Salganik
M.P.
,
1997
.
Centroid-moment tensor solutions for October-December 1995
,
Phys. Earth planet. Inter
,
101
,
1
12
.

Ekström
G.
,
1989
.
A very broad band inversion method for the recovery of earthquake source parameters
,
Tectonophysics
,
166
,
73
100
.

Ekström
G.
Tromp
J.
Larson
E.W.
,
1997
.
Measurements and global models of surface wave propagation
,
J. geophys. Res
,
102
,
8137
8157
.

Gaherty
J.B.
Jordan
T.H.
Gee
L.S.
,
1996
.
Seismic structure of the upper mantle in a central pacific corridor
,
J. geophys. Res
,
101
,
22 291
22 309
.

Gee
L.S.
Jordan
T.H.
,
1992
.
Generalized seismological data functionals
,
Geophys. J. Int
,
111
,
363
390
.

Gilbert
F.
,
1970
.
Excitation of the normal-modes of the Earth by earthquake sources
,
Geophys. J. R. astr. Soc
,
23
,
119
123
.

Gusev
A.A.
Pavlov
V.M.
,
1988
.
Determination of space-time structure of a deep earthquake source by means of power moments
,
Tectonophysics
,
152
,
319
334
.

Ide
S.
Takeo
M.
,
1997
.
Determination of constitutive relations of fault slip based on seismic wave analysis
,
J. geophys. Res
,
102
,
37 379
27 391
.

Ide
S.
Takeo
M.
Yoshida
Y.
,
1996
.
Source process of the 1995 Kobe earthquake: determination of spatio-temporal slip distribution by Bayesian modeling
,
Bull. seism. Soc. Am
,
86
,
547
566
.

Ihmlé
P.F.
,
1998
.
On the interpretation of subevents in teleseismic waveforms: the 1994 bolivia deep earthquake revisited
,
J. geophys. Res
,
103
,
17 919
17 932
.

Katzman
R.
Zhao
L.
Jordan
T.H.
,
1998
.
High-resolution, two-dimensional vertical tomography of the central pacific mantle using scs reverberations and frequency-dependent travel times
,
J. geophys. Res
,
103
,
17 933
17 971
.

Klinger
Y.
Rivera
L.
Haessler
H.
Christopre Maurin
J.
,
1999
.
Active faulting in the Gulf of Aqaba: new knowledge from the mw 7.3 earthquake of 22 November 1995
,
Bull. seism. Soc. Am
,
89
,
1025
1036
.

McGuire
J.J.
Zhao
L.
Jordan
T.H.
,
2000
.
Rupture dimensions of the 1998 antarctic earthquake from low-frequency waves
,
Geophys. Res. Lett
,
27
,
2305
2308
.

Melbourne
T.
Carmichael
I.
DeMets
C.
Hudnut
K.
Sanchez
O.
Stock
J.
Suarez
G.
Webb
F.
,
1997
.
The geodetic signature of the M8.0 Oct. 9, 1995 Jalisco subduction earthquake
,
Geophys. Res. Lett
,
24
,
715
718
.

Okal
E.A.
,
1982
.
Higher moment excitation of normal modes and surface waves
,
J. Phys. Earth
,
30
,
1
31
.

Pavlov
V.M.
,
1994
.
On non-uniqueness of the inverse problem for a seismic source—II. Treatment in terms of polynomial moments
,
Geophys. J. Int
,
119
,
487
496
.

Pinar
A.
,
1997
.
Source inversion of the 1993 and 1995 Gulf of Aqaba earthquakes
,
Tectonophysics
,
283
,
279
288
.

Silver
P.G.
Jordan
T.H.
,
1983
.
Total-moment spectra of fourteen large earthquakes
,
J. geophys. Res
,
88
,
3273
3293
.

Su
W.J.
Woodward
R.L.
Dziewonski
A.M.
,
1994
.
Degree 12 model of shear velocity heterogeneity in the mantle
,
J. geophys. Res
,
99
,
6945
6980
.

Toda
S.
Stein
R.
,
2000
.
Did stress triggering cause the large off-fault aftershocks of the 25 March 1998 Mw=8.1 Antarctic plate earthquake?
,
Geophys. Res. Lett
,
27
,
2301
2304
.

Tukey
J.W.
,
1984
.
The Collected Works of John W. Tukey
,
Wadsworth
, Belmont, CA.

Vandenberghe
L.
Boyd
S.
,
1996
.
Semidefinite programming
,
SIAM Rev
,
38
,
49
95
.

Wald
D.J.
Heaton
T.H.
,
1994
.
Spatial and temporal distribution of slip for the 1992 landers, california, earthquake
,
Bull. seism. Soc. Am
,
84
,
668
691
.

Wessel
P.
Smith
W.H.F.
,
1991
.
Free software helps map and display data
,
EOS, Trans. Am. geophys. Un
,
72
,
441, 445
446
.

Woodhouse
J.H.
Dziewonski
A.M.
,
1984
.
Mapping the upper mantle: three-dimensional modeling of Earth structure by inversion of seismic waveforms
,
J. geophys. Res
,
89
,
5953
5986
.

Zhao
L.
Jordan
T.H.
,
1998
.
Sensitivity of frequency-dependent travel-times to laterally heterogeneous, anisotropic earth structure
,
Geophys. J. Int
,
133
,
683
704
.