<<

Geophys. J. Int. (2000) 141, 374–390

How are vertical measurements affected by variations in the orientation of azimuthal with depth?

Rebecca L. Saltzer, James B. Gaherty* and Thomas. H. Jordan Department of , Atmospheric and Planetary Sciences, Massachusetts Institute of T echnology, Cambridge, MA 02139, USA. E-mail: [email protected]

Accepted 1999 December 1. Received 1999 November 22; in original form 1999 May 20

SUMMARY Splitting measurements of teleseismic shear waves, such as SKS, have been used to estimate the amount and direction of upper anisotropy worldwide. These measurements are usually made by approximating the anisotropic regions as a single, homogeneous layer and searching for an apparent fast direction (w˜) and an apparent splitting time (Dt˜) by minimizing the energy on the transverse component of the back- projected seismogram. In this paper, we examine the validity of this assumption. In particular, we use synthetic seismograms to explore how a vertically varying anisotropic medium affects shear wave splitting measurements. We find that weak heterogeneity causes observable effects, such as frequency dependence of the apparent splitting parameters. These variations can be used, in principle, to map out the vertical variations in anisotropy with depth through the use of Fre´chet kernels, which we derive using perturbation theory. In addition, we find that measurements made in typical frequency bands produce an apparent orientation direction that is consistently different from the average of the medium and weighted towards the orientation of the anisotropy in the upper portions of the model. This tendency of the measurements to mimic the anisotropy at the top part of the medium may explain why shear wave splitting measurements tend to be correlated with surface geology. When the heterogeneity becomes stronger, multiple scattering reduces the amplitude of the tangential-component seismogram and the associated splitting time, so that a null result may be obtained despite the fact that the waves have travelled through a strongly anisotropic medium. Regardless of the amount of vertical heterogeneity, we find that there is very little dependence on backazimuth for the measured fast-axis direction or splitting time if the top and bottom halves of the medium average to similar fast-axis directions. If, however, the average fast-axis direction in the top half of the model differs from that in the bottom half, splitting-time measurements will show a significant dependence on backazimuth, but fast-axis direction measurements will remain relatively constant. Key words: Fre´chet derivatives, layered media, , shear-wave splitting.

(e.g. Forsyth 1975; Nataf et al. 1984; Tanimoto & Anderson 1 INTRODUCTION 1985; Montagner & Tanimoto 1990), the last decade has seen Measurements of seismic anisotropy are used to infer mantle an explosion in shear wave splitting studies using vertically deformation and flow patterns. While several different methods propagating shear waves (see reviews by Silver 1996 and for constraining anisotropy have been developed, Savage 1999). Typically, these analyses are performed on waves such as Pn refraction surveys (e.g. Raitt et al. 1969; Shearer such as SKS or SKKS because they have a known & Orcutt 1986) and surface wave polarization analyses direction (SV) as a result of passing through the outer core. The standard procedure is to find the inverse splitting * Now at: School of Earth and Atmospheric Sciences, Georgia Institute operator C−1 which, when applied to the observed waveform, of Technology, Atlanta, GA, USA. minimizes the energy on the tangential component (Silver &

374 © 2000 RAS Shear wave splitting with variable anisotropy 375

Chan 1991). When other phases such as S and ScS are used, the 2 PROBLEM FORMULATION splitting parameters are found either by assuming a rectilinear source mechanism (Ando & Ishikawa 1982; Ando 1984) or Because our purpose is to investigate some elementary aspects by explicitly diagonalizing the covariance matrix of surface- of vertical shear , we adopt a very simple corrected horizontal particle motions (Vidale 1986; Fouch & model for the mantle comprising a heterogeneous, anisotropic Fischer 1996). layer of thickness d overlying a homogeneous, isotropic half- A basic assumption in interpreting measurements using these space (Fig. 1). The anisotropy is assumed to be hexagonally techniques is that the splitting operator C corresponds to a single symmetric with a horizontal axis of symmetry, and the lateral ffi homogeneous layer in which the anisotropy has a horizontal heterogeneity is assumed to be su ciently smooth that hori- symmetry axis and a constant magnitude. The parameters zontal gradients in the wave velocities can be ignored. Vertically used to describe this model (the splitting parameters) are the propagating shear waves can thus be represented as linear polarization azimuth of the fast eigenwave, w, and the travel- combinations of orthogonal eigenwaves with shear velocities v and v that depend on the depth coordinate z. To simplify time difference between the fast and slow eigenwaves, Dt.Itis 1 2 the problem further, we assume that the mean velocity straightforward to construct the splitting operator for an v: =(v +v ) 2 and the velocity difference v=v −v are con- arbitrary stack of layers with depth-dependent properties and 1 2 / D 1 2 stants, and we label the eigenwaves such that Dv>0. The more general forms of anisotropy using propagator matrices heterogeneity in the medium is specified by a single function (e.g. Keith & Crampin 1977; Mallick & Frazer 1990), but it is of depth that we take to be the azimuth of the fast (v ) axis, less clear how one might use such constructions to make 1 w(z), measured clockwise from the x-axis. inferences about anisotropic structure. A potentially fruitful For the calculations in this paper, we adopt a layer thick- direction is to fit the waveform data by optimizing the homo- ness of d=200 km and a mean velocity of v: =4.54 km s−1, genous-layer operator and then interpret the two recovered and we take the velocity of the isotropic half-space to equal ˜ quantities, denoted here by w and Dt˜, as apparent splitting this mean velocity. The maximum splitting time for shear waves parameters which are functionals of the vertical structure. This propagating from the base of the anisotropic layer to the approach was adopted by Silver & Savage (1994), who showed surface—we ignore the —is Dt=d(v −v )/v v #dDv/v:2, ˜ 1 2 1 2 how an approximation to the variation of w and Dt˜ with the which represents the ‘splitting strength’ of the model. We refer incident polarization angle could be inverted for a two-layer to Dt in some of our numerical experiments as the ‘true’ anisotropic model. They also discussed the generalization of splitting time. their approximate functional relations, which are valid for forward scattering at low frequencies (wave periods &Dt), to an arbitrary layer stack. Ru¨mpker & Silver (1998) have recently 2.1 Forward problem: synthetic seismograms expanded this theoretical discussion of vertical heterogeneity We use a stack of thin, homogeneous layers to represent the to include expressions for the apparent splitting parameters medium, and a propagator-matrix method to propagate shear valid at high frequencies, as well as some statistical properties waves vertically through the layers (e.g. Kennett 1983). For all of the parameters for random layer stacks, and they have tested our calculations, these layers are less than 1 km thick, in order various aspects of their theory with numerical calculations. to ensure that seismic wavelengths do not approach layer thick- In this paper, we consider several additional aspects of this ness. Boundary conditions restrict displacements and tractions interpretation problem. Using a propagator-matrix method to be continuous at the interfaces between the layers, and that includes both forward- (upgoing) and back-scattered tractions to be zero at the surface. The Fourier-transformed, (downgoing) waves, we compute synthetic seismograms for various types of depth dependence, including smooth models as well as those with discontinuous variations in the anisotropy axis. We investigate the behaviour of the apparent splitting parameters with increasing amounts of vertical heterogeneity x in the azimuthal anisotropy, and use the results to define three wave-propagation regimes corresponding to weak, inter- z = 0 y mediate, and strong scattering. For weakly heterogeneous media, z Anisotropic layer we employ perturbation theory to calculate the sensitivity φ( z ) v , v constant 12 (Fre´chet) kernels for band-limited, apparent-splitting measure- axis of hexagonal ments, and show how these measurements sample the depth symmetry dependence as a function of frequency and incident polarization z = d angle. In realistic situations, the centre frequencies of the observations are sufficiently small that the kernels are one- Isotropic halfspace sided, and we can define an apparent depth of sampling that we demonstrate is biased towards the upper part of the plane structure. In principle, the Fre´chet kernels can be used to set up the problem of inverting frequency-dependent splitting Figure 1. Model used in the calculations. A vertically travelling, rectilinearly polarized shear wave impinges at depth d on the base of measurements for depth-dependent anisotropy. We show that a heterogeneous, anisotropic layer in which the fast-axis direction w in practice, however, strong scattering by vertical heterogeneity varies as a function of depth z. The two eigenvelocities, v1 and v2, can invalidate the assumptions that underlie this linearized are constant throughout the layer, and the velocity of the isotropic approach, especially at higher frequencies. half-space is taken to be equal to their mean.

© 2000 RAS, GJI 141, 374–390 376 R. L . Saltzer, J. B. Gaherty and T . H. Jordan

= −1 T density-normalized stress vector t(z, v) r [TxzTyz] is related to the depth derivative of the displacement vector (a) input = T ff u(z, v) [uxuy] by the Christo el matrix R 2 2 + 2 2 2− 2 v1 cos w(z) v2 sin w(z) (v1 v2) cos w(z) sin w(z) C(z)= . C 2− 2 2 2 + 2 2 D (v1 v2) cos w(z) sin w(z) v1 sin w(z) v2 cos w(z) T (1) ∂ = The are z f Af, where the displacement- stress vector and system matrix are given by

u(z, v) (b) homogeneous layer f (z, v)= , (2) C t(z, v) D R 0C−1(z) A(z, v)= . (3) C−v2I0D T The rotation operator

cos w(z) −sin w(z) U(w(z))= (4) C sin w(z) cos w(z) D o ff ˆ ¬ 2 2 = T (c) fast axis rotates 30 diagonalizes the Christo el matrix: C diag[v1, v2] UCU . The propagator matrix for this problem and some of its approximations are discussed in Appendix A. For an upgoing R wave uI(v) incident at the base of the anisotropic layer, the free-surface displacement vector can be written T = + + − u(0, v) [Puu ivv: Put (Puu ivv: Put)R]uI(v), (A16) × where Puu and Put are 2 2 submatrices of the propagator matrix (A7), R is the 2×2 matrix of reflection coefficients (A15), and U =U(w(z)). z o The pulse shape at the base of the anisotropic layer in (d) fast axis rotates 100 = all of our calculations is taken to be of the form uI(t) − − − − − = R exp[ a/(t t0) (t t0)/b] H(t t0) with a duration a 2s and a decay constant b=4 s. The convolution of this initial pulse shape with the broad-band instrument response and pre- T filter is given in Fig. 2(a). There is no energy on the tangential component because the initial pulse is radially polarized; how- ever, propagation of the pulse through an anisotropic layer produces energy on both the radial and tangential components of the surface seismogram. An example of a synthetic seismo- gram calculated for a homogeneous anisotropic layer is shown (e) fast axis rotates 1000o in Fig. 2(b) for a velocity contrast of Dv/v: =4.54 per cent. This corresponds to a splitting strength of Dt=2 s, which R lies towards the high end of the observations summarized by Silver (1996) and Savage (1999). The azimuth of the fast T axis, measured clockwise from the radial (xˆ) direction, is 45°. The seismograms in this example are ‘broad-band’ with a corner at 50 mHz and at 300 mHz, and a centre frequency of ~140 mHz. 40 60 80 100 120 A simple model of a heterogeneous anisotropic layer, used Time (s) extensively in our numerical illustrations, is one in which w(z) varies linearly with depth: Figure 2. Radial- and tangential-component seismograms at (a) the base of the anisotropic layer, (b) the surface after passing through a = + = + − w(z) w0 kz wd k(z d). (5) homogeneous anisotropic layer, (c) the surface after passing through a weakly heterogeneous anisotropic layer in which the fast-axis The constant k=dw/dz is the vertical rotation rate. The direction (Dw) linearly rotates 30°, (d) intermediate heterogeneous orientation of the fast axis varies from w at the surface to w 0 d anisotropic layer in which the fast-axis direction (Dw) linearly rotates at the base of the layer, and the total change in the orientation 100°, and (e) strongly heterogeneous anisotropic layer in which the = − = Dw wd w0 kd measures the strength of the heterogeneity. fast-axis direction (Dw) linearly rotates 1000°. Seismograms are band- Figs 2(c) to (e) show synthetic seismograms for the linear- pass filtered with a Butterworth filter to frequencies between 50 rotation model with a splitting strength of Dt=2 s and and 300 mHz.

© 2000 RAS, GJI 141, 374–390 Shear wave splitting with variable anisotropy 377 increasing amounts of heterogeneity: Dw=30°, 100°, and 1000°, The behaviours illustrated in Fig. 3 are typical of three respectively. In all three examples, the incident polarization scattering regimes that can be qualitatively described as ‘weak’, was chosen such that the mean orientation of the fast axis was ‘strong coherent’, and ‘strong incoherent’. 45°; that is,

1 d : ¬ = + = w P w(z)dz w0 Dw/2 p/4 . (6) 3 WEAK-SCATTERING REGIME d 0 When the scattering is weak, the effects of the heterogeneity In the case where Dw=30°, the surface seismograms (Fig. 2c) on the apparent splitting parameters can be approximated are very similar to those produced in the homogeneous with a linearized perturbation theory. In this section, we derive anisotropic case (Fig. 2b). Both the radial- and tangential- analytical expressions for perturbations from a homogeneous component seismograms have a comparable amount of energy. starting model, test their applicability with numerical calcu- As Dw increases to 100°, the energy on the radial com- lations, and use them to gain insight into how the sensitivity ponent becomes greater than that on the tangential component of the apparent splitting parameters varies with depth. (Fig. 2d), and when Dw is as large as 1000°, the net effect of the anisotropy is to put very little energy onto the tangential component (Fig. 2e). 3.1 Fre´chet kernels The linearized equations that relate a structural perturbation 2.2 Inverse problem: apparent splitting parameters dw(z) to perturbations in the apparent splitting parameters, dw˜ and dt˜, can be written as integrals over the layer: If the incident pulse is known to be radially polarized, then ˜ the apparent splitting parameters w and Dt˜ can be defined as d ˜ the values that minimize the energy on the transverse (yˆ) dw# P Gw(z)dw(z)dz , (8) component of the displacement field back-projected to z=d 0 using a homogeneous-layer splitting operator (Silver & Chan d 1991). Parseval’s theorem allows the transverse-component dt˜# P Gt(z)dw(z)dz . (9) energy to be written as a frequency-domain integral: 0

2 This approximation ignores terms of order dw2. G (z) and G (z) e2(w∞, Dt∞)= |yˆTC−1(w∞, Dt∞)u(0, v)|2dv . (7) w t P h are the sensitivity functions, or Fre´chet kernels. They generally −2 depend on the structural model that is being perturbed, as −1 Here Ch is the inverse of the splitting operator given by well as on the spectral properties of the waves being measured. eq. (A24). In practice, the determination of these so-called For the structures considered here, a model is specified by the splitting parameters requires a search over a grid of fast-axis fast-axis orientation function {w(z): 0≤z≤d} and the two directions and delay times. For typical teleseismic observations, velocity constants v and v . ± ° 1 2 these parameters can be determined to within 10 and 0.15 s The Fre´chet kernels can be numerically approximated for (Fouch & Fischer 1996; Silver & Chan 1991). Fig. 3 shows the an arbitrary starting model by computing the small change in ∞ ° ° ‘energy map’ contoured as a function of w (0 to 180 ) and the splitting parameters due to a small perturbation in the ∞ Dt (0–4 s) for the three linear-gradient examples shown in the fast-axis orientation distributed over a thin layer. However, in previous section, and Fig. 4 shows the inferred seismograms at = the special case of a homogeneous starting model (w(z) w0), the base of the layer, calculated by back-projecting the splitting the kernels for narrow-band pulses can be derived analytically. parameters and assuming propagation of waves through a The details are relegated to Appendix B. The approximate homogeneous, anisotropic layer. results for a pulse with centre frequency v and half-bandwidth The energy map for the seismograms computed for a homo- 0 s are geneous layer (Fig. 3a) shows a well-defined minimum at the correct values of splitting parameters (w˜ =45°, Dt˜=2s).For + + 2 2 4 g4 [g5 cos 4w0 g6](s /v0) s weak heterogeneity (Dw=30°), the energy minimum remains G = Dk+O , (10) w g +(g cos 4w +g )(s2/v2) Av4B close to the layer mean (w˜ =48°, Dt˜=1.9 s), and the bulk of 1 2 0 3 0 0 the tangential-component energy has been removed (Fig. 4c). g +[g cos 4w +g ](s2/v2) s4 At intermediate values of the heterogeneity (Dw=30°), the = 7 8 0 9 0 + Gt + + 2 2 Dk cot 2w0 O A 4B , energy minimum is still well defined (Fig. 3c), but it is displaced g1 (g2 cos 4w0 g3)(s /v0) v0 ° ˜ = ° away from the layer mean by 18 (w 63 ). Moreover, the (11) scattering from the vertical gradients in the anisotropy is ffi su cient to reduce the apparent splitting time significantly where Dk=v (v−1−v−1) is the differential wavenumber. = 0 2 1 below its true value (Dt˜ 1.5 s). When the heterogeneity gets The three parameters appearing in the denominator of these = ° ffi to be very large (Dw 1000 ), the scattering is su ciently strong expressions are independent of depth z: as to cause destructive interference that nearly wipes out the arrivals on the transverse component (Fig. 2e). The resulting = − 2 g1 (1 cos Dkd) , (12a) energy map (Fig. 3d) is characteristic of a ‘null measurement’, = 2 + − with the lowest values occurring near the horizontal axis g2 Dkd[sin Dkd (Dkd 2) sin Dkd cos Dkd] , (12b) where Dt∞=0 and along vertical ridges corresponding to the = ° ° = + 2 2 + − degenerate azimuths of w 0 and 90 . g3 g2 2(Dkd) [sin Dkd (1 cos Dkd) cos Dkd ] . (12c)

© 2000 RAS, GJI 141, 374–390 378 R. L . Saltzer, J. B. Gaherty and T . H. Jordan

HOMOGENEOUS LAYER WEAK SCATTERING (a)∆φ=0ο (b) ∆φ=30ο 4 4

0.8

3 3 0.8 0.8

0.8 1.2

t (s) 2 2 1.2 0.4 ∆

0.4

1 1

0 0 0 30 60 90 120 150 180 0 30 60 90 120 150 180 φ (deg) φ (deg)

STRONG COHERENT SCATTERING STRONG INCOHERENT SCATTERING (c)∆φ=100ο (d) ∆φ=1000ο 4 4

2 3 3

0.8

1.6

t (s) 2 2

0.8 ∆

1.2 1 0.4 1

0 0 0 30 60 90 120 150 180 0 30 60 90 120 150 180 φ (deg) φ (deg)

Figure 3. Energy diagrams for (a) a homogeneous anisotropic layer, (b) a weakly heterogeneous anisotropic layer (Dw=30°), (c) an intermediate heterogeneous anisotropic layer (Dw=100°), and (d) a strongly heterogeneous anisotropic layer (Dw=1000°).

The other six can be written in terms of trigonometric functions The relative bandwidth s/v0, which is less than 0.5 in most of the height variable r=d−z: seismological applications (e.g. Silver 1996; Fouch & Fischer 1996; Wolfe & Solomon 1998), is sufficiently small that it is g (r)=(1−cos Dkd) sin Dkr , (13a) 4 safe to ignore the fourth-order terms in (10) and (11). Fig. 5 = − + − g5(r) (Dkd 1) sin Dkd cos Dkr (Dkd DkdDkr) sin Dkd displays kernels computed under this approximation for a range of initial polarizations and centre frequencies. The ×sin Dkr−Dkd cos Dkd cos Dkr , (13b) sinuosity of the kernels increases with frequency, reflecting the = + 2 + g6(r) g5(r) (Dkd) cos Dkd sin Dkr 2(Dkd)(Dkr) sin Dkd first-order trigonometric dependence on Dkz. The kernel for × − 2 − the apparent splitting azimuth satisfies the lower boundary cos Dkr (Dkr) (1 cos Dkd) sin Dkr , (13c) = condition, Gw(d) 0, which can be verified from the analytical = − 2 − − g7(r) (1 cos Dkd) cos Dkr (1 cos Dkd) sin Dkd sin Dkr , expressions. Integration of these expressions show that the kernel for dw˜ (14a) is unimodular and the kernel dt˜ averages to zero: g (r)=sin Dkd(sin Dkd−2 cos Dkd) 8 d = ×[Dkd cos Dkr−(Dkd)(Dkr) sin Dkr] P Gw(z)dz 1, (15) 0 +( kd)2 sin kd cos kd cos kr , (14b) D D D D d = g (r)=g (r)+2(Dkd)2 P Gt(z)dz 0. (16) 9 8 0 ×[sin2 Dkd+(1−cos Dkd) cos Dkd ] cos Dkr These properties must apply to the exact forms of the Fre´chet −(1−cos Dkd)2[Dkr sin Dkr+(Dkr)2 cos Dkr] . (14c) kernels, not just to their narrow-band approximations given

© 2000 RAS, GJI 141, 374–390 Shear wave splitting with variable anisotropy 379

by (10) and (11), because a constant perturbation maintains (a) input the structural homogeneity of the starting model. In other words, setting the perturbation in (8) and (9) to a constant R value of dw must always yield dw˜ =dw and dt˜=0.

T 3.2 Limiting forms of the kernels By considering the zero-bandwidth limit, we gain additional insight into the nature of the sensitivity kernels:

sin k(d−z) 0 ¬ = D Gw(z) lim Gw(z) Dk, (17) (b) homogeneous layer s0 1−cos Dkd 0 ¬ R Gt (z) lim Gt(z) s0 sin Dkd sin Dk(d−z) T = cos Dk(d−z)− Dk cot w . (18) C 1−cos Dkd D 0

The splitting-time kernel (18) varies like the cotangent of 2w0 = and is thus singular at w0 np/2, which corresponds to incident polarizations aligned with an eigenwave orientation in the (c) fast axis rotates 30 o unperturbed model. In contrast to this singular behaviour, the splitting-azimuth kernel (17) is independent of the initial R polarization, and remains well defined even at its degenerate values. T This important theoretical point deserves special emphasis. A shear wave with a polarization aligned with one of the eigenwave directions is not split by propagation through the reference model, and the tangential-component energy of the back-projected displacement field is thus identically zero ∞= = ∞> at w w0 np/2 for arbitrary values of Dt 0. Consequently, (d) fast axis rotates 100o the energy map displays vertical nodal lines at these azimuths, as well as a horizontal nodal line at Dt∞=0, and the inversion R of the seismograms for the apparent splitting parameters via the minimization of (7) becomes unstable (e.g. Silver & Chan T 1991). Nevertheless, the apparent splitting azimuth remains formally defined in this limit by the orientation of the appro- ˜ + priate vertical node; that is, w equals either w0 or w0 p/2. Moreover, w˜ is Fre´chet differentiable, because a small per- turbation to the model will result in a small, well-defined perturbation of the node in the w∞ direction. The splitting-time (e) fast axis rotates 1000o functional Dt˜, on the other hand, is not Fre´chet differentiable at the nodes, which is why its kernel is singular. This behaviour R generalizes to pulse shapes with finite bandwidths, as is evident from eq. (11). The cotangent dependence of the splitting-time kernel also T = = implies that Gt(z) 0 for w0 np/4, which means that at polarization angles near 45° the apparent splitting time is only weakly dependent on perturbations to the local splitting 40 60 80 100 120 orientation. Time (s) Eqs (17) and (18) show that, in the zero-bandwidth limit, both kernels become unbounded at Dkd=2 np, where the total Figure 4. Radial- and tangential-component seismograms. After deter- splitting time Dt is an integral multiple of the wave period mining the apparent splitting parameters, the seismograms can be ~ 2 2 2p/v0. A finite bandwidth introduces a weak ( s /v0) depen- back-projected to the base of the layer again using those same dence of Gw and Gt on the initial azimuth through the cos 4w0 parameters and assuming a homogeneous layer. (a) Input pulse at the term, which suppresses this resonance singularity. base of the anisotropic layer to be compared with back-projected The expressions for the kernels presented thus far apply to seismograms from (b) homogeneous anisotropic layer, (c) weakly splitting strengths that are arbitrarily large. At low frequencies, heterogeneous anisotropic layer (Dw linearly rotates 30°), (d) inter- mediate heterogeneous anisotropic layer (Dw linearly rotates 100°), when the splitting strength is much less than the wave period, and (e) strongly heterogeneous anisotropic layer (Dw linearly rotates the maximum phase shift between the two eigenwaves is small, = 1000°). Seismograms are band-pass filtered with a Butterworth filter Dkd v0Dt%1. If we expand the trigonometric functions and to frequencies between 50 and 300 mHz. retain only the leading terms, the single-frequency kernels

© 2000 RAS, GJI 141, 374–390 380 R. L . Saltzer, J. B. Gaherty and T . H. Jordan

(a)K (b) K φ ∆t 0 0 20 deg

100 100 depth (km)

200 200 0.04 0.02 0 0.02 0.04 0.04 0.02 0 0.02 0.04

0 0 40 deg

100 100 depth (km)

200 200 0.04 0.02 0 0.02 0.04 0.04 0.02 0 0.02 0.04

0 0 60 deg

100 100 depth (km)

200 200 0.04 0.02 0 0.02 0.04 0.04 0.02 0 0.02 0.04

0 0 80 deg

100 100 depth (km)

200 200 0.04 0.02 0 0.02 0.04 0.04 0.02 0 0.02 0.04

Figure 5. Fre´chet kernels for the apparent splitting parameters computed from eqs (10) and (11) to second order in the relative bandwidth s/v0. = The reference model is a 200-km thick homogeneous layer with Dt 2 s. Left panels show the apparent azimuth kernels Gw(z), and right panels ° ° show the apparent splitting-time kernels Gt(z). The initial polarization w0 increases downwards from 20 (top panel) to 80 (bottom panel). In each panel, the centre frequencies v0 range from 0.1 to 0.8 Hz (lowest frequencies are solid lines, intermediate are dashed, and highest frequencies = are dotted), while the relative bandwidth is held constant at s/v0 0.125. become linear functions of depth: of the apparent splitting time to azimuthal heterogeneity is zero in the middle of the layer, and it is of equal magnitude 2(d−z) G0(z)# , (19) and opposite sign at the top and bottom of the layer. w d2 2z−d G0(z)#2Dt cot w . (20) 3.3 Apparent depth of sampling t A d2 B 0 The previous discussion shows that the Fre´chet kernel for the Thus, in the low-frequency limit, the apparent splitting azimuth apparent splitting azimuth will be non-negative when the centre is insensitive to heterogeneity at the base of the layer and most period of the wavegroup, 2p/v0, is greater than or equal to sensitive to heterogeneity at the top of the layer. The sensitivity twice the splitting time Dt (i.e. Dkd≤p). Under this condition,

© 2000 RAS, GJI 141, 374–390 Shear wave splitting with variable anisotropy 381 which applies to most observations of teleseismic shear wave prediction, corresponding to the fast-axis direction approxi- splitting, we can define an apparent depth of sampling by the mately one-third of the way down the layer. This increase in centroid of the kernel: sensitivity to the fast-axis direction near the surface may explain d why shear wave splitting measurements tend to correlate with = tectonic deformation observed at the surface (Silver 1996). zapp P Gw(z)zdz . (21) 0 Most shear wave splitting measurements are made on seismo- < Using (15), we obtain for the zero-bandwidth limit grams with relatively low centre frequencies ( 200 mHz), so that the apparent depth of sampling is less than the mean sin kd− kd cos kd 0 = = D D D thickness of the layer. There are significant variations in the zapp lim zapp . (22) s0 Dkd(1−cos Dkd) centre frequencies used by different researchers, however, so this apparent depth varies from study to study. In the low-frequency limit, where the w˜ kernel becomes a linear function of depth, the apparent depth of sampling goes to one-third of the layer thickness. This value increases to half 3.4 Numerical tests of the layer thickness as the centre period approaches the We conducted a series of numerical experiments to test the splitting time. Fig. 6 shows three low-frequency w˜ kernels and perturbation theory derived for a homogeneous model. To their apparent depths of sampling for a 200-km thick, aniso- investigate the sensitivity of the kernels to the reference tropic layer. In the low-frequency limit, the apparent depth of structure, we have computed them by numerical perturbation sampling is just 66 km, whereas, for data low-pass filtered at to heterogeneous starting models. The results for starting 100 mHz, the depth of sampling increases to 71 km, and, for models with a linear gradient and a step-wise discontinuity in data low-pass filtered at 200 mHz, it increases to 88 km. w(z) are compared with the homogeneous-layer case in Fig. 7. This bias in the sensitivity to near-surface structure explains The average orientation was chosen to be the same for all why the recovered value for w˜ in the weak scattering case three models, w: ¬d−1 ∆d w(z)dz=45°, while the total variation (Fig. 3b) is greater than the layer mean by about 3°. In this 0 in w(z) was taken to be 20° for the two heterogeneous models. model, the fast axis rotates linearly from 40° at the base of the The kernels are very similar, indicating only a weak dependence layer to 50° at the top. The kernels predict that the upper part on the starting model when the heterogeneity is of this magni- of the model, where the fast axis ranges between 45° and 50°, tude. In particular, the apparent depths of sampling for the will dominate the shear wave splitting measurement, and layered and linear models are 91 km and 90 km, respectively, indeed the value found numerically (w˜ =48°) agrees with this essentially the same as the value of 88 km calculated for the homogeneous model. The properties found for the homo- Low−Frequency φ kernels geneous-layer kernels, such as their dependence on frequency, 0 bandwidth, and incidence azimuth, should therefore pertain more generally in the weak-scattering regime. We also compared the perturbations calculated from the 20 Fre´chet kernels using eqs (8) and (9) with the results of a direct numerical calculation that minimized the tangential- 40 component energy on back-projected synthetic seismograms. Fig. 8 shows several examples of these comparisons as a : 60 function of the average polarization direction w for an initial 66 km pulse with a centre frequency at 60 mHz and corners at 45 mHz 71 km and 75 mHz. Fig. 9 shows the results for w: =45° for increasing 80 values of the centre frequency v0. We note that care must 88 km be taken in the numerical calculations when evaluating the 100 apparent splitting parameters near the azimuthal nodes at = w0 np/2, because the energy surfaces can be very flat in the depth (km) Dt∞ direction, and the location of the minimum is susceptible 120 to numerical inaccuracies that can cause a p/2 ambiguity. When the heterogeneity is small (Dw=10°) in the linear- 150 gradient models, the kernels do a good job of predicting w˜ and Dt˜ for all backazimuths (Fig. 8a). For heterogeneity with ° 160 a total rotation angle as large as 60 (Fig. 8b), the kernels typically overestimate the apparent splitting time by about 0.3 s, and, near the nodes, w˜ can be off by as much as 20°. This 180 failure of the kernels with greater heterogeneity reflects a breakdown in the small-angle approximations (e.g. eq. B5). 200 Inclusion of the back-scattering terms in the calculation of 0 0.005 0.01 synthetic seismograms (solid dots in Fig. 8) does not produce 1/km significantly different results from those obtained using just Figure 6. Low-frequency Fre´chet kernels and the apparent depth of the forward-scattering terms. sampling computed with eq. (20). The solid line is the kernel in the In a second set of comparisons, we use what is perhaps a low-frequency limit, the dashed line is the low-passed kernel for more geologically relevant model comprising 10 layers, each 100 mHz, and the dotted line is the low-passed kernel at 200 mHz. 20 km thick, of differing orientations constrained such that the

© 2000 RAS, GJI 141, 374–390 382 R. L . Saltzer, J. B. Gaherty and T . H. Jordan

K K reference models φ ∆t 0 0 0 (a) (b) (c) 50 50 50

100 100 100 depth (km) 150 150 150

200 200 200 35 40 45 50 55 0 0.005 0.01 −0.01 −0.005 0 0.005 0.01 phi (deg) 1/km 1/km

Figure 7. A comparison of Fre´chet kernels for three starting models. Panel (a) is a plot of w(z) for the homogeneous-layer (solid line), constant- gradient (dotted line), and two-layer (dashed line) models. Panels (b) and (c) show the corresponding kernels for these models, Gw(z) and Gt(z), = = respectively. The kernels were calculated by a numerical perturbation scheme for v0 0.14 Hz, s v0/6 Hz, which are similar to the values used : = ° = in the processing of teleseismic shear waves. The models have the same average azimuth, w 45 . Eq. (10) shows that Gt(z) 0 for a homogeneous layer with this initial azimuth. The agreement illustrates the weak dependence of the kernels on the starting model.

fast-axis directions varies between 0° and 30°. When the layer at higher frequencies (Figs 9b and d) than at lower frequencies. orientations are distributed randomly, such that the average In these examples, Dt=2.14 s. At frequencies approaching 1/Dt orientation in the top half of the model is similar to that in (~0.46 Hz) the kernels predict highly oscillatory behaviour in the bottom half, the agreement between the numerical results w˜. This also corresponds to the frequency at which the kernels and the predictions of the analytical kernels is usually very become unbounded in the single-frequency limit. good (Fig. 8c). In addition, both the exact values of the apparent splitting parameters and perturbation-theory pre- 4 STRONG-SCATTERING REGIME dictions show very little dependence on the polarization angle, with the variations in the apparent splitting time associated The numerical experiments demonstrate that the first-order with nodal singularities compressed into a narrow range of perturbation theory expressed in eqs (8) and (9) provides an azimuths. accurate description of the apparent splitting parameters in When the layer orientations are skewed, however, such that situations where the magnitude of the vertical heterogeneity is the average orientation in the top half of the model differs small. As this magnitude increases, the perturbation theory significantly from that in the bottom half (Fig. 8d), the p/2 fails because the small-angle approximations employed in periodicity in Dt˜ associated with the nodal singularities obtaining the scattering matrix (B5) and the linearized mini- becomes more pronounced. This difference in the variation of mization condition (B12) become inaccurate, owing to the ff Dt˜ with initial azimuth results from the fact that Gt(z) is accumulating e ects of multiple forward-scattering. The com- approximately a linear function of depth that averages to bination of these strong-scattering effects causes the behaviour zero, as seen from its low-frequency form (18); that is, the of the apparent splitting parameters to deviate from the perturbation to the apparent splitting time will be small and weak-scattering results. the p/2 periodicity will be suppressed when the first moment Aspects of this behaviour were noted in the previous ∆d (z)zdz is small. For a specified level of heterogeneity, the discussion of Fig. 3, which displays the numerical results for 0 w constant-gradient case has the largest first moment of any a constant-gradient model. In these calculations, Dt=2s, model, which is why the initial-azimuth dependence in Figs w: =45°, and all forward- and back-scattering terms were 8(a) and (b) is so pronounced. retained. For a 45° average polarization, the general form of These results can be used to qualify Silver & Savage’s (1994) the kernel (11) shows that the splitting-time perturbation argument that a p/2 periodicity in initial azimuth should should be zero to first order; that is, the apparent splitting be diagnostic of vertical heterogeneity. This periodicity will be time Dt˜ should equal the total splitting strength Dt. Strong relatively weak for heterogeneous structures where the azimuth scattering acts to reduce Dt˜ below this theoretical limit, so that of the anisotropy does not vary systematically with depth. the ratio (Dt−Dt˜)/Dt, to the extent that it can be accurately On increasing the heterogeneity in the 10 random layers so estimated, measures the higher-order effects. For the 30° that the fast-axis direction ranges over 120°, we find azimuthal rotation in Fig. 2(c), this reduction is only about 5 per cent, discrepancies of up to ±5° and splitting-time discrepancies consistent with the weak-scattering approximations. The exceeding 1 s (Figs 8e and f ). At this level of heterogeneity, 120° rotation in Fig. 2(d) gives a much more substantial effect back-scattering effects, given by the differences between the (~35 per cent), indicating that these approximations are not open and solid circles, begin to become important. accurate for heterogeneity of this magnitude. For the 1000° A comparison of the analytical and numerical results for rotation, the scattering is sufficiently large that the tangential- different frequencies at a single backazimuth shows that the component arrivals are incoherent, so that the long-period kernels do a nice job of predicting w˜ and Dt˜ up to 0.5 Hz when amplitude is nearly zero, and the energy diagram looks nodal. the heterogeneity is weak (Figs 9a and c), but that when the In this case, there is no well-defined energy minimum, and it heterogeneity gets stronger, the kernels break down more quickly is difficult to measure the apparent splitting parameters.

© 2000 RAS, GJI 141, 374–390 Shear wave splitting with variable anisotropy 383 ) in which ° 120 = w 80û 80û D 80û 60û 60û 60û t avg avg avg φ φ φ ∆ 40û 40û 40û 20û 20û 20û 0û 0û 0û

4 3 2 1 0

4 3 2 1 0 4 3 2 1 0

3.5 2.5 1.5 0.5 3.5 2.5 (s) 1.5 0.5

3.5 2.5 1.5 0.5 t (s) t (s) t ∆ ∆ ∆ ) linear rotation model. (b) Strongly heterogeneous ° 10 = w ) ) D ) o o o 80û 80û 80û 60û 60û 60û φ avg avg avg φ φ φ 40û 40û 40û forwardscattering fullscattering erent from those in the bottom half. (e) Strongly heterogeneous 10-random-layer forwardscattering fullscattering forwardscattering fullscattering 20û 20û ff 20û 10 Skewed Layers (max rotation = 30 10 Random Layers (max rotation = 120 10 Skewed Layers (max rotation = 120

0û 0û 0û

90û 80û 70û 60û 50û 40û 30û 20û 10û 90û 80û 70û 60û 50û 40û 30û 20û 10û 90û 80û 70û 60û 50û 40û 30û 20û 10û φ φ φ , plotted as a function of backazimuth. Solid lines are analytical predictions from the kernels. ˜ t ) in which the average fast-axis directions in the top and bottom halves are similar. (d) Weakly ° D (f) (e) (d) 30 = w D 80û 80û 80û 60û 60û 60û t avg avg avg φ φ φ , and right panels show ∆ ˜ w 40û 40û 40û 20û 20û 20û 0û 0û

4 3 2 1 0 4 3 2 1 0 4 3 2 1 0

3.5 2.5 1.5 0.5 3.5 2.5 1.5 0.5 (s) 3.5 2.5 1.5 0.5 t (s) t

(s) t ∆ ∆ ∆ ) in which the layer orientations in the top half are di ° 30 = w ) D o 80û 80û 80û erent from those in the bottom half. ff 60û 60û 60û φ avg avg avg φ φ φ 40û 40û 40û Constant Rotation Constant Rotation o o 10 60 forwardscattering fullscattering forwardscattering fullscattering ) in which the average fast-axis directions in the top and bottom halves are similar. (f ) Strongly heterogeneous 10-random-layer model (maximum forwardscattering fullscattering ° 20û 20û 20û 120 10 Random Layers (max rotation = 30 =

w

0û 0û 0û

90û 80û 70û 60û 50û 40û 30û 20û 10û 90û 80û 70û 60û 50û 40û 30û 20û 10û 90û 80û 70û 60û 50û 40û 30û 20û 10û φ φ φ D (c) (a) (b) Comparison of numerical and analytical results. Left panels show ) linear rotation model. (c) Weakly heterogeneous 10-random-layer model (maximum ° 60 = w D the layer orientations in the top half are di ( model (maximum heterogeneous 10-random-layer model (maximum Solid circles are the numerical results including back-scattering, and open circles include just forward-scattering. (a) Weakly heterogeneous ( Figure 8. © 2000 RAS, GJI 141, 374–390 384 R. L . Saltzer, J. B. Gaherty and T . H. Jordan

φ ∆t

10o Constant Rotation (a) 80û 4 70û 3.5 60û forwardscattering 3 50û fullscattering 2.5 40û

φ 2 (s) 30û t ∆ 1.5 20û 1 10û

0û 0.5

10û 0 0.05 0.1 0.2 0.3 0.5 0.05 0.1 0.2 0.3 0.5 frequency (Hz) frequency (Hz)

60o Constant Rotation (b) 80û 4 70û forwardscattering 3.5 60û fullscattering 3 50û 2.5 40û (s)

t 2 φ

30û ∆ 1.5 20û 1 10û

0û 0.5

10û 0 0.05 0.1 0.2 0.3 0.5 0.05 0.1 0.2 0.3 0.5 frequency (Hz) frequency (Hz)

10 random layers (max = 30o) (c) 90û 4 80û 3.5 70û forwardscattering 3 60û fullscattering 2.5 50û (s) φ

t 2

40û ∆ 1.5 30û 1 20û

10û 0.5

0û 0 0.05 0.1 0.2 0.3 0.5 0.05 0.1 0.2 0.3 0.5 frequency (Hz) frequency (Hz)

10 random layers (max = 120o) (d) 90û 4 80û 3.5

70û 3 60û 2.5 50û (s) φ t 2

40û ∆ 1.5 30û forwardscattering 1 20û fullscattering

10û 0.5

0û 0 0.05 0.1 0.2 0.3 0.5 0.05 0.1 0.2 0.3 0.5 frequency (Hz) frequency (Hz)

Figure 9. Comparison of numerical and analytical results. Left panels show w˜ and right panels show Dt˜, plotted as a function of frequency. Solid lines are analytical predictions from the kernels. Solid circles are the numerical results with both forward- and back-scattering, and open circles include just forward-scattering. (a) Weakly heterogeneous (Dw=10°) linear rotation model. (b) Strongly heterogeneous (Dw=60°) linear rotation model. (c) Weakly heterogeneous 10-random-layer model (maximum Dw=30°). (d) Strongly heterogeneous 10-random-layer model (maximum Dw=120°).

In Fig. 10, we extend these calculations to constant-gradient Section 3 is valid. The deviation of these contours towards the models with w: =45° and a range of splitting and heterogeneity horizontal defines a region where the scattering is too strong strengths. The ordinate is taken to be 1/k, a quantity pro- for perturbation theory to apply but not so strong as to portional to the inverse of the heterogeneity gradient, which prohibit the estimation of the apparent splitting parameters. defines a vertical correlation length. The contours of apparent The boundary between these two scattering regimes, indicated splitting time on this plot can be used to delineate the three by the dashed line in Fig. 10, is given by a correlation length scattering regimes. The region with nearly vertical contours that increases exponentially with the anisotropy strength. As at large values of the correlation length corresponds to weak the correlation length of the vertical heterogeneity decreases scattering, where Dt˜#Dt and the perturbation theory of at constant Dt, the apparent splitting time decreases, at first

© 2000 RAS, GJI 141, 374–390 Shear wave splitting with variable anisotropy 385

−1° 100 0.5s 1.0s 1.5s 2.0s 2.5s weak scattering −3° 32

, (km/deg) −10° κ 10 −30° 3.2 small strong, coherent scattering anisotropy −100° 1 −300°

vertical correlation length, 1/ 0.3 strong, incoherent scattering −1000° 0.1 0 0.5 1 1.5 2 2.5 3 strength of anisotropy, ∆t=∆s⋅d (s)

Figure 10. Scattering diagram contour of splitting-time measurements Dt˜ for smoothly varying models. The horizontal axis shows the range of anisotropy in the models (Dt∞), and the vertical axis shows the range of vertical heterogeneity in the models (k is the rotation rate). Splitting measurements cannot be made when Dt∞<0.5 s or 1/k<1 km deg−1. slowly then rapidly. Below some critical value of the correlation The applicability of this procedure is likely to be limited by length (~50 km in this example, corresponding to k=1° km−1), difficulties in extracting apparent splitting parameters at higher the scattering becomes so strong that Dt˜ cannot be defined. frequencies. Above 0.1 Hz, observed shear waveforms become As this diagram makes clear, it is not possible to distinguish increasingly complex due to microseisms, crustal scattering, on the basis of low-frequency splitting observations the differ- and other sources of ‘noise’. In addition, split shear waves have ence between highly heterogeneous anisotropy (Dt and k large) distinct ‘holes’ in their amplitude spectra at frequencies with and weak anisotropy (Dt small). Analysis of horizontally integer multiples of 1/Dt (Silver & Chan 1991), which com- propagating surfaces waves (e.g. Jordan & Gaherty 1996) is plicate analysis at higher frequencies. As a result, most shear one way to distinguish between these two cases. wave splitting analyses in the literature utilize centre frequencies that fall within a relatively narrow frequency band of approxi- mately 0.05–0.2 Hz (e.g. Silver 1996; Fouch & Fischer 1996; 5DISCUSSION Wolfe & Solomon 1998). Our numerical experiments in the For the case of weak scattering, differences in the sensitivity weak-scattering regime indicate that, across this bandwidth, kernels as a function of frequency can, in principle, be exploited variations in Dt˜ and w˜ are generally less than 0.1 s and 5°, to invert frequency-dependent shear wave splitting measure- respectively (Figs 8a and 9a). These variations are smaller than ments for a picture of anisotropic variation with depth. To do the typical error estimates in observational studies, and thus so, one applies standard splitting analysis to broad-band they cannot resolve changes in anisotropy with depth. recordings in order to obtain an apparent fast direction to In the case of strong scattering, the approximations made be used as a starting estimate for w:. Frequency-dependent in deriving the kernels are no longer valid, so that the kernels apparent splitting parameters are then extracted by applying cannot be used to solve the inverse problem. We can, how- the back-projection procedure to narrow-band filtered seismo- ever, utilize the numerical results in the interpretation of grams, and their kernels constructed from (10) and (11). splitting observations. Marson-Pidgeon & Savage (1997) report Adherence to the weak-scattering regime can be checked by frequency-dependent shear wave splitting results from New confirming that minimal signal remains on the tangential- Zealand (between 50 and 200 mHz) that are consistent with component seismogram after back-projection via the apparent our numerical results for the strong coherent scattering regime splitting parameters (Fig. 4). The frequency-dependent splitting (Figs 9b and d), implying significant vertical heterogeneity parameters can then be inverted for w(z). (In all of our with depth. In addition, splitting results from cratons in South calculations, the difference between the speeds of the two Africa (Gao et al. 1998), Australia (Clitheroe & van der Hilst eigenwaves remained constant throughout the model. In the 1998; O¨ zalaybey & Chen 1999), India (Chen & O¨ zalaybey real world, this parameter, like the anisotropy orientation, 1998), and Tanzania (Hill et al. 1996) all find splitting times probably varies with depth. It is a simple matter, however, to that are smaller (generally <0.6 s) than for many other con- extend the theory to depth-dependent wave speeds.) tinental environments (e.g. Silver 1996). Such observations are

© 2000 RAS, GJI 141, 374–390 386 R. L . Saltzer, J. B. Gaherty and T . H. Jordan typically interpreted as evidence for little or no anisotropy. of the figures were generated using GMT software freely Our calculations provide an alternative explanation for these distributed by Wessel & Smith (1991). This research was null results in terms of strong incoherent scattering in an upper funded by NSF Grant EAR-9526702. mantle that is anisotropic, but has a high degree of vertical heterogeneity. Other data, such as horizontally propagating surface waves, are necessary to distinguish between these two possibilities. REFERENCES In at least two cratonic regions where null or near-null Ando, M., 1984. ScS polarization anisotropy around the Pacific Ocean, results are reported (Australia and Southern Africa), analyses J. Phys. Earth, 32, 179–195. using surface waves find that the upper mantle is anisotropic Ando, M. & Ishikawa, Y., 1982. Observations of shear-wave velocity between the Moho and 200–250 km depth, with Dvs/vs of polarization anisotropy beneath Honshu, Japan: two masses with approximately 3–4 per cent (Gaherty & Jordan 1995; Saltzer different polarizations in the upper mantle, J. Phys. Earth, 30, et al. 1998). If perfectly aligned, such anisotropy would produce 191–199. over 2 s of splitting. We interpret the apparent discrepancy Chen, W.-P. & O¨ zalaybey, S., 1998. Correlation between seismic between the splitting and surface wave results as evidence anisotropy and Bouguer gravity anomalies in Tibet and its for strongly heterogeneous anisotropy in these regions. In implications for lithospheric structures, Geophys. J. Int., 135, 93–101. Clitheroe, G. & van der Hilst, R., 1998. Complex anisotropy in the particular, both sets of observations can be explained by a Australian lithosphere from shear-wave splitting in broad-band SKS model in which the local anisotropy axis remains, on average, records, in Structure and Evolution of the Australian Continent, eds close to horizontal, but varies in azimuth as a function Braun, J., Dooley, J., Goleby, R., van der Hilst, R. & Klootwijk, C., of position, with a vertical correlation length of the order of Am. geophys. Un. Geodynamic Series, 26, 39–57. 50 km (Jordan et al. 1999). Forsyth, D.W., 1975. The early structural evolution and anisotropy of the oceanic upper-mantle, Geophys. J. R. astr. Soc., 43, 103–162. Fouch, M.J. & Fischer, K.M., 1996. Mantle anisotropy beneath 6 CONCLUSIONS northwest Pacific subduction zones, J. geophys. Res., 101, 15 987–16 002. In both weakly and strongly heterogeneous media we find that Gaherty, J.B. & Jordan, T.H., 1995. The Lehman discontinuity as the shear wave splitting measurements made in the low-frequency base of an anistropic layer beneath the continents, Science, 268, bands typically used to make observations are more sensitive 1468–1471. to the upper portions of the model than to the lower portions. Gao, S.S., Silver, P.G. & James, D.E. & the Kaapvaal Working Group, Consequently, if the orientation of the anisotropy is hetero- 1998. Seismic structure and tectonics of southern Africa—progress geneous, the measured splitting direction will be more reflective report, EOS, T rans. Am. geophys. Un., 79, 45. of the fast-axis direction near the top of the upper mantle, Hill, A.A., Owens, T.J., Ritsema, J., Nyblade, A.A. & Langston, C.A., and the measured splitting time will vary as a function of 1996. Constraints on the upper mantle structure beneath the backazimuth. This effect may explain why global shear wave Tanzanian Craton from teleseismic body waves, EOS, T rans., Am. splitting measurements tend to correspond to the local tectonic geophys. Un., 77, F477. fabric in crust beneath the observing station (Silver 1996). Jordan, T.H. & Gaherty, J.B., 1996. Stochastic modeling of small- scale, anisotropic structure in the continental upper mantle, in Proc. We have derived analytic expressions for sensitivity kernels 17th Ann. seismic Res. Symp., pp. 433–451, eds Lewkowicz, J.F., that relate a perturbation in the measured splitting parameters McPhetre, J.M. & Reide, D.T., Phillips Labs, Hanscom Air Force to a perturbation in the anisotropy of the model. For weakly Base. heterogeneous media, frequency-dependent shear wave splitting Jordan, T.H., Saltzer, R.L. & Gaherty, J.B., 1999. Small-scale aniso- measurements can be inverted using these kernels to determine tropic heterogeneity in the continental upper mantle, Seism. Res. how the anisotropy varies as a function of depth. Variations L ett., 70, 259. in the strength of the anisotropy as a function of depth can be Keith, C.M. & Crampin, S., 1977. Seismic body waves in anisotropic accounted for with a weighting function. Practically speaking, media: synthetic seismograms, Geophys. J. R. astr. Soc., 49, 225–243. however, a sufficient number of measurements may not be Kennett, B.L.N., 1983. Propagation in Stratified Media, available at high enough frequencies or with enough precision Cambridge University Press, Cambridge. (errors are typically ±10° or more) for such an inversion to Mallick, S. & Frazer, N.L., 1990. Computation of synthetic seismo- be feasible. grams for stratified, azimuthally anisotropic media, J. geophys. Res., Strongly heterogeneous media (i.e. where the fast axis 95, 8513–8526. Marson-Pidgeon, K. & Savage, M.K., 1997. Frequency-dependent direction varies anywhere between 0° and 180°) have the anisotropy in Wellington, New Zealand, Geophys. Res. L ett., 24, additional property that they cause strong scattering. This 3297–3300. scattering will cause the tangential-component seismograms to Montagner, J.P. & Tanimoto, T., 1990. Global anisotropy in the have very little energy, and a null-like energy measurement upper mantle inferred from the regionalization of phase velocities, will be made despite the fact that the medium may be highly J. geophys. Res., 95, 4797–4819. anisotropic. Our numerical experiments show that the greatest Nataf, H.C., Nakanishi, I. & Anderson, D.L., 1984. Anisotropy and diagnostic of strong vertical heterogeneity is the drop-off in Dt˜ shear-velocity heterogeneities in the upper mantle, Geophys. Res. relative to Dt. L ett., 11, 109–112. O¨ zalaybey, S. & Chen, W.-P., 1999. Frequency-dependent analysis of SKS/SKKS waveforms observed in Australia: Evidence for null ACKNOWLEDGMENTS , Phys. Earth planet. Inter., 114, 197–210. Raitt, R.W., Shor, G.G., Francis, T.J.G. & Morris, G.B., 1969. Many thanks to Martha Savage, Steve Ward and an anonymous Anisotropy of the Pacific upper mantle, J. geophys. Res., 74, reviewer for useful comments that improved the paper. Many 3095–3109.

© 2000 RAS, GJI 141, 374–390 Shear wave splitting with variable anisotropy 387

Ru¨mpker, G. & Silver, P.G., 1998. Apparent shear-wave splitting where * indicates complex conjugation and parameters in the presence of vertically varying anisotropy, Geophys. J. Int., 135, 790–800. − exp[ik1(z z0)] 0 Saltzer, R.L., Jordan, T.H., Gaherty, J.B., Zhao, L. & the Kaapvaal E(z, z )= . (A5) 0 C − D Working Group, 1998. Anisotropic structure of the Kaapvaal craton 0 exp[ik2(z z0)] from surface wave analysis, EOS, T rans. Am. geophys. Un., 79, 45. Savage, M.K., 1999. Seismic anisotropy and mantle deformation: what Therefore, the displacement–stress propagator in the eigenwave have we learned from shear wave splitting?, Rev. Geophys., 37, coordinate system is ˆ h(z, z )= (z, z ) −1, and the general 65–106. P 0 DE 0 D homogeneous-layer propagator matrix can be written as Shearer, P.M. & Orcutt, J.A., 1986. Compressional and shear wave anisotropy in the oceanic lithosphere—the Ngendei seismic h h refraction experiment, Geophys. J. R. astr. Soc., 87, 967–1003. Puu(z, z0) Put(z, z0) Ph(z, z )¬ =U(w)Pˆ h(z, z )UT(w). Silver, P.G., 1996. Seismic anisotropy beneath the continents: probing 0 C h h D 0 the depths of geology, Ann. Rev. Earth. planet. Sci., 24, 385–432. Ptu(z, z0) Ptt(z, z0) Silver, P.G. & Chan, W.W., 1991. Shear wave splitting and mantle (A6) deformation, J. geophys. Res., 96, 16 429–16 454. Silver, P.G. & Savage, M.K., 1994. The interpretation of shear-wave The propagator for the anisotropic, heterogeneous interval splitting parameters in the presence of two anisotropic layers, ≤ ≤ Geophys. J. Int., 119, 949–963. 0 z d can be constructed by approximating the hetero- Tanimoto, T. & Anderson, D.L., 1985. Lateral heterogeneity and geneity as a stack of homogeneous layers and multiplying azimuthal anisotropy of the upper mantle: Love and Rayleigh waves propagators in the form of (A6). Better analytical insight into 100–250 S, J. geophys. Res., 90, 1842–1858. the effects of the heterogeneity can be obtained by proceeding Varshalovich, D.A. & Moskalev, A.N., 1988. Quantum T heory of in a different fashion, however. We make the ansatz (cf. Kennett Angular Momentum, ed. Kheronski, V.K., World Scientific. 1983, p. 53), Vidale, J.E., 1986. Complex plarzation analysis of particle motin, Bull. seism. Soc. Am., 76, 1393–1405. P(z, z∞)=U(w(z))DE(z, z )Q(z, z )D−1UT(w(z )), (A7) Wessel, P. & Smith, W.H.F., 1991. Free software helps map and display 0 0 0 data, EOS, T rans. Am. geophys. Un., 72, 441–446. and find that (A7) obeys the system equation ∂ = if Wolfe, C.J. & Solomon, S.C., 1998. Shear-wave splitting and zP AP implications for mantle flow beneath the MELT region of the East and only if Q satisfies Pacific Rise, Science, 280, 1230–1232. ∂ = zQ(z, z0) S(z)Q(z, z0), (A8a) = Q(z0,z0) I , (A8b) APPENDIX A: PROPAGATOR MATRICES AND SPLITTING OPERATORS where The system matrix A(z) is rotated into the eigenwave coordinate system by the 4×4 block-diagonal matrix U=diag[U, U], 0 a(z) 0 b*(z) −a*(z) 0 b*(z) 0 dw 0C−1 S(z)= . (A9) Aˆ =UAUT= , (A1) 0 b(z) 0 a*(z) dz C−v2I0D C D b(z) 0 −a(z) 0 and is diagonalized by a matrix whose columns represent the downgoing and upgoing eigenwaves: D−1Aˆ D= The depth-dependent coefficients a and b depend exponentially − − = = = − i diag[k1, k2, k1, k2]. Here, kj v/vj ( j 1, 2) are the on the wavenumber Dk k2 k1 and the wavenumber average : = + eigenwavenumbers, and k (k1 k2)/2, respectively:

−1 −1 v: Du Du Du Dt a(z)= eiDkz , (A10a) D= , D−1= , (A2) √ C *D C −1 *−1D v1v2 Dt Dt Du Dt Dv where the 2×2 blocks are = 2ik:z b(z) √ e . (A10b) 2 v1v2 e 0 ivv e 0 = 1 = 1 1 The propagator Q(z, z )isa4×4differential scattering Du C D , Dt C D . (A3) 0 0 e2 0 ivv2e2 matrix for the eigenwaves, which we write in block form as

Kennett’s (1983, eq. 2.63) energy normalization procedure Q++ Q+− = −1/2 = . (A11) yields ej (2vj) . Q C D In a homogeneous layer [w(z)=constant], the eigenwave Q−+ Q−− propagator from z0 to z is The submatrices Q++ and Q+− describe, respectively, the forward scattering of downgoing (+) and upgoing (−) eigen- E(z, z ) 0 = 0 waves by gradients in w(z), and Q and Q describe the E(z, z0) C D , (A4) −+ +− 0E*(z, z0) corresponding backward scattering. These scattering operators

© 2000 RAS, GJI 141, 374–390 388 R. L . Saltzer, J. B. Gaherty and T . H. Jordan satisfy the reciprocal relations In the case of a homogeneous layer, these expressions reduce ˆ h = ˆ h = − ˆ h = −1 = * to Puu Ptt diag[cos kj(z z0)], Put diag[(vvj) sin kjd], Q−− Q++ , (A12a) ˆ h = − and Ptu diag[ (vvj) sin kjd]. = * Q+− Q−+ . (A12b) A useful approximation, almost always employed in vertical shear wave splitting analysis, is to ignore back-scattering and From eqs (A8) to (A10) we can see that the strength of forward reverberations within the anisotropic layer. This amounts to scattering in a depth increment dz is proportional to ignoring terms of order Dv/v: in eq. (A17). Under this approxi- a exp(iDkz)dw(z), while the strength of the back scattering 0 mation, a =1 and b =0, so that Q =Q =0. A little goes like b exp(2ik:z)dw(z). If the relative difference in the 0 0 +− −+ 0 algebra obtains eigenvelocities is small, then forward scattering will tend to dominate because b %a ~1. Moreover, in this situation of 0 0 = T 2 T : R UdQ++E Q++Ud , (A18a) small anisotropy, Dk%2k, so that the back-scattering kernel (no back-scattering). will be more oscillatory and its integral contributions will tend = T ∞ H u(0) 2U0EQ++Ud uI(d) . (A18b) to cancel if w(z) is smooth. = At zero frequency, S S0 depends on z only through Eq. (A18b) shows that, when back-scattering can be ignored, ˙ ¬ w dw/dz. Therefore, S0 commutes with its integral, and the the eigenwave propagator is just EQ It will be convenient = = − ++ solution to (11) is Q(z, z0) exp(SDw), where Dw w(z) w(z0). to pull out the phase factor corresponding to the mean Using the fact that traveltime through the layer, t:=k:d/v, and rewrite these 2n= − n 2n 2n+1= − n 2n expressions in terms of the eigenwave splitting matrix, S0 ( 1) Dw I and S0 ( 1) Dw S0 , we can sum the exponential series. This yields a good approxi- exp(−ivDt/2) 0 mation to the propagator across a layer that is thin compared H=exp(−ivt:)E= C D , (A19) with a wavelength; that is, for k:Dz%1, 0 exp(ivDt/2)

cos Dw a(z) sin Dw which is unimodular; that is, det[H]=1. We define the + Q++(z Dz, z)# , (A13a) splitting operator, C−a*(z) sin Dw cos Dw D

0 b(z) sin Dw C=U HQ UT . (A20) + 0 ++ d Q−+(z Dz, z)# C D . (A13b) b(z) sin Dw 0 = : ∞ The surface displacement is thus u(0) 2 exp(ivt)CuI(d), and We note that this approximation is independent of the form the reflection matrix is R=exp(2iv:t)CTC. The factor of two of w(z) and, for example, does not require w(z) to be a smooth in the former comes from the constructive interference of the function of depth. Indeed, it provides the generalization of upward-going wave and its surface reflection. In the case of a ˙ = the propagator to self-affine (fractal) media for which w may homogeneous layer, Q++ I, and (A20) becomes not be well defined. Eqs (A7) and (A13) are the basis of our = T computational algorithm. Ch(w, Dt) U(w)H(Dt)U (w)(homogeneous layer). (A21) ~ − : For an upgoing wave uI(z) exp( ikz) incident at the base of the anisotropic layer, the displacement–stress vector in the All of the matrices in (A20) are both unitary and unimodular; half-space can be expressed as for example, C−1=C†¬(C*)T, det[C]=1. (The unimodularity = of Q++ follows from tr[S++(z)] 0; see Kennett 1983, p. 42.) (I+R) ∞ = Therefore, all of the matrix operations associated with forward- f (d) uI(d) , (A14) Civv:(I−R)D scattering belong to the group SU(2). This symmetry can be used to simplify the analysis. Any member of this group can where R is a 2×2 matrix of reflection coefficients. Satisfying be written in it terms of two complex numbers, the zero-traction boundary conditions at the surface yields R=U (ivv:Pˆ −Pˆ )−1(ivv:Pˆ +Pˆ )UT . (A15) ab d tt tu tt tu d , where |a|2+|b|2=1. (A22) = C−b* a*D From here on, Uz U(w(z)) and it is understood that, unless otherwise specified, the propagators are taken from the base of the anisotropic layer to the surface; for An SU(2) matrix thus depends on three real parameters and ˆ ¬ T can be written in the following general forms (Varshalovich & example, Puu U0 Puu(0, d)Ud. In this notation, the free-surface displacement vector is Moskalev 1988): u(0)=U [Pˆ +ivv:Pˆ +(Pˆ −ivv:Pˆ )Rˆ ]UTu∞ (d) . (A16) 0 uu ut uu ut d I exp[−i(a+c)/2] cos b/2 exp[i(a+c)/2] sin b/2 = The symmetries in (A2), (A4) and (A12) can be used to C Cexp[−i(a+c)/2] sin b/2 exp[i(a+c)/2] cos b/2D express the propagator submatrices as the following (real-valued) expressions: cos V/2−i cos H sin V/2 −i exp(iW) sin H sin V/2 = . ˆ = + −1 C− − + D Puu DuRe[E(Q++ Q+−)]Du , (A17a) i exp( iW) sin H sin V/2 cos V/1 i cos H sin V/2 ˆ = + −1 Put iDuIm [E(Q++ Q+−)]Dt , (A17b) (A23) Pˆ =iD [E(Q +Q )]D−1 , (A17c) tu tIm ++ +− u SU(2) is homomorphic (with a two-fold ambiguity) to O+(3), ˆ = + −1 Ptt DtRe[E(Q++ Q+−)]Dt . (A17d) the group of proper orthogonal transformations in 3-space,

© 2000 RAS, GJI 141, 374–390 Shear wave splitting with variable anisotropy 389 which allows the splitting operations to be visualized as 3-D Since the perturbed model is homogeneous above z and rotations. In the first form in (A23), the parameters (a, b, c) below z+dz, its splitting operator can be written in the form correspond to the Euler angles of the 3-D rotation; in the = T second, the 3-D rotation is through an angle V about an axis C U(w0)H(Dt)Q++U (w0), (B3) with polar coordinates (H, W). Eq. (A21) can be recast as where H is given by (A19). For a constant perturbation dw in + = = cos vDt/2−i cos 2w sin vDt/2 −i sin 2w sin vDt/2 a layer (z, z dz), Q satisfies (A8) with a0 1, b0 0, and C = . h C − + D i sin 2w sin vDt/2 cos vDt/2 i cos 2w sin vDt/2 w˙(f)=dw[d(f−z)−d(f−z−dz)] . (B4) (A24) Integrating up from the base of the layer across the discon- The splitting matrix for a homogeneous layer thus corresponds tinuities yields an expression for Q++ that is the product of to a 3-D rotation through an angle vDt about an axis located two matrices in the form of (A13a), one for an azimuthal + − at colatitude 2w and zero longitude change of dw at z dz, and one for a change of dw at z. Multiplying these out and using the small-angle approxi- mations, we can express the forward-scattering matrix in terms of a perturbation parameter dX¬Dk dz dw exp[iDk(d−z)]: APPENDIX B: FRE´ CHET KERNELS FOR A HOMOGENEOUS STARTING MODEL 1 −idX = Q++ . (B5) To calculate the Fre´chet kernels defined by (8) and (9), we C−idX* 1 D perturb the splitting orientation function w(z) by a small constant amount dw in a thin layer of thickness dz at a depth The apparent splitting parameters minimize the energy on the < < ˜ 0 z d and compute the corresponding perturbations dw(z) transverse component of the back-projected displacement field, and dt˜(z). The kernels Gw(z) and Gt(z) are then given as given by the quadratic form (7). In the present notation, this ˜ the limiting values of the ratios dw(z)/dwdz and dt˜(z)/dwdz, integral becomes respectively. This calculation can be done numerically for arbitrary starting models and pulse shapes (e.g. Fig. 7). Here 2 2 ∞ ∞ = | T −1 ∞ ∞ |2| |2 we present an analytical derivation for narrow-band pulses in e (w , Dt ) 2 P yˆ Ch (w , Dt )Cxˆ uI(v) dv . (B6) the special case of a homogeneous starting model. 0 Because the perturbations are small, the forward-scattering The homogeneous-layer splitting operator in this expression approximation applies. Vertical propagation through a homo- corresponds to the perturbed apparent splitting parameters geneous anisotropic layer with a fast-axis orientation w 0 w∞=w +dw∞ and Dt∞=Dt+dt∞, which can be expressed in a and splitting time Dt yields the Fourier-transformed vertical 0 = : form similar to (B3): displacement u(0, v) 2 exp(ivt )Ch(w0, Dt)uI(v), where the homogeneous-layer splitting operator C is given by eq. (A24). h (w∞, t∞)=U(w )H( t)Q∞ UT(w ). (B7) We assume that the incident pulse is radially polarized, Ch D 0 D ++ 0 u (v)=u (v)xˆ =[u (v)0]T, and we approximate its energy I I I Equating (B6) with (w∞, t)=U(w∞)H( t∞)UT(w∞) yields the spectrum by a Gaussian: Ch D D scattering matrix 1 |u (v)|2= exp[−(v−v )2/2s2] c−is cos 2dw∞−is sin 2dw∞ I s√8p 0 ∞ = Q++ , (B8) C −is* sin 2dw∞ c*+is* cos 2dw∞D 1 + exp[−( + )2/2 2]. (B1) √ v v0 s s 8p with complex coefficients,

This spectrum has peaks of half-bandwidth s centred at c=exp(ivDt/2) cos[v(Dt+dt∞)/2] , (B9a) ± frequencies of v0, and it is normalized to have unit total ∆2 | |2 = energy: −2 uI(v) dv 1. If the pulse is narrow band in the s=exp(ivDt/2) sin[v(Dt+dt∞)/2] . (B9b) sense that s%v0, then the integral of its energy spectrum against any reasonably smooth function can be approximated Eq. (B8) is exact and does not require the perturbations by integrating a truncated Taylor expansion of the function ∞ ∞ = = dw and dt to be small. When, w0 Dt 0, for example, about the centre frequency v : = ∞ = ∞ ∞ ∞ 0 c cos vdt /2, s sin vdt /2, and (B7) reduces to Ch(dw , dt ) in the form given by (A24). 2 From (B3) and (B6) and the fact that Q∞ is unitary, we 2 f(v)|u (v)|2dv ++ P I obtain −1( ∞, t∞) = ( ) ∞† T( ). The energy 0 Ch dw d C U w0 Q++Q++U w0 (B6) involves an integration over the (2,1) component of this 2 + ˙ + 2 ¨ | |2 matrix. The product Q∞† Q can be expressed in the general #2 P [f(v0) v f (v0) v f (v0)] uI(v) dv ++ ++ 0 SU(2) form (A22), where # f(v )+s2 ¨f (v ). (B2) 0 0 a=c*+s*(dX* sin 2dw∞+i cos 2dw∞), (B10) 4 =− + ∞+ ∞ The terms dropped in this approximation are of order (s/v0) . b idXc* s*(dX cos 2dw i sin 2dw ).

© 2000 RAS, GJI 141, 374–390 390 R. L . Saltzer, J. B. Gaherty and T . H. Jordan

Working out the appropriate matrix element in terms of the 2 = | |2 real and imaginary parts of these coefficients, we find A21 sin 2w0 cos 2w0 P (v/v0) sin vDt uI(v) dv , (B13c) 0 2 2 2 ∞ ∞ = 2+ + 2 = − 2| |2 e (dw , dt ) 2 P {Re(b) [Im (b) cos 2w0 Re(a) sin 2w0] } A22 P (1 cos vDt) uI(v) dv 0 0 ×|u ( )|2d . (B11) 2 I v v + 2 2 | |2 cos 2w0 P sin vDt uI(v) dv , (B13d) To find the energy minimum, we differentiate (B11) with 0 respect to the perturbations dw∞ and dt∞ and set the results 2 D (z)=sin 2w cos 2w (v/v ) cos k(d−z)|u (v)|2dv , equal to zero, which gives two equations for the apparent 1 0 0 P 0 D I 0 splitting parameters. Linearizing these equations in dw˜ and dt˜, we obtain a 2×2 system for the Fre´chet kernels: (B14a) 2 D (z)= (1−cos v t) sin k(d−z)|u (v)|2dv A11 A12 Gw(z) D1(z) 2 P D D I C DC D =Dk C D , (B12) 0 A21 A22 Gt(z) D2(z) 2 + 2 − ∞| |2 2 cos 2w0 P (v/v0) sin Dk(d z) cos vDt uI(v) dv . = 2 | |2 A11 sin 2w0 P (v/v0) uI(v) dv , (B13a) 0 0 (B14b) 2 = | |2 Approximating these integrals using (B2) and solving for the A12 sin 2w0 cos 2w0 P sin vDt uI(v) dv , (B13b) 0 kernels leads to the expressions (10) to (14) given in the text.

© 2000 RAS, GJI 141, 374–390