11 Global Imaging of the Earth’s Deep Interior: Seismic ...

27
11 Global Imaging of the Earth’s Deep Interior: Seismic Constraints on (An)isotropy, Density and Attenuation JEANNOT TRAMPERT AND ANDREAS FICHTNER Department of Earth Sciences, Utrecht University, Utrecht, The Netherlands Summary Seismic tomography is our principal tool to probe the deep interior of the Earth. Over the past decades, it has drawn the picture of a vigor- ously convecting planet, with large-scale up- and down-wellings convincingly imaged by isotropic velocity variations. Models of seismic anisotropy induced by crystal alignment provide insight into the underlying convective motion, and variations of density allow us to discriminate between ther- mal and compositional heterogeneities. However, despite substantial progress, enor- mous challenges remain: The strength of im- aged anisotropy trades off nearly perfectly with the roughness of isotropic heterogeneity, which complicates any quantitative interpretation. Den- sity variations in the Earth, except for spherical harmonic degrees 2, 4 and 6, are still largely unknown, and error estimates are dominated by subjective regularization. Concerning attenua- tion, any pair of global 3D models is uncorrelated, with mutual differences larger than the attributed uncertainties. Most of these difficulties can be traced to the complicated nature of multi-observable/multi- parameter inverse problems that suffer from weak constraints and complex trade-offs. Promising di- rections to strengthen constraints and reduce trade-offs, include (1) the addition of new data at both the short-period (scattered waves) and Physics and Chemistry of the Deep Earth, First Edition. Edited by Shun-ichiro Karato. © 2013 John Wiley & Sons, Ltd. Published 2013 by John Wiley & Sons, Ltd. long-period (mode splitting) ends of the spectrum, (2) the implementation of fully probabilistic ap- proaches, (3) the joint inversion of phase and amplitude information for elastic and anelastic structure, (4) the design of misfit functionals tar- geted at specific aspects of Earth structure, and (5) the continuing development of full waveform in- version techniques. Furthermore, much progress can be made by objectively quantifying the uncer- tainties of our inferences. This can lead to both more consistent models of the Earth and more reliable interpretations in terms of its thermo- chemical structure and evolution. 11.1 Introduction Seismic tomography is central to the multi- disciplinary effort aimed at understanding the structure and thermo-chemical evolution of the Earth. Indeed, as seismic waves produced by earthquakes travel through the deep Earth, they absorb information from the (an)elastic and density structure along their way which can be extracted by seismic imaging. Compared to geological time scales, seismic waves record an instantaneous snapshot of the current Earth, although anisotropy provides some constraints on geological events in the Earth’s recent history. The process of absorbing the information is a well posed forward problem and currently, with

Transcript of 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

Page 1: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

11 Global Imaging of the Earth’sDeep Interior: Seismic Constraints

on (An)isotropy, Densityand Attenuation

J E A N N O T T R A M P E RT A N D A N D R E A S F I C H T N E RDepartment of Earth Sciences, Utrecht University, Utrecht, The Netherlands

Summary

Seismic tomography is our principal tool to probethe deep interior of the Earth. Over the pastdecades, it has drawn the picture of a vigor-ously convecting planet, with large-scale up- anddown-wellings convincingly imaged by isotropicvelocity variations. Models of seismic anisotropyinduced by crystal alignment provide insight intothe underlying convective motion, and variationsof density allow us to discriminate between ther-mal and compositional heterogeneities.

However, despite substantial progress, enor-mous challenges remain: The strength of im-aged anisotropy trades off nearly perfectly withthe roughness of isotropic heterogeneity, whichcomplicates any quantitative interpretation. Den-sity variations in the Earth, except for sphericalharmonic degrees 2, 4 and 6, are still largelyunknown, and error estimates are dominatedby subjective regularization. Concerning attenua-tion, any pair of global 3D models is uncorrelated,with mutual differences larger than the attributeduncertainties.

Most of these difficulties can be traced to thecomplicated nature of multi-observable/multi-parameter inverse problems that suffer from weakconstraints and complex trade-offs. Promising di-rections to strengthen constraints and reducetrade-offs, include (1) the addition of new dataat both the short-period (scattered waves) and

Physics and Chemistry of the Deep Earth, First Edition. Edited by Shun-ichiro Karato.© 2013 John Wiley & Sons, Ltd. Published 2013 by John Wiley & Sons, Ltd.

long-period (mode splitting) ends of the spectrum,(2) the implementation of fully probabilistic ap-proaches, (3) the joint inversion of phase andamplitude information for elastic and anelasticstructure, (4) the design of misfit functionals tar-geted at specific aspects of Earth structure, and (5)the continuing development of full waveform in-version techniques. Furthermore, much progresscan be made by objectively quantifying the uncer-tainties of our inferences. This can lead to bothmore consistent models of the Earth and morereliable interpretations in terms of its thermo-chemical structure and evolution.

11.1 Introduction

Seismic tomography is central to the multi-disciplinary effort aimed at understanding thestructure and thermo-chemical evolution ofthe Earth. Indeed, as seismic waves producedby earthquakes travel through the deep Earth,they absorb information from the (an)elastic anddensity structure along their way which canbe extracted by seismic imaging. Compared togeological time scales, seismic waves record aninstantaneous snapshot of the current Earth,although anisotropy provides some constraintson geological events in the Earth’s recent history.The process of absorbing the information is awell posed forward problem and currently, with

Page 2: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

Global Imaging of the Earth’s Deep Interior 325

the help of large computers, state-of-the-arttechniques allow us to model quasi-exact wavepropagation in quite complex Earth models. Seis-mic tomography is the corresponding ill-posedinverse problem and still quite a challenge. Themain limiting factors are the uneven illuminationof the Earth’s interior by seismic waves and thecomputational resources. The former is addressedby using more of the available data, althoughthe nonuniform distribution of earthquakes andseismic stations cannot entirely be undone. Thecomputational burden is reduced by implement-ing various theoretical approximations duringimaging.

The most common approach to seismic tomog-raphy is based on travel times of identified bodyand surface waves, which are inverted for ve-locities using approximations of elastic isotropyand high frequency wave propagation (ray theory).Recent reviews provide extensive reference listsgoing back to the beginning of this research fieldin the 1970s (e.g., Romanowicz, 2003; Trampert &van der Hilst, 2005; Nolet, 2008; Rawlinson et al.,2010). Sometimes, a collection of normal-modefrequencies is added to provide a global long-wavelength constraint on the structure (Ritsemaet al., 2011). While these studies have shaped aclear picture of the Earth’s internal structure, theyhave not provided unambiguous information ontheir origin and evolution. Trampert & van derHilst 2005 argued that estimates of uncertaintyand information on the density structure are es-sential to distinguish between temperature andchemical origins. A convincing case can furtherbe made that information on intrinsic attenuationand anisotropic parameters significantly helps ourquest to understand the thermo-chemical evo-lution of the Earth (e.g., Karato, 2008). Mantlediscontinuities also provide interesting informa-tion and is discussed in chapter 10 by Deusset al.

The purpose of this chapter is to review onwhich class of models there is agreement andwhere further investigation is required. The mainemphasis will be on the imaging itself and notso much its thermo-chemical interpretation,although we will provide some indications. In

turn we will treat the case of isotropic velocitytomography, anisotropy, density and attenuationimaging in decreasing order of consensus. Finally,we will make some suggestions for future work.

11.2 An Introduction to LinearisedInverse Theory

Assume that we have a collection of measure-ments gathered in a data vector d. First, we haveto decide on a mathematical description of the pa-rameter field we want to image. Often the modelparameters are expanded onto a finite number ofglobal or local basis functions, e.g. spherical har-monics, blocks and many others. The coefficientsfor these basis functions are called the modelparameters and are collected in a vector m. Math-ematically, d and m belong to vector spaces, thedata space D and the model space M, respectively.The mapping from M to D is called the forwardoperator. It is, in the case of a linear problem,represented by the matrix G. For instance, if weuse ray theory, the data could be travel time resid-uals, the model parameters constant slownessesin blocks, and a row of G contains the lengths ofa particular ray in each block. Formally, we maywrite

d = Gm + e. (11.1)

The data are not perfect and therefore we addedan error vector e which is implicitly recordedtogether with the data d. The inverse problemconsists of finding a linear mapping S from Dto M. Since G is usually not invertible, S is notthe exact inverse of G, and thus we only find anestimator m of m. This is formally written as

m = m0 + S(d − Gm0), (11.2)

where we have introduced an optional startingmodel m0. The interesting question is how ourestimated model m is related to the true modelm that we intended to find. We therefore insertEquation (11.1) into Equation (11.2) to get

m − m0 = R(m − m0) + Se, (11.3)

Page 3: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

326 jeannot trampert and andreas fichtner

where we have defined the resolution operator asR = SG. It is more instructive to introduce thetarget model m explicitly, and we find that

m − m = (I − R)(m0 − m) + Se, (11.4)

with the identity matrix I. The last equationclearly shows that our estimated model m devi-ates from the target model m by two terms. Thefirst one is due to imperfect resolution, meaningthat the resolution matrix is not equal to theidentity matrix (R �= I). The second term is the re-sult of data errors propagating into the estimatedsolution.

There is considerable freedom and choice in-volved in the construction of the approximateinverse operator S (e.g., Parker, 1994; Tarantola,2005). The most general approach starts by assign-ing a probability density σM(m) to each model, i.e.

σM(m) = kρM(m)L(m) , (11.5)

where ρM is the prior distribution in the modelspace, L the likelihood function which measureshow well the model explains the data within theiruncertainty, and k a normalizing constant (e.g.,Tarantola, 2005). Assuming that both, data uncer-tainty and our prior knowledge, can adequatelybe described by Gaussian distributions, Equation(11.5) takes the form

σM(m) = k e− 12 χ (m), (11.6)

with the misfit functional

χ (m) = (d − Gm)TC−1d (d − Gm)

+ (m − m0)TC−1m (m − m0) . (11.7)

The superscript T denotes vector transposition,and Cd and Cm are the covariance matrices inthe data and model space, respectively. On thebasis of Equations (11.6) and (11.7) we define ourestimator m as the maximum-likelihood model,i.e. the model that maximizes (11.6) and mini-mizes (11.7). Requiring that the derivative of χ

with respect to m vanishes at the position of the

maximum-likelihood model m, we find that S isdetermined by

S =(

GTG + σ 2d

σ 2m

I

)−1

GT , (11.8)

where we assumed, for simplicity, that the co-variance matrices are diagonal, i.e. Cd = σ 2

d I andCm = σ 2

mI. The symbol σd denotes the standarddeviation of the data uncertainty, and σm is thestandard deviation of the prior model range. Theposterior model covariance is then simply

Cm = σ 2d

(GTG + σ 2

d

σ 2m

I

)−1

, (11.9)

where we note that such explicit expressions canonly be obtained on the basis of Gaussian statis-tics. Equation (11.8) reveals a dilemma in thesolution of inverse problems: For most realis-

tic applications, the matrix (GTG + σ2d

σ2m

I) is badly

conditioned or not invertible at all, unless the

ratioσ2

dσ2

mis artificially increased. In this case,

the initial variances are used to regularize theinversion – and not to objectively quantify dataerrors and prior knowledge, as it was originallyintended. Decreasing σm for the purpose of regu-larization also reduces the posterior covariance,therefore providing an unrealistically optimisticestimate of the errors in our inferred model m.Expression (11.8) nicely reveals the effect that reg-ularization has on the estimated model m. A largedata uncertainty (σd large) and a narrow searcharound the prior model (σm small), result in asmall approximate inverse S, and hence little up-date of m0. This explains why most tomographicinversions recover only a fraction of the ampli-tudes of the actual heterogeneities. Furthermore,the regularization employed in the constructionof S reduces the resolution, because R = SG.Moreover, as seen from Equation (11.4), regular-ization acts as a trade-off between the error prop-agation and the imperfect resolution. For a strongregularization, Se is small and (I − R)(m0 − m) islarge and vice versa. The knowledge of both terms

jeannot
Highlight
jeannot
Sticky Note
remove 'tilde' above m
Page 4: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

Global Imaging of the Earth’s Deep Interior 327

is needed to assess the quality of m, unless one ofthem is dominant.

The only way to avoid possibly unrealisticGaussian distributions and the effects that reg-ularization has on our inferences, is to directlysample the probability distribution (11.5), using,for instance, Monte Carlo techniques (e.g. Sam-bridge & Mosegaard, 2002) or neural networks(e.g. Meier et al., 2007). These are, however, com-putationally much more involved than solvingthe regularized analytical expressions based onGaussian statistics.

Finally, we note that seismic tomography isin reality not a linear inverse problem becauseour data generally depend nonlinearly on theproperties of the Earth. Within a probabilisticframework, non-linearity can naturally be ac-commodated in the model space sampling. Indeterministic problems, the nonlinearity is mostoften addressed using a perturbation approach.This involves updating the model iteratively:

mn = mn−1 + Pn−1(d − Gn−1mn−1)

+ Q(mn−1 − m0) , (11.10)

where n indicates the iteration step and Gn−1the Frechet derivatives of the nonlinear forwardfunctional with respect to mn−1. Commonly usedalgorithms are, for instance, conjugate gradientor Newton schemes which each define separateoperators Pn−1 and Q. It is worth noting thatsolution (11.5) holds for linear as well as nonlinearproblems. In the following, we will refer to thesedifferent ways of solving the inverse problem inthe context of (an)isotropic and (an)elastic seismictomography.

11.3 Isotropic Velocity Tomography

Over the last three decades seismic tomographyhas produced a large number of models with ahigh degree of overlap as documented in manyreview articles (e.g., Woodhouse & Dziewonski,1989; Ritzwoller & Lavely, 1995; Masters et al.,2000; Romanowicz, 2003; Trampert & van derHilst, 2005; Rawlinson et al., 2010).

P-velocity models are almost exclusivelyconstructed from large collections of traveltime residuals. The uneven distribution ofearthquakes and seismic stations implies thatmainly tectonically active regions are sampledby these data, and the main features to be imagedare descending lithospheric plates (e.g., van derHilst et al., 1997). The depth extent of theimaged slabs varies, some reaching the lower-most mantle, others stagnant in the transitionzone (Fukao et al., 2009). If S-wave travel timeresiduals are used, remarkably similar imagesare retrieved (Grand et al., 1997). Travel timemodels are often referred to as high-resolutionmodels. This is misleading since the fine blockor pixel parametrization, usually employed inthe construction of these models, only indicatesa potential maximum resolution. The actualachieved resolution is frequently unknown,despite the many proposed synthetic tests.

S-velocity models are more often constructedfrom long-period waveforms and/or a combi-nation of long-period body wave travel timeresiduals, surface wave dispersion measurementsand normal-mode splitting functions (e.g. Megnin& Romanowicz, 2000; Kustowski et al., 2008;Ritsema et al., 2011). The main robust featuresappearing are low-velocity mid-oceanic ridgesrecognizable down to a depth of 100–150 km(Figure 11.1, left). There is a clear ocean-continentdifference disappearing at around 250 km depth.The transition zone and the lower-most mantleare dominated by large low-velocity zonesbeneath Africa and the Pacific Ocean and highervelocities in a circum-Pacific belt (Figure 11.1,middle and right). These features have beenknown since the pioneering studies of (Masterset al. 1982), (Woodhouse & Dziewonski 1984) and(Dziewonski 1984). They have since been con-firmed by virtually all successive studies. TheseS-velocity models employ data with a more evencoverage than high-frequency P-wave travel timeresiduals, and the parametrization emphasizesthe long-wavelength structure by employing low-order spherical harmonic expansions. Overall,the resolution is more even over the globe, butthat is not to say that it is known precisely.

Page 5: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

328 jeannot trampert and andreas fichtner

S40RTS at 100 km depth S40RTS at 500 km depth S40RTS at 2800 km depth

−7.5 +7.5 −3.0 +3.0 −3.0 +3.0

−1.5 +1.5 −0.6 +0.6 −0.6 +0.6

null-space componentat 100 km depth

null-space componentat 500 km depth

null-space componentat 2800 km depth

Fig. 11.1 Top Relative S velocity variations, d ln vs, in the global model S40RTS (Ritsema et al., 2011) at 100, 500and 2800 km depth. Bottom: The corresponding null-space component mnull. The null-space component containsshort-wavelength structure that can be scaled and added to the model without changing the misfit. (See ColorPlate 11).

With a few notable exceptions (e.g., Bijwaard& Spakman, 2000; Panning & Romanowicz,2006), all classical P- and S-velocity models areobtained by a linear inversion using expression(11.2). The most important tool to assess theamplitude and shape of the models obtained bylinearized inversions is the resolution operator,provided that the influence of the data errors issmall (Equation 11.3). The latter is the case formost strongly regularized models. The resolutionis an operator which tells us how the obtainedmodel parameters are linearly related to eachother. Ideally we would like to construct amodel such that the resolution is the identitymatrix, meaning that the data can constrainall the chosen parameters separately with thecorrect amplitude. A typical global seismictomography model consists of few thousand toa few hundred thousand model parameters, andthis number squared is the number of entriesin the resolution matrix. It is therefore easilyunderstandable that the latter can computation-ally be a challenge, although not impossible(e.g., Soldati et al., 2006). The importance ofcalculating the resolution matrix was put forward

for comparing seismic tomography to geodynamicmodels (e.g, Ritsema et al., 2007), but more oftenthan not simple synthetic tests of the chequerboard type are used as a proxy for the resolutionmatrix. (Leveque et al. 1993) convincingly arguedthat such a simple test has to be interpreted withcaution as it can be quite misleading and hardlyrepresentative of the true resolution.

As a result of regularization, most resolutiontests indicate that only a quarter to a third of theamplitude of heterogeneities is recovered (e.g., Liet al., 2008; Ritsema et al., 2007). This observa-tion is crucial, but often forgotten, when combin-ing seismic tomography and mineral physics datato estimate the thermo-chemical structure of themantle. Equation-of-state modeling (e.g., Karato& Karki, 2001; Trampert et al., 2001; Stacey& Davis, 2004; Stixrude & Lithgow-Bertelloni,2005; 2011) of mineral physics data allows usto infer sensitivities (partial derivatives) of ve-locity variations to temperature and chemicalvariations. To make the conversion, amplitudeand position of the seismic anomalies has to beknown with great precision. For instance, if onlytemperature is changing at a depth of 2800 km,

Page 6: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

Global Imaging of the Earth’s Deep Interior 329

a relative shear velocity variation d ln vs = 0.01would correspond to a temperature variation ofthe order of 400 K, using the temperature sen-sitivities of (Trampert et al., 2004), a plausiblevalue from current geodynamic models. The un-derestimation of amplitude in tomography hasprofound consequences on this conversion. If wehave to multiply the amplitudes by a factor of3–4, the resulting temperature variation wouldbe 1200–1600 K. This, however, is difficult toreconcile with mantle flow models.

The shape of the anomalies is mainly influ-enced by the off-diagonal elements of the reso-lution matrix. They can be very significant andusually oscillate strongly. This is most easily vi-sualized by evaluating the null-space componentof a given model. Every model vector can be de-composed into a part that lies in the range of theforward operator G and in its null-space. The null-space component of a model is defined by findingthe part for which Gm = 0. It can be expressed as(Deal & Nolet, 1996)

mnull = (I − R) m . (11.11)

By definition, this null-space component does notchange the fit to the data. In practice, Equation(11.11) is only exact if the estimated model ex-plains the data exactly. For most regularizations,such as in Equation (11.7) where there is a com-promise between data fit and model norm, thisis not the case. The data can then not distin-guish between m and m + γ mnull for a certainrange of scalars γ determined by the data uncer-tainty, or equivalently the tolerance on the misfit(de Wit et al., 2012). Nevertheless expression(11.11) is useful for illustrating certain proper-ties of the resolution. If it is a scaled versionof the original model, R has little off-diagonalelements and it can be used to directly visual-ize the amplitude uncertainty mentioned above.If the resolution matrix oscillates, so does thenull-space operator (I − R), and hence (I − R)malso has a strong short wave-length component.Figure (11.1) shows, as an example, the null-space component of the shear velocity modelS40RTS (Ritsema et al., 2011). It has a significantshort wave-length component which is mostly

dermined by regularization. A similar result isseen for P velocity models (de Wit et al., 2012).

It appears thus that the amplitude of isotropicwave speed models is significantly underesti-mated and the smallest wave-lengths in eachmodel strongly distorted. In any thermo-chemicalinterpretation this should, but has not yet, beentaken into account.

11.4 Anisotropic Velocity Tomography

When physical properties depend on the directionin which they are measured, the medium is saidto be anisotropic. A medium with properties de-pending on position only is called isotropic. Inseismology we treat the Earth as having isotropicmass density, but velocities are often found to beanisotropic which can easily be modeled with anappropriate elastic tensor, with two independentparameters for an isotropic medium and between3 and 21 for an anisotropic medium. The effect ofanisotropy on elastic wave propagation is three-fold: obviously wave propagation depends on thedirection of propagation, it alters the wave polar-ization away from parallel or perpendicular to thepropagation direction and it provokes shear wavesplitting, a phenomenon similar to birefringencein optical media. All three properties are readilyobserved for seismic waves. We will only cite se-lected examples to show the diversity of observa-tions without the intention of being exhaustive.An excellent overview on theory and observationof seismic wave propagation in anisotropic mediais by Maupin and Park (2007).

There is large agreement amongst seismologiststhat seismic wave propagation is anisotropic.Although in global studies the crust is usuallyassumed to be isotropic, there are many reports oflocal wave propagation requiring an anisotropicdescription of the crust. Seismic laminationobserved in many parts of the lower continentalcrust (e.g., Meissner et al., 2006), polarizationanomalies of Lg waves (e.g., Maupin, 1990)and S-wave splitting from local earthquakes(e.g., Paulssen, 2004) are all best explained withcrustal anisotropy. Most observations concernwaves propagating in the upper mantle or

Page 7: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

330 jeannot trampert and andreas fichtner

interpretations concerning the upper mantleonly. Significant azimuthal anisotropy has beenobserved for Pn waves beneath oceans (Hess,1964) and continents (Smith & Ekstrom, 1999) aswell as for travel time residuals of teleseismic Pwaves (e.g., Babuska et al., 1998). There are clearobservations of polarization anomalies of longperiod P-waves which have been used to inferupper mantle anisotropy (Schulte-Pelkum et al.,2001), but the biggest wealth of observationscomes from SKS-splitting measurements (for areview see e.g., Long & Silver, 2009). Surfacewaves also exhibit a clear anisotropic behaviour.Both Love and Rayleigh waves show azimuthaldependencies for fundamental modes (e.g., Tram-pert & Woodhouse, 2003; Ekstrom, 2011) as wellas overtones (Visser et al., 2008). The azimuthallyaveraged phase velocities of Love and Rayleighwaves see a different vertically averaged elasticstructure. This is known as the Love-Rayleighdiscrepancy and was first interpreted by (Ander-son 1961) in terms of anisotropy. A transverseisotropic medium with vertical symmetry issufficient the reconcile Love and Rayleigh wavepropagation. Most joint inversions of Love andRayleigh wave phase and/or group velocities forthree-dimensional structure therefore employ atransverse isotropic description of the uppermostmantle (e.g., Montagner & Tanimoto, 1991; Gunget al., 2003; Kustowski et al., 2008). There arefew clear inferences of anisotropy in the mantletransition zone, but surface wave overtonessuggest azimuthal (Trampert & van Heijst, 2002)and radial (Visser et al., 2008) anisotropy. There isa large consensus that the lower mantle is devoidof anisotropy with the exception of D’’ (e.g.,Maupin et al., 2005; Panning & Romanowicz,2006). Wave propagation through the inner coreis best explained using anisotropy although thedetails are not fully mapped yet (e.g., Morelliet al., 1986; Woodhouse et al., 1986; Ishii &Dziewonski, 2002; Beghein & Trampert, 2003;Deuss et al., 2010).

In mathematical terms, seismic anisotropy iseasily understood. From continuum mechan-ics we know that if the local elastic tensorhas more than two independent parameters,

wave propagation depends on the direction ofpropagation. With seismic observations we can-not estimate the elastic tensor at a specific pointin space, but only spatial averages from tens tothousands of kilometres depending on the waves.This makes the seismic inferences difficult tointerpret. From geodynamics and mineral physicsmodeling we know that anisotropic mineralsor isotropic inclusions can align themselvesin preferred strain orientations. The former isknown as lattice preferred orientation (LPO) andthe latter as shape preferred orientation (SPO)(e.g., Karato, 2008). Because the orientation ofthese crystals and inclusions is not instanta-neous, the observation of seismic anisotropypotentially gives us valuable constraints onthe geological history of the strain field (e.g.,Wenk et al., 2011). As mentioned above, seismicobservations cannot image individual crystal orinclusion, but only spatial averages of them.The (an)isotropic description therefore is scale-dependent. Backus (1962) showed that a stackof thin (sub-wavelength) isotropic layers is seenby seismic waves as a large-scale anisotropicmedium. This concept can be generalized, andin fact any large scale description of small scaleisotropic heterogeneity has to include anisotropy(e.g., Capdeville et al., 2010). While this isa fundamental property of the mathematicaldescription at a limited wavelength, anotherambiguity is often forgotten. To infer seismicanisotropy from travel time or polarizationanomalies, an inverse problem is usually solved.Starting from Equation (11.1), let us explicitlypartition the model parameters into isotropic (i)and anisotropic (a) parts. Neglecting data errorsand the starting model for notational simplicity,we find

d = (Gi|Ga)(mi|ma)T. (11.12)

From Equation (11.8), we see that the inverseoperator will also partition and hence(

mi

ma

)=

(SiGi SiGa

SaGi SaGa

)(mi

ma

)(11.13)

The ill-posedness of the inverse problem requiresregularization which implies that the resolution

Page 8: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

Global Imaging of the Earth’s Deep Interior 331

operator is not the identity matrix and hencetrade-offs between isotropic and anisotropicparameters are artificially introduced. Partialderivatives for isotropic and anisotropic param-eters are different and so are the partial inverseoperators, and hence the resolution operatoris not symmetric. This means that the trade-off from isotropic to anisotropic parameters,SaGi, is different from anisotropic to isotropicparameters, SiGa. The isotropic-anisotropictrade-off can have large consequences for ourinterpretation of the inferred anisotropy. If weconstruct global phase velocity maps using apurely isotropic description, i.e. Sa forced to 0,the models appear blobby. Once anisotropy isintroduced, both the isotropic and anisotropicparts of the models appear smooth, with a similarfit to the data (Trampert & Woodhouse, 2003).Hence, the strength of anisotropy trades off withthe roughness of isotropic velocity variations.Another example is that significant apparenttransverse isotropy is generated if the crustalmodel is inappropriate (Bozdag & Trampert,2008), or when apparent splitting is seen onSdiff caused by isotropic velocity gradients in D’’(Komatitsch et al., 2010).

We thus have to keep in mind, when seismol-ogists report anisotropy, they explain the datausing a description involving the least possibleparameters. They implicitly or explicitly employOccam’s razor. In splitting measurements, the ra-zor is implicit as the interpretation assumes aone- or multi-layered medium a priori (e.g., Long& Silver, 2009). A similar situation holds whennormal-mode splitting functions are used as themedium is parametrized with a small numberof unknowns (e.g., Beghein & Trampert, 2003;Deuss et al., 2010). With surface wave measure-ments, the razor is more explicit, as we searchfor a parametrization which explains the data sig-nificantly better (e.g., Trampert & Woodhouse,2003). Due to the wavelength and resolution op-erator averaging, this is the best the seismologistscan do. However, the interpretation of the mod-els evoking LPO or SPO is only meaningful if thisaveraging is taken into account. There are manystudies which show that anisotropy can develop

during convection in the upper (e.g., Kaminskiet al., 2004; Becker et al., 2008) and lower mantle(e.g., McNamara et al., 2002). The comparisonsto seismology, so far, are qualitative, but can beformulated in a quantitative way using the res-olution operator. This is important, because theresolution strongly influences the recovered mag-nitude of anisotropy and properly accounts for thetrade-offs.

This leaves the burning question: is the Earthanisotropic? As argued above, seismologists findthat given the observations and the used modelparametrization, the data are usually explainedmore efficiently, i.e. with less parameters, usinganisotropy. They present mathematical equiva-lents of the physical Earth, filtered by a limitedfrequency band and a nonsymmetric resolutionoperator. There is consensus that the upper fewhundred kilometres of the Earth are transverselyisotropic, with decaying amplitude and a pos-sible sign change around 200 km (Figure 11.2).There is also agreement that there is azimuthalanisotropy, although there is little similarity be-tween the inferred models (e.g., Becker et al.,2007). There are many reports of anisotropy in theD’’ layer, but the different models tend to havefew features in common. There is a large con-sensus that the inner core is anisotropic, but thedetails have yet to emerge. We concur with Beckeret al. (2007) that the reason for this disparity aredifferent strategies in the inverse modelling andwe are currently comparing models with widelydiffering averaging properties. Unless the seismol-ogists come to a similar degree of consensus asfor isotropic structures, it seems difficult to inferany geodynamic constraints from the anisotropicmodels. Anisotropy needs far more parametersfor a proper description than isotropy. A possi-ble strategy is therefore to use implicit scalingsor fix the symmetry a priori based on mineralphysics arguments (e.g., Panning & Romanowicz,2006; Chevrot, 2006; Long et al. 2008; Panning& Nolet, 2008). Another strategy is to be guidedby the data and invert for parameters they aremost sensitive to (e.g. Sieminski et al. 2009). Thelatter is easier to interpret because it is difficult

Page 9: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

332 jeannot trampert and andreas fichtner

200

400

600

800

1000

1200

1400

0.94 0.96 0.98 1

ξ1.02 1.04 1.06

dept

h [k

m]

uncertaintymean

PREM

Fig. 11.2 Probability density function of thespherically averaged anisotropic parameter ξ , modifiedafter (Visser et al. 2008). Plotted is the mean model and2 standard deviations around it.

to quantify isotropic-anisotropic trade-offs if theparametrization is fixed a priori.

11.5 Density Tomography

Lateral variations in density are the source ofmass transport in the Earth on all scales rangingfrom the upwelling of salt bodies to the sub-duction of lithospheric plates and the ascent ofsuper-plumes. They are key to the resolutionof long-lasting debates including the distribu-tion of compositional heterogeneity, the natureof continental lithosphere, and the contributionof mantle convection to surface tectonics. How-ever, despite its relevance, the detailed 3D densitystructure of the Earth remains largely unknown.

In contrast to seismic velocities that can be in-ferred from travel times, most surface observables

bear little information on density. While be-ing most directly related to density, gravity pro-vides only weak constraints, because solutionsto the gravitational inverse problem are inher-ently non-unique. The use of free air gravity orthe geoid is further complicated by the effects ofboundary perturbations. Volumetric density het-erogeneities by themselves affect gravity, but theyalso generate topography on the Earth’s surfaceand internal interfaces, such as the core-mantleboundary. The strength of the boundary pertur-bations depends on viscosity, which is, however,poorly constrained. Furthermore, the volumetricand boundary perturbation effects nearly cancel atmid-mantle depths, which precludes any robustinference on 3D density structure.

The travel times of P and S waves - the ma-jor source of information on the Earth’s velocitystructure – are practically insensitive to densityheterogeneities. This surprisingly complete ab-sence of information is closely related to thescattering characteristics of velocity and densityperturbations (e.g. Wu & Aki, 1985; Tarantola,1986). When a body wave impinges upon a ve-locity perturbation, it generates a scattered wavethat travels along with the incident wave. Theinterference with the scattered wave modifiesthe shape of the original wave. This modifica-tion is then perceived as a travel time shift thatcan be used in travel time tomography to trackthe causative velocity perturbation (e.g. Dahlenet al., 2000). Density heterogeneities, in contrast,generate scattered waves that travel opposite tothe direction of the incident wave. Therefore,the incident and the scattered waves can neverinterfere, meaning that the original pulse shaperemains unaffected. It follows that travel times ofthe classical body wave phases are not influencedby density perturbations, and that informationon density must be sought in the later parts ofseismograms. This phenomenon is illustrated inFigure 11.3. We note that the backward scatter-ing off density heterogeneities depends on theparametrization of the Earth model. For instance,choosing ρ together with the elastic parame-ters λ and μ, instead of ρ, vp and vs, leadsto forward-scattering off density heterogeneities,

Page 10: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

Global Imaging of the Earth’s Deep Interior 333

velocity perturbation (dvs ≠ 0 or dvp ≠ 0 and dρ = 0)

before scattering after scattering

density perturbation (dvs = 0, dvp = 0 and dρ ≠ 0)

before scattering

superposition perceivedas traveltime shift

scattered wave

incident wave scattered wave scattered waveincident wave

after scattering

Fig. 11.3 Scattering characteristics of velocity and density perturbations (e.g. Wu & Aki, 1985; Tarantola, 1986).Left: An incoming body wave impinges upon a velocity perturbation (black star) with either dvs �= 0 or dvp �= 0 anddρ = 0. The scattered wave, indicated by the double dashed curve, propagates in the same direction as the incomingwave. The superposition of both waves is perceived as a travel time delay or advance, depending on the sign of theperturbation (e.g. Dahlen et al., 2000). Right: The same as to the left, but for a density perturbation with dvs = 0,dvp = 0 and dρ �= 0. The scattered wave propagates opposite to the direction of the incoming wave. The pulse shapeof the incoming wave remains unperturbed.

i.e., scattering in the propagation direction of theincoming wave. While this parametrization in-troduces body wave sensitivity to density, it alsoleads to undesirable trade-offs between the modelparameters, similar to Equation (11.13).

The frequency-dependent travel times of sur-face waves (dispersion) are mildly sensitive todensity variations. Unfortunately, however, thissensitivity tends to be oscillatory. A density per-turbation may therefore interact with both posi-tive and negative parts of the sensitivity distribu-tion. The net effect is then nearly zero. Additionalcomplications arise from the strong trade-offsbetween velocity structure and density in themulti-parameter inverse problem. Perturbationsin seismic velocities and density influence sur-face wave dispersion simultaneously, and the two

effects cannot easily be distinguished. This ledmany authors to either ignore density variationsor to scale vs to ρ using various multiplicative fac-tors (e.g. Panning & Romanowicz, 2006; Ritsemaet al., 2011).

At the long-period end of the seismic spec-trum, around 0.5–3.0 mHz, the spheroidal freeoscillations of the Earth are sensitive to the long-wavelength density distribution in the wholemantle because of the gravitational restoringforce (Figure 11.4). Lateral heterogeneities leadto the splitting of normal-mode eigenfrequenciesthat can be used for tomography. However, sincenormal modes result from the constructive inter-ference of waves traveling in opposite directionsaround the globe, they are primarily sensitiveto even-degree structure, because odd-degree

Page 11: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

334 jeannot trampert and andreas fichtner

Fig. 11.4 Sensitivity of thefrequencies of the spheroidal modes

1S4 and 2S6 to perturbations in theradial distributions of vp (dotted), vs(solid) and ρ (dashed) away from the1D reference Earth model ak135(Kennett et al., 1995). The peakamplitude of the density sensitivity isgenerally smaller than the sensitivityto vs. Similar to surface waves, thenormal-mode sensitivity to density isoscillatory, which further reduces thenet effect of large-scale densityvariations. centre

660 km

surface

CMB

ICB

centre

660 km

surface

CMB

ICB

centre

1 S 4 2 S 6

vp

vs

ρ

contributions tend to cancel. While coupledmodes provide some information on odd-degreestructure, they have so far not been used intomographic inversions for density.

The weak sensitivities and strong trade-offsmap directly into large uncertainties, even inthe radially symmetric density structure thatis additionally constrained by the Earth’s massand moment of inertia. Exploring the space ofone-dimensional Earth models with Monte Carlosampling, Kennett (1998) found that density vari-ations of around 1% over 200 km depth inter-vals are compatible with the frequencies of thegravest spheroidal modes and their associated er-rors. These uncertainties must be kept in mindwhen interpreting lateral variations relative to aone-dimensional average.

Compared to the abundance of tomographicinversions for velocity structure, few attemptshave been made to exploit the scarce informationon density from surface observations.

By far the most common approach to infer3D density, is to transform tomographic P andS velocity models using a depth-dependentscaling. This approach relies on mineral physicsmodeling (e.g. Karato, 1993) and works under the

assumption that density heterogeneities are ofpurely thermal origin. While being a convenientchoice in the absence of compositional infor-mation, the resulting density models are ofteninconsistent with fundamental geodynamicobservations such as free air gravity anomalies(Forte, 2007). The neglect of compositionalcontributions to density heterogeneity can fur-thermore lead to incorrect predictions of mantleflow and the associated surface deformation.

To overcome such inconsistencies, Tondi et al.(2000, 2009) jointly inverted seismic travel timesand Bouguer anomalies for lithospheric structureby imposing linear relations between velocitiesand density. The scaling links the two other-wise independent data sets by assuming that agiven seismic velocity uniquely specifies densityand vice versa. Using inverse problem termi-nology, the scaling reduces the number of freemodel parameters, thereby alleviating the inher-ent non-uniqueness of pure gravity inversions.This approach produces models of the lithospherethat explain both gravity and seismic data sets,but it does not allow for a decorrelation of ve-locities and density that is likely to result fromcompositional heterogeneities.

Page 12: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

Global Imaging of the Earth’s Deep Interior 335

Along similar lines, Simmons et al., 2010 as-similated seismic travel times, the global free airgravity field, the divergence of tectonic plates, dy-namic surface topography and topography of thecore-mantle boundary into a joint velocity anddensity model. Their method relies on the radialviscosity profile of the Earth (Mitrovica & Forte,2004), and on the assumption that lateral hetero-geneities are most likely of thermal origin. Whilebeing a significant step forward, this approach islimited by the bias towards thermally induceddensity variations, the large uncertainties in theviscosity profile, and the absence of a formal erroranalysis that accounts for plausible variations ofthe forward modeling assumptions.

Recognizing that enforced velocity-density cor-relations may prevent important inferences onthermo-chemical structure from the outset, Ishii& Tromp (1999, 2001, 2004) inverted normal-mode splitting and free air gravity for a globaldegree-6 model of velocities and density that werenot related a priori. Their results suggest a decor-relation of velocity and density variations, andin particular the presence of high-density piles inthe deep mantle beneath the African and Pacificsuper-plumes.

The robustness of Ishii and Tromp’s densitymodels was at the centre of a long debate thatreflects the difficulty of solving a deterministicmulti-parameter inverse problem that is char-acterized by weak constraints, strong trade-offsand the nearly complete absence of true physicalprior information. Resovsky & Ritzwoller (1999)pointed out that the resolution analysis in a de-terministic inverse problem depends on the priorknowledge, i.e. the choice of the prior model co-variance σ 2

m in Equations (11.8) and (11.9). Sincewe have little prior knowledge on density vari-ations in the Earth, σ 2

m corresponding to densityshould be very large. The weakness of our con-straints furthermore implies that the entries ofthe forward modeling or sensitivity matrix G aresmall. It follows that the matrix GTG + σ 2

d σ−2m I

in Equation (11.8) may practically not be invert-ible, unless σ 2

m is chosen smaller than it actuallyis on the basis of our prior information. Theprior model covariance is then used to regularize

the inversion in a subjective way, and not torepresent an objective state of knowledge. Whenthe data constraints are weak, this subjectivitydominates the perceived resolution, expressed, forinstance, in terms of the posterior covariance Cm(Equation 11.9). Based on a series of test inversionswith different choices of the prior information,Resovsky & Ritzwoller 1999 concluded that adecorrelation of S velocity and density could notbe detected reliably. A similar conclusion wasreached by Romanowicz (2001), who found thatdensity in the lower mantle trades off stronglywith the topography of the core-mantle bound-ary, and that gravity data hardly discriminatebetween different density models. However, Ro-manowicz (2001) was able to infer rough boundsfor the depth-dependent relationship between thedegree-2 structure of density and S velocity. Therobustness of 3D density variations in determinis-tic inversions was furthermore questioned by Kuo& Romanowicz (2002) on the basis of syntheticexperiments.

To circumvent the subjectivity of determinis-tic tomographies, Resovsky & Trampert (2003)worked with a probabilistic formulation of the in-verse problem, that does not require explicit regu-larization to stabilize and reduce trade-offs. Usinga combined set of normal-mode and surface-wavedata they explored the model space with the helpof a neighbourhood algorithm (Sambridge, 1999a;b) to produce probability distributions for velocityand density heterogeneities. These probabilitydensities provide a complete description of ourstate of knowledge, including uncertainties. Theresults of Resovsky & Trampert (2003) suggestthat long-wavelength variations in vs and ρ areunlikely to be correlated as much as variationsin vp and vs anywhere in the mantle. Especiallywithin the transition zone the data favour ad ln vs-d ln ρ anti-correlation. Using mineralphysics relations between seismic properties,temperature and composition, Trampert et al.2004 were able to discriminate between thermaland compositional contributions to observeddensity variations, suggesting that high-densityanomalies in the deep mantle (2000–2891 kmdepth) are most likely of compositional origin.

Page 13: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

336 jeannot trampert and andreas fichtner

dln Vs [%]

950 km (+/− 0.82)

dln rho [%]

950 km (+/− 1.1)

1 0.8 0.6 0.4 0.2 0 0.2 0.4 0.6 0.8 1

1000

1200

1400

1600

1800

2000

2200

2400

2600

2800

3000

Correlation

Dep

th [k

m]

dln vs dln ρdln vs dln vp

dln vs dln vc

1500 km (+/− 0.85) 1550 km (+/− 1.0)

2600 km (+/− 0.95) 2600 km (+/− 1.0)

2891 km (+/− 1.0) 2891 km (+/− 0.99)

−1.8 1.80.0 −1.8 1.80.0

Fig. 11.5 Left: Lateral variations in vs and ρ at various depths in the maximum-likelihood model of Mosca et al.2012. The laterally averaged standard deviations are indicated in brackets. Note the pronounced anti-correlation ofd ln vs and d ln ρ around 2600 km depth beneath the central Pacific and Africa. (See Color Plate 12). Right:Correlation coefficients as a function of depth between the most likely models of d ln vs, d ln ρ, d ln vp andd ln vc, where vc denotes the bulk sound velocity. Figure modified after Mosca et al. 2012.

Recently, the studies on probabilistic tomog-raphy were extended by Mosca et al. (2012)who incorporated improved splitting functionmeasurements, as well as a large collection ofbody wave travel times (figure 11.5).

In the light of the weak constraints, it is notsurprising that widely accepted seismological in-ferences on the Earth’s density structure are rel-atively few in number. That subducting slabsand ascending plumes in the uppermost mantleare comparatively dense or light, respectively,follows from geodynamic arguments, but un-equivocal seismic evidence on length scales abovedegree 6 is missing.

The most reliable results concern the corre-lation between velocity and density structureat degrees 2, 4 and 6. In the mid-mantle, be-tween 660 and around 1800 km depth, d ln vsand d ln ρ are most likely to be uncorrelated.Within the lower mantle, seismic data prefer a

mild anti-correlation of d ln vs and d ln ρ, witha correlation coefficient around −0.4 (see Figure11.5). The only concrete structural inference re-lates to the presence of high-density material inthe lowermost mantle beneath the African andPacific superplumes that has been found consis-tently by several independent studies (e.g. Ishii& Tromp, 2001, 2004), Trampert et al., 2004;Simmons et al., 2010; Mosca et al., 2012). All ofthese results must be interpreted with caution,given the relatively large standard deviations thatrange around 50% of the maximum relative den-sity perturbations, even when large data sets ofsurface waves, normal modes and body waves arecombined.

Finally, we note that robust seismic informa-tion on the density in the upper mantle is so farmissing. This is because structure above 660 kmdepth cannot be adequately described in terms ofthe lowest-degree spherical harmonics that we are

Page 14: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

Global Imaging of the Earth’s Deep Interior 337

able to recover reliably from long-period normal-mode data.

It is becoming increasingly evident that newinferences on the details of Earth’s 3D densitystructure will force us to progress beyond bothtraditional data analysis and modeling. The accu-rate measurement of splitting functions on datafrom recent megathrust earthquakes, combinedwith proper mode coupling, has the potential toyield more robust information, also about odd-degree density structure (Deuss et al., 2011).Several nonseismological developments may alsohelp to improve our knowledge concerning den-sity. Lateral variations of the Earth’s body tideshave recently been shown to be above the ob-servational error, and may thus yield additionalconstraints on global density structure (Latychevet al., 2009). Also on a global scale, neutrino to-mography et al., originally conceived in the early1980s (e.g. De Rujula et al., 1983) – may providevery direct insight into density structure, that isfree from the trade-offs that plague seismologicalapproaches (e.g. Gonzalez-Garcia et al., 2008). Onsmall scales (up to a few kilometres) muon trans-mission tomography has been used successfullyto image low-density magma conduits inside ac-tive volcanoes (e.g. Tanaka et al., 2003, 2010).Both tidal and elementary particle tomographiesare, however, in their infancy.

11.6 Attenuation Tomography

Seismic waves propagating through the Earth areattenuated due to various relaxation mechanismsthat lead to the transformation of elastic energyinto heat. The large interest in the absorptionproperties of the Earth, parametrized in terms ofthe quality factor Q, is mostly related to theirtemperature dependence. While seismic veloci-ties are quasi-linearly related to temperature, T,the temperature-dependence of Q is well approxi-mated by an exponential Arrhenius-type law (e.g.Jackson, 2000)

Q−1 ∝ e− ERT , (11.14)

where E and R denote an activation energy andthe gas constant, respectively. Before delving intothe details of the Earth’s Q structure, we givea brief summary of the description of seismicwave attenuation, complemented by the mostfundamental observations.

11.6.1 Description of seismic waveattenuation, basic observables and observations

Assuming that all seismologically relevant re-laxation mechanisms can be modeled by lin-ear rheologies, shear attenuation is commonlydescribed by the frequency-domain visco-elasticstress-strain relation

σ (ω) = μ(ω)ε(ω), (11.15)

where μ denotes the complex shear modulus inthe frequency domain. The ratio between the realand imaginary parts of μ defines the shear qualityfactor Qμ:

Qμ(ω) = �μ(ω)ζ μ(ω)

. (11.16)

The bulk quality factor Qκ is defined similarlyon the basis of the complex bulk modulus κ.The observability of Q in the Earth is closelyrelated to the amplitude decay of seismic waves.For instance, when Qμ >> 1, the amplitude A of aplane S wave propagating through a homogeneousand isotropic medium behaves as

A ∝ exp

(−ω x

2vsQμ

), (11.17)

where x denotes the distance traveled. Relation(11.17) is frequently adapted in ray-theoreticalstudies that replace x/Qμvs by

t∗ =∫C

dsQμ(x) vs(x)

, (11.18)

with C and ds denoting the ray path and a pathsegment, respectively. Analogous relations canbe derived for plane P waves. Most body wave

jeannot
Highlight
jeannot
Sticky Note
replace with symbol for ' imaginary part'
Page 15: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

338 jeannot trampert and andreas fichtner

attenuation tomographies are based on variants ofEquations (11.17) and (11.18) because they allowus to eliminate the contribution of the potentiallyuncertain seismic moment by measuring ampli-tude ratios of different seismic phases (e.g. Reidet al., 2001; Cheng & Kennett, 2002; Kennett &Abdullah, 2011).

Early studies of Q in the Earth were primar-ily based on the decay of free oscillation peaksand long-period surface wave amplitudes (e.g. An-derson et al., 1965; Canas & Mitchell, 1978;Anderson & Hart, 1978; Sailor & Dziewonski,1978. They established that attenuation in shearlargely dominates over attenuation associatedto bulk deformation, meaning that Qμ << Qκ

almost anywhere in the Earth. Attenuation mea-surements of the Chandler wobble at sub-seismicfrequencies on the one hand, and surface waves onthe other hand furthermore indicated that Qμ ismildly frequency-dependent, with Qμ ∝ ωα and α

in the range of 0.1–0.5 (e.g. Anderson & Minster,1979). These results were confirmed by numer-ous body wave analyses (e.g. Sipkin & Jordan,1979; Flanagan & Wiens, 1998; Cheng & Kennett,2002) and laboratory experiments (e.g. Jackson,2000, 2007; Karato, 2008). In a recent study,

Lekic et al. (2009) developed a method that sep-arates the depth- and frequency-dependencies ofnormal-mode and surface-wave attenuation mea-surements. Their results suggest that the effectiveα in the mantle is negative (≈ −0.4) at periodslonger than 1000 s, transitioning to positive val-ues around 0.3 for periods shorter than 500 s.

Despite the unequivocal evidence for a power-law frequency dependence, tomographic inver-sions mostly assume Q to be constant across theseismic frequency band, i.e. from ≈ 1 to ≈ 10−3

Hz. This simplification is motivated by the dif-ficulty to robustly constrain Q variations in theEarth even within a narrow frequency band. Theconstant-Q model is closely related to the notionof a continuous absorption band, i.e. a continu-ous distribution of relaxation mechanisms thatlead to nearly frequency-independent absorption(Liu et al., 1976), as shown in Figure 11.6. Thisis in contrast to the appearance of isolated ab-sorption peaks that are, for instance, commonlyfound in metals (e.g. Zener, 1948. Outside theabsorption band, Q is predicted to be proportionalto ω−1 at the low-frequency end, and to ω at thehigh-frequency end.

A direct consequence of attenuation in gen-eral is dispersion, i.e. the frequency-dependence

Fig. 11.6 Phase velocity (dashed)and Q−1 (solid) as a function offrequency in an absorption bandmodel with a nearly constant Q of400 in the frequency band from≈ 10−5 Hz to 1 Hz. Afrequency-dependent Q withinthe absorption band is shownschematically in the form of agray line. A power-law frequencydependence, Q ∝ ωα with aslightly positive α around 0.3 isconsistently found in seismic andlaboratory studies, but commonlyignored in tomographicinversions.

1/400

1/500

1/1000

Q−1

Q−1

Phase velocity

5.08

5.04

Pha

se v

eloc

ity

5.00

−8 −6 −4 −2 0

Log (frequency)

2 2

jeannot
Inserted Text
)
Page 16: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

Global Imaging of the Earth’s Deep Interior 339

of phase velocity (Figure 11.6). In the constant-Q model (Liu et al., 1976), the phase velocitydispersion relation is given by

vphase(ω)

vphase(ωref )= 1 + 1

π Qln

ωref

), (11.19)

where ωref is a reference frequency. Equation(11.19) holds, provided that Q >> 1, and that ω

lies within the absorption band. Equation (11.19)suggests that phase velocity increases monoton-ically with increasing frequency; and this effectis most pronounced in highly attenuative mediawhere Q is small.

It is important to make a clear distinction be-tween velocity as a property of a traveling wave,and velocity as a material property. The phasevelocity in Equation (11.19) is strictly speakingthe propagation speed of a monochromatic planewave in an isotropic and homogeneous medium.This is not generally identical with a velocity asmaterial property, given by the square root of anelastic modulus divided by density.

11.6.2 The nature of the inverse problem

The dependence of the phase velocity on both Qand ω has profound implications on our abilityto infer the absorption properties of the Earth.On the one hand, there is no seismological ob-servable that is sensitive to variations in Q only,while being practically insensitive to variationsin any other parameter. On the other hand, thereis also no seismological observable that is unaf-fected by Q. The amplitudes of both long-period(50–200 s) surface waves and intermediate-period(5–30 s) body waves are found to be dominatedby focusing induced by 3D variations of purelyelastic structure (Sigloch, 2008; Zhou, 2009; Tianet al., 2009). This does not mean that Equation(11.17) is incorrect, but it is a reminder that aproportionality derived from plane wave analysisin homogeneous media should not be mistakenfor an equality that holds in the real Earth. Theimpact of attenuation on travel times should alsonot be underestimated. Based on the analysis offinite-frequency sensitivity kernels, Zhou (2009)

infers that 15–20% of the observed phase delays inlong-period surface waves may in fact be the resultof 3D heterogeneity in Qμ. It follows, in conclu-sion, that any inference on the spatial distributionof Q requires the solution of an intrinsically cou-pled multi-observable/multi-parameter problem.

The inverse problem is complicated by the dif-ficult nature of seismic wave amplitudes that arefrequently used as the only measurement to con-strain Q. In addition to attenuation, amplitudesare strongly affected by focusing. Waves trav-eling through a low-velocity region are focusedand amplified, whereas defocusing and amplitudereduction occur in high-velocity zones. Further-more, amplitude measurements are affected bymiscalibrated instruments, scattering and seismicsource characteristics. Small-scale heterogeneityin the immediate vicinity of the receiver, suchas layers of nearly saturated sediments, may giverise to nonlinear amplitude effects. These fac-tors add frequency-dependent source and receivercorrections to the list of unknowns.

The daunting complexity of the inverse prob-lem for Q explains why substantial simplifica-tions are common. Instead of jointly invertingfor all the necessary free parameters, both elasticfocusing effects and source/receiver correctionsare frequently ignored. While convenient, this ap-proach was shown to produce incorrect estimatesof attenuation structure in body and surface wavestudies (Dalton et al., 2008; Tian et al., 2009). Toisolate the signature of attenuation, several stud-ies proposed to remove the focusing effect fromobserved amplitudes with the help of 3D elasticvelocity models derived from travel time tomog-raphy. While being a significant step forward,this strategy suffers from the subjective choiceof an elastic model, and its smoothness proper-ties in particular. Overly rough elastic modelsmay lead unrealistically strong focusing, whileoverly smooth models will compensate the un-derestimated elastic effect by excessive Q vari-ations. Interestingly, this scale-length problemis similar to the one encountered in anisotropictomography where the roughness of isotropic ve-locity variations trades off with the strength ofanisotropy. In any case, the resulting combination

Page 17: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

340 jeannot trampert and andreas fichtner

of elastic velocity and attenuation models will beincompatible, because the deduced 3D variationsin Q were not accounted for in the interpretationof travel times used to derive the velocity modelthat was initially needed to remove the elasticeffect from amplitudes. This discrepancy under-lines the intrinsic nonseparability of elastic andanelastic effects that precludes any consistentremoval of focusing effects from seismic waveamplitudes.

Both the multi-observable/multi-parameter na-ture of the inverse problem for Q structure andthe inconsistency of many simplifications explainwhy our progress in attenuation tomography hasbeen rather slow, compared to tomography forelastic structure.

11.6.3 Constraints on the radial attenuationstructure

Constraints on the spherically averaged attenu-ation in the Earth are mostly derived from theamplitude decay of body waves, long-period sur-face waves and free oscillation peaks.

Estimates of the radial shear attenuation byvarious authors, summarized in Figure 11.7, arerelatively consistent. The only notable exceptionis the lithosphere where values range betweenQμ ≈ 200 (Durek & Ekstrom, 1996) and Qμ ≈ 600(Dziewonski & Anderson, 1981; Okal & Jo, 1990).This variability partly reflects the presence ofstrong heterogeneity in the lithosphere that pre-vents the determination of a physically meaning-ful average. Within a narrow asthenospheric layer,Qμ is generally found to be low (60–90). This dropis followed by a rapid increase through the transi-tion zone to Qμ ≈ 300 near 660 km depth. Withinthe lower mantle some authors favour a slightdecrease (Okal & Jo, 1990; Widmer et al., 1991),while others prefer a constant Qμ ≈ 300 downto the core-mantle boundary (Dziewonski & An-derson, 1981; Durek & Ekstrom, 1996). Reliableuncertainty estimates for the radial Qμ profilewere provided by Resovsky et al. (2005), who ap-plied a probabilistic inversion to normal-modeand surface-wave data. Typically, the standard

deviations were found to range between 10% and20%, depending on depth.

The distribution of bulk attenuation, requiredby the decay of radial free oscillations, is thesubject of a long-lasting debate, in the courseof which a large variety of models have been pre-sented. Some of these are displayed in Figure 11.7.Dziewonski & Anderson (1981) confined a com-paratively large bulk attenuation in PREM tothe inner core, where Qκ = 1327.7 compared toQκ = 57823 anywhere else in the Earth. Theynoted, however, that ‘‘this should be only under-stood as a way to lower Q of radial modes inorder to make them more compatible with obser-vations. The problem is highly non-unique and itsearly resolution is not likely.’’ Similar to PREM,Anderson & Hart (1978) and Okal & Jo 1990 lo-cate high bulk attenuation (small Qκ ) in the innercore. In contrast, more recent models based onhigher-quality data clearly prefer a distribution oflow Qκ in the upper mantle and the transitionzone (Durek & Ekstrom, 1996), or in the wholemantle and the outer core (Widmer et al., 1991;Resovsky et al., 2005). The variety of Qκ modelsis proportional to the ill-posedness of the inverseproblem. As in the case of density tomography,the observational constraints on Qκ are weak, sothat the solution of a deterministic inverse prob-lem is dominated by the unavoidable and largelysubjective regularization. Any particular value ofQκ should be interpreted with caution, since plau-sible values of bulk attenuation may span nearlyone order of magnitude (Resovsky et al., 2005).

All of the radial Qμ models shown in Figure11.7 are based on the assumption of frequency-independent attenuation, despite the well-knownpower law relation Qμ ∝ ωα with α mostly be-tween 0.1 and 0.5. Additional free parameters todescribe frequency dependence are unlikely to beresolvable. As pointed out by Lekic et al. (2009),this simplification may lead to incorrect Qμ pro-files, because higher-frequency waves with higherQμ constrain shallower structure, whereas lower-frequency waves with lower Qμ constrain deeperstructure. Thus, attenuation estimates may be bi-ased towards low Qμ values in the deep mantleand towards high Qμ values in the upper mantle.

Page 18: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

Global Imaging of the Earth’s Deep Interior 341

PREM (Dziewonski & Anderson, 1981)QL6 (Durek & Ekström, 1996)

PAR3C (Okal & Jo, 1990)QM1 (Widmer et al , 1991)

mean + σmeanmean − σ

Resovsky et al. (2005)80

220

670 670

2891

5150

6371103 104 105

2891

5150

637170 100 200

400 700Qκ

dept

h [k

m] de

pth

[km

]

Fig. 11.7 Comparison of the radial Q models PREM (Dziewonski & Anderson, 1981), PAR3C (Okal & Jo, 1990),QM1 (Widmer et al., 1991), QL6 (Durek & Ekstrom, 1996), and the model by Resovsky et al. (2005) with ±σ

uncertainties indicated by the light-gray shaded areas.

11.6.4 3D attenuation models

Over the past four decades, numerous regional-scale studies made use of diverse data setsincluding surface wave amplitudes and differen-tial t∗ measurements on pairs of seismic phases.While individual results differ substantially dueto the complexity of the inverse problem and thevarious ways of dealing with it, some robust linksbetween near-surface tectonic features and the

attenuation properties of the uppermost mantleare emerging from the ensemble of available mod-els: Low attenuation, i.e. large Q, is often foundwithin stable continents (e.g. Chan & Der, 1988;Sipkin & Revenaugh, 1994; Tian et al., 2009;Kennett & Abdullah, 2011). This is in contrastto the typically high attenuation in the vicinityof hotspots (e.g. Tian et al., 2009; Kennett &Abdullah, 2011). Along convergent margins,several studies inferred high attenuation in the

Page 19: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

342 jeannot trampert and andreas fichtner

mantle wedge above subducting slabs and lowerattenuation within the downgoing plate (e.g.Flanagan & Wiens, 1990; Stachnik et al., 2004). Adecrease of uppermost mantle attenuation withincreasing age of ocean floor was found, for in-stance, by Canas & Mitchell (1978) and Sheehan& Solomon (1992). These observations suggestthat attenuation is primarily temperature-related,at least within the shallow mantle.

Global models of attenuation in the wholemantle have also been derived mostly from long-period surface wave amplitudes and amplituderatios of various body wave phases. These modelsdiffer strongly in the treatment of elastic ef-fects and source/receiver corrections. Romanow-icz (1990) and Durek et al. (1993), for instance,considered the amplitudes of four consecutive ar-rivals of multiply-orbiting long-period Rayleighwaves (e.g. R1, R2, R3, R4), in order to reduce theeffects of elastic focusing and source uncertainty.While this approach helps to pronounce the con-tribution of attenuation, it only yields informa-tion on even-degree Qμ structure. An improveddata analysis scheme was used by Romanowicz(1994), who selected both R1 and R2 amplitudesthat did not appear to be contaminated by sourceerrors and elastic effects. The selection criterionenforced consistency between the attenuation co-efficients inferred from four consecutive wavetrains, and those measured on minor and majorarc amplitudes only. This strategy led to mapsof both even- and odd-degree structure, that wereused to construct the first global model of shearattenuation (Romanowicz, 1995).

Billien et al. (2000) and Selby & Woodhouse(2000; 2002) explicitly treated elastic focusingusing a linear approximation derived from raytheory (Woodhouse & Wong, 1986). The con-tribution from focusing to the amplitudes ofRayleigh waves was found to be considerable,even for long periods between 70 s and 170 s.This motivated Billien et al. (2000) to jointlyinvert phase and amplitude measurements ofRayleigh waves for degree-20 maps of phase ve-locities and attenuation. However, being con-cerned that focusing may not be predicted withsufficient accuracy by current velocity models,

Selby & Woodhouse (2002) decided not to ac-count for elastic effects in their inversion for 3Dshear attenuation. A similar approach was takenby Gung & Romanowicz (2004) who neglectedfocusing as well as source/receiver corrections,because these factors seemed to have little effecton their degree-8 model in synthetic tests.

As an alternative to surface waves, Reid et al.(2001) estimated t∗ from amplitude ratios of glob-ally recorded S, SS and SSS waves. Also neglectingthe effect of focusing, the t∗ measurements wereinverted for a degree-8 model of Qμ in the uppermantle.

The currently most sophisticated study onwhole-mantle shear attenuation was initiated byDalton & Ekstrom (2006) who simultaneouslyinverted Rayleigh wave amplitudes and phasedelays for maps of shear attenuation and phase ve-locity, as well as for amplitude correction factorsfor each source and receiver. They demonstratedthat attenuation estimates depend significantlyon each of the amplitude corrections, and thatthe neglect of elastic focusing translates intoinaccurate Q structure. The phase delay mapswere then used to remove the focusing effect fromthe amplitude data, which were then invertedfor a global degree-12 Qμ model (Dalton et al.,2008). More than previous models, the work ofDalton et al. (2008) reveals a clear correlationbetween shallow-mantle attenuation and surfacetectonics. In particular, high attenuation is foundaround 100 km depth beneath the circum-Pacificvolcanic arc, the Lau basin, and along mostmid-ocean ridges, including the East Pacific Rise,the Indian-Antarctic ridge and the Mid-Atlanticridge. Precambrian lithosphere is characterizedby lower than average attenuation.

In contrast to elastic P or S tomographies, thereis little agreement between attenuation models.Dalton et al. (2008) compared their global Qμ

model to those of Romanowicz (1995), Reid et al.(2001), Warren & Shearer (2002), Selby & Wood-house (2002) and Gung & Romanowicz (2004).When truncated at degree 8, the correlation ofthese models was found to be mostly below 0.4throughout the upper mantle. Furthermore, Dal-ton et al. (2008) noted that global Qμ models differ

Page 20: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

Global Imaging of the Earth’s Deep Interior 343

most in the vicinity of mid-ocean ridges wherecharacteristic low-velocity regions can lead to astrong focusing of wave energy that must be ac-counted for correctly. The strength of the lateralvariations in global attenuation models varies bya factor of up to 4, which is significantly above theerror estimates of around 10%, given, for instance,by Romanowicz (1995). The various models aretherefore, technically speaking, inconsistent. Thelarge discrepancies between Qμ models and theoverly optimistic error estimates are again relatedto the properties of an ill-posed inverse problemwhere weak constraints and strong trade-offs leadto a dominance of subjective and often implicitregularization.

Despite the difficulties involved in attenuationtomography, some large-scale robust features canbe deduced from the combined analysis of allavailable global attenuation models. Above 250km depth, most studies find a correlation betweenQμ and surface tectonics. The most notable con-sistency is the appearance of high Qμ values instable continental interiors, including Australia’sPilbara and Yilgarn cratons, the East EuropeanPlatform and the Canadian Shield. Below 250km depth the attenuation pattern changes, andlow Qμ can roughly be associated with the posi-tion of hotspots as imaged in isotropic velocitytomographies.

The variability of Qμ models is due to a strongmethodology-dependence that reflects the diffi-culty of extracting a robust signal from seismicwave amplitudes that can be attributed to attenu-ation with sufficient confidence. From the currentperspective, much progress remains to be madein order to arrive at a stage where attenuationmodels can be interpreted quantitatively in termsof the Earth’s thermo-chemical structure.

11.7 Promising Future Directions

From the above discussions it is clear that globaltomography is, in general, a multi-observable/multi-parameter problem. The complicatednature of this problem has to be addressed verycarefully, especially when inverting for weakly-constrained and scale-dependent properties such

as anisotropy, density and attenuation. Theconsensus amongst seismologists decreases fromisotropic to anisotropic, density and attenuationmodels. The reason for this is the nonsymmetricnature of the resolution operator (Equation11.13). The Frechet derivatives are such thattrade-offs to isotropic parameters are smallest.This explains the considerable overlap betweendifferent studies using different approximationsand data. Good data exist to image the elasticpart of the problem, but density is weaklyconstrained. The Frechet derivatives to anelasticparameters are much smaller with commonparametrizations and misfit functionals. Wesee mainly three general future directions: (1)progressing towards full waveform inversionwith appropriate resolution analysis, (2) principalcomponent analysis and (3) finding new observ-ables which are sensitive to a limited range ofparameters.

Full waveform inversion uses gradient-type im-plementations of Equation (11.10) with exactFrechet derivatives computed with purely numer-ical solutions of the forward problem (Bambergeret al., 1982; Tarantola, 1988; Igel et al., 1996).This approach accounts for the nonlinear rela-tion between structure, phase and amplitudes inthe construction of tomographic models, thusprogressing beyond linear approximations com-monly used. Since the update of the model is justa scaled version of the misfit kernel, there is no ex-plicit regularization required, beyond smoothingthe kernels and terminating the inversion aftera finite number of iterations. The first regional-to continental-scale seismological applications offull waveform inversion indicate that resolutionindeed improves (e.g. Chen et al., 2007; Ficht-ner et al., 2009; 2010; Tape et al., 2010). Inparticular, the amplitudes of the heterogeneitiesincrease, and more small-scale features are ro-bustly imaged. We anticipate that full waveforminversion techniques will make particularly valu-able contributions to attenuation tomography.The replacement of source/receiver correctionfactors by more elaborate forward and inversemodeling techniques that correctly account for

Page 21: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

344 jeannot trampert and andreas fichtner

local structure would constitute a very signifi-cant progress. This is technically possible thanksto advanced numerical wave propagation tools for3D heterogeneous media (e.g. Moczo et al., 2007;Dumbser et al., 2007; Peter et al., 2011). Poten-tially, we can calculate the misfit kernels for allparameters simultaneously and update the com-plete model. Not all parameters will, of course,be equally well determined, and trade-offs exit.This approach will require a detailed resolutionanalysis. Tools in the framework of full waveforminversion are just becoming available (Fichtner &Trampert, 2011a, b).

While a simultaneous inversion for all possibleparameters is tempting, we must realize that notall Frechet derivatives are equally strong. Princi-pal component analysis (Sieminski et al., 2009)opens the opportunity to determine a priori thoselinear combinations of elastic parameters that arebest constrained by the data set, i.e., that generatethe largest sensitivities. For the elastic problem,not more than 6 linear combinations mostly ac-count for about 90% of the total sensitivity,meaning that all remaining linear combinationscan hardly be constrained. Thus, rather than up-dating all elastic parameters, only the first 6linear combinations of parameters could be up-dated instead. Principal component analysis is avaluable tool to optimize the design of an inverseproblem by finding the maximum number of well-constrained parameters. While the optimal linearcombinations of elastic parameters may not al-ways have a specific meaning in terms of wavepropagation, they can still be interpreted in amineral physics context, thereby providing robustinsight into the thermochemical and deformationstate of the Earth.

A last promising direction is the incorpora-tion of new observables, either in the form ofcompletely new measurements or as specificallydesigned misfit functionals. Examples of the for-mer include measurements of rotational groundmotions (e.g. Ferreira & Igel, 2009), Bernaueret al., 2009), Earth tides (Latychev et al., 2009)and highly accurate long-period mode splitting(Deuss et al., 2011). The design of targeted misfitfunctionals is based on the realization that not all

parametrizations are equivalent. It is well under-stood, for instance, that travel times are sensitiveto velocities, but hardly to density. If the problemis reformulated with elastic parameters and den-sity instead, the sensitivity to density increases,but now the problem has to be solved for elas-tic parameters and density simultaneously, withpossible trade-offs. In the proposed approach it isfurther important to realize that the sensitivitiesare determined by the definition of the misfitfunctional or the measurement. The idea is thento find a misfit functional which maximizes, oreven better, which diagonalizes the sensitivity tothe parameter of interest. This can be set up asa design problem in combination with principalcomponent analysis (Sieminski et al., 2009) orin a fully nonlinear fashion (van den Berg et al.,2003).

Most of these directions are certain to sig-nificantly increase our imaging capabilities. Inthe more distant future, sampling directly theposterior (i.e. solving Equation 11.5) includingthe full nonlinearity of the forward problem,should become feasible with exa-computing.Guided Monte Carlo or Neural Networks (e.g.,Meier et al., 2007) seem promising.

References

Anderson, D. L., 1961. Elastic wave propagation inlayered anisotropic media, J. Geophys. Res., 66,2953–2963.

Anderson, D. L., Ben-Menahem, A., & Archambeau, C.B., 1965. Attenuation of seismic energy in the uppermantle, J. Geophys. Res., 70, 1441–1448.

Anderson, D. L. & Hart, R. S., 1978. Q of the Earth, J.Geophys. Res., 83, 5869–5882.

Anderson, D. L. & Minster, J. B., 1979. The frequencydependence of Q in the Earth and implications formantle rheology and the Chandler wobble, Geophys.J. Roy. Astr. Soc., 58, 431–440.

Babuska, V., Montagner, J., Plomerov, J., & Girardin, N.,1998. Age-dependent large-scale fabric of the mantlelithosphere as derived from surface-wave velocityanisotropy, Pure and Applied Geophysics, 151(2-4),257–280.

Page 22: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

Global Imaging of the Earth’s Deep Interior 345

Backus, G. E., 1962. Long-wave elastic anisotropy pro-duced by horizontal layering, J. Geophys. Res., 67,4427–4440.

Bamberger, A., Chavent, G., Hemons, C., & Lailly, P.,1982. Inversion of normal incidence seismograms,Geophysics, 47, 757–770.

Becker, T. W., Ekstrom, G., Boschi, L., & Woodhouse,J. H., 2007. Length scales, patterns and origin ofazimuthal seismic anisotropy in the upper mantle asmapped by Rayleigh waves, Geophys. J. Int., 171(1),451–462.

Becker, T. W., Kustowski, B., & Ekstrom, G., 2008.Radial seismic anisotropy as a constraint for uppermantle rheology, Earth Planet. Sci. Lett., 267(1-2),213–227.

Beghein, C. & Trampert, J., 2003. Robust normal modeconstraints on inner-core anisotropy from modelspace search, Science, 299(5606), 552–555.

Bernauer, M., Fichtner, A., & Igel, H., 2009. Infer-ring Earth structure from combined measurementsof rotational and translational ground motions, Geo-physics, 74, WCD41–WCD47.

Bijwaard, H. & Spakman, W., 2000. Non-linear globalP-wave tomography by iterated linearized inversion,Geophys. J. Int., 141(1), 71–82.

Billien, M., Leveque, J.-J., & Trampert, J., 2000. Globalmaps of Rayleigh wave attenuation for periods be-tween 40 and 150 seconds, Geophys. Res. Lett., 27,3619–3622.

Bozdag, E. & Trampert, J., 2008. On crustal correctionsin surface wave tomography, Geophys. J. Int., 172(3),1066–1082.

Canas, J. A. & Mitchell, B. J., 1978. Lateral variationsof surface wave attenuation across the Pacific, Bull.Seis. Soc. Am., 68, 1637–1650.

Capdeville, Y., Guillot, L., & Marigo, J. J., 2010. 2Dnonperiodic homogenization to upscale elastic mediafor P-SV waves, Geophys. J. Int., 182, 903–922.

Chan, W. W. & Der, Z. A., 1988. Attenuation of multipleScS in various parts of the world, Geophys. J., 92,303–314.

Chen, P., Zhao, L., & Jordan, T. H., 2007. Full 3D to-mography for the crustal structure of the Los Angelesregion, Bull. Seismol. Soc. Am., 97, 1094–1120.

Cheng, H.-X. & Kennett, B. L. N., 2002. Frequencydependence of seismic wave attenuation in the uppermantle beneath the Australian region, Geophys. J.Int., pp. 45–57.

Chevrot, S., 2006. Finite-frequency vectorial tomogra-phy: A new method for high-resolution imaging ofupper mantle anisotropy, Geophys. J. Int., 165(2),641–657.

Dahlen, F., Hung, S.-H., & Nolet, G., 2000. Frechetkernels for finite-frequency traveltimes – I. Theory,Geophys. J. Int., 141, 157–174.

Dalton, C. A. & Ekstrom, G., 2006. Global modelsof surface wave attenuation, J. Geophys. Res., 111,doi:10.1029/2005JB003997.

Dalton, C. A., Ekstrom, G., & Dziewonski, A. M., 2008.The global attenuation structure of the upper mantle,J. Geophys. Res., 113, doi:10.1029/2007JB005429.

De Rujula, A., Glashow, S. L., Wilson, R. R., & Charpak,G., 1983. Neutrino exploration of the Earth, Phys.Rep., 99, 341–396.

de Wit, R., Trampert, J., & van der Hilst, R., 2012.Towards quantifying uncertainty in travel time to-mography using the null space shuttle, J. Geophys.Res., 117, B03301, doi:10.1029/2011JB008754.

Deal, M. M. & Nolet, G., 1996. Nullspace shuttles,Geophys. J. Int., 124(2), 372–380.

Deuss, A., Irving, J., & Woodhouse, J., 2010. Regionalvariation of inner core anisotropy from seismic nor-mal mode observations, Science, 328, 1018–1020.

Deuss, A., Ritsema, J., & van Heijst, H. J., 2011. Splittingfunction measurements for Earth’s longest-periodnormal-modes using recent earthquakes, Geophys.Res. Lett., 38, L04303, doi:10.1029/2010JB046115.

Dumbser, M., Kaser, M., & de la Puente, J., 2007. Ar-bitrary high-order finite volume schemes for seismicwave propagation on unstructured meshes in 2D and3D, Geophys. J. Int., 171, 665–694.

Durek, J. J. & Ekstrom, G., 1996. A radial model ofanelasticity consistent with long-period surface waveattenuation, Bull. Seis. Soc. Am., 86, 144–158.

Durek, J. J., Ritzwoller, M. H., & Woodhouse, J. H.,1993. Constraining upper mantle anelasticity usingsurface wave amplitude anomalies, Geophys. J. Int.,114, 249–272.

Dziewonski, A. M., 1984. Mapping the lower mantle:Determination of lateral heterogeneity in P veloc-ity up to degree and order 6, J. Geophys. Res., 89,5929–5952.

Dziewonski, A. M. & Anderson, D. L., 1981. Preliminaryreference Earth model, Phys. Earth Planet. Inter., 25,297–356.

Ekstrom, G., 2011. A global model of Love and Rayleighsurface-wave dispersion and anisotropy, 25-250 sec-onds, Geophys. J. Int., 187, 1668–1686.

Page 23: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

346 jeannot trampert and andreas fichtner

Ferreira, A. M. G. & Igel, H., 2009. Rotational motionsof seismic surface waves in a laterally heterogeneousEarth, Bull. Seis. Soc. Am., 99, 1429–1436.

Fichtner, A., Kennett, B. L. N., Igel, H., & Bunge, H.-P.,2009. Full seismic waveform tomography for upper-mantle structure in the Australasian region usingadjoint methods, Geophys. J. Int., 179, 1703–1725.

Fichtner, A., Kennett, B. L. N., Igel, H., & Bunge,H.-P., 2010. Full waveform tomography for radiallyanisotropic structure: New insights into present andpasts states of the Australasian upper mantle, EarthPlanet. Sci. Lett., 290, 270–280.

Fichtner, A. & Trampert, J., 2011a. Hessian kernelsof seismic data functionals based upon adjoint tech-niques, Geophys. J. Int., 185, 775–798.

Fichtner, A. & Trampert, J., 2011b. Resolution analysisin full waveform inversion, Geophys. J. Int., 187,1604–1664.

Flanagan, M. P. & Wiens, D. A., 1990. Attenuationstructure beneath the Lau backarc spreading centerfrom teleseismic S phases, Geophys. Res. Lett., 17,2117–2120.

Flanagan, M. P. & Wiens, D. A., 1998. Attenuationof broadband P and S waves in Tonga, Pure. Appl.Geophys., 153, 345–375.

Forte, A., 2007. Constraints on seismic models fromother disciplines – implications for mantle dynam-ics and composition, in: Treatise of Geophysics, 1,805–858.

Fukao, Y., Obayashi, M., & Nakakuki, T., 2009. Stag-nant slab: A review, Annu. Rev. Earth Planet. Sci.,37, 19–46.

Gonzalez-Garcia, M. C., Halzen, F., Maltoni, M., &Tanaka, H. K. M., 2008. Radiography of Earth’s coreand mantle with atmospheric neutrinos, Phys. Rev.Lett., 100, doi:10.1103/PhysRevLett.100.061802.

Grand, S., van der Hilst, R., & Widiyantoro, S., 1997.High resolution global tomography: a snapshot ofconvection in the Earth, GSA Today, 7, 1–7.

Gung, Y., Panning, M., & Romanowicz, B., 2003. Globalanisotropy and the thickness of continents, Nature,422(6933), 707–711.

Gung, Y. C. & Romanowicz, B., 2004. Q tomography ofthe upper mantle using three component long periodwaveforms, Geophys. J. Int., 157, 813–830.

Hess, H., 1964. Seismic anisotropy of the uppermostmantle under the oceans, Nature, 203, 629–631.

Igel, H., Djikpesse, H., & Tarantola, A., 1996. Wave-form inversion of marine reflection seismograms forP impedance and Poisson’s ratio, Geophys. J. Int.,124, 363–371.

Ishii, M. & Dziewonski, A. M., 2002. The innermostinner core of the Earth: Evidence for a change inanisotropic behavior at the radius of about 300 km,Proceedings of the National Academy of Sciences ofthe United States of America, 99(22), 14026–14030.

Ishii, M. & Tromp, J., 1999. Normal-mode and free-airgravity constraints on lateral variations in veloc-ity and density of Earth’s mantle, Science, 285,1231–1236.

Ishii, M. & Tromp, J., 2001. Even-degree lateral vari-ations in the Earth’s mantle constrained by freeoscillations and the free-air gravity anomaly, Geo-phys. J. Int., 145, 77–96.

Ishii, M. & Tromp, J., 2004. Constraining large-scalemantle heterogeneity using mantle and inner-coresensitive normal modes, Phys. Earth Planet. Int.,146, 113–124.

Jackson, I., 2000. Laboratory measurements of seismicwave dispersion and attenuation: Recent progress, In:Earth’s Deep Interior: mineral physics and tomogra-phy from the atomic to the global scale, GeophysicalMonograph, 117, 265–289.

Jackson, I., 2007. Physical origins of anelasticity and at-tenuation in rock, In: Treatise on Geophysics, editedby G. D. Price, 2, 496–522.

Kaminski, E., Ribe, N. M., & Browaeys, J. T., 2004. D-rex, a program for calculation of seismic anisotropydue to crystal lattice prefferred orientation in theconvective upper mantle, Geophys. J. Int., 158(2),744–752.

Karato, S.-I., 1993. Importance of attenuation in theinterpretation of seismic tomography, Geophys. Res.Lett., 20, 1623–1626.

Karato, S.-I., 2008. Deformation of Earth materials: Anintroduction to the rheology of solid Earth, Cam-bridge University Press, Cambridge.

Karato, S.-I., & Karki, B. B., 2001. Origin of lateral varia-tion of seismic wave velocities and density in the deepmantle, J. Geophys. Res., 106(B10), 21771–21783.

Kennett, B. L. N., 1998. On the density distributionwithin the Earth, Geophys. J. Int., 132, 374–382.

Kennett, B. L. N. & Abdullah, A., 2011. Seismic waveattenuation beneath the Australasian region, Austr.J. Earth Sci., 58, 285–295.

Kennett, B. L. N., Engdahl, E. R., & Buland, R., 1995.Constraints on seismic velocities in the Earth fromtraveltimes., Geophys. J. Int., 122, 108–124.

Komatitsch, D., Vinnik, L. P., & Chevrot, S., 2010.SHdiff-SVdiff splitting in an isotropic Earth, J. Geo-phys. Res., 115, B07312.

Page 24: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

Global Imaging of the Earth’s Deep Interior 347

Kuo, C. & Romanowicz, B., 2002. On the resolutionof density anomalies in the Earth’s mantle usingspectral fitting of normal-mode data, Geophys. J. Int.,150, 162–179.

Kustowski, B., Ekstrom, G., & Dziewonski, A. M.,2008. Anisotropic shear-wave velocity structure ofthe Earth’s mantle: a global model, J. Geophys. Res.,113, B06306, doi:10.1029/2007JB005169.

Latychev, K., Mitrovica, J. X., Ishii, M., Chan, N.-H., &Davis, J. L., 2009. Body tides on a 3D elastic Earth:Toward a tidal tomography, Earth Planet. Sci. Lett.,227, 86–90.

Lekic, V., Matas, J., Panning, M., & Romanowicz, B.,2009. Measurement and implications of frequencydependence of attenuation, Earth Planet. Sci. Lett.,282, 285–293.

Leveque, J. J., Rivera, L., & Wittlinger, G., 1993. Onthe use of the checkerboard test to assess the resolu-tion of tomographic inversions, Geophys. J. Int., 115,313–318.

Li, C., van der Hilst, R. D., Engdahl, E. R., & Burdick, S.,2008. A new global model for P wave speed variationsin Earth’s mantle, Geochem Geophys Geosys, 9(5),doi:10.1029/2007GC001806.

Liu, H.-P., Anderson, D. L., & Kanamori, H., 1976.Velocity dispersion due to anelasticity: implicationsfor seismology and mantle composition, Geophys. J.Roy. Astr. Soc., 47, 41–58.

Long, M. D., De Hoop, M. V., & van der Hilst, R. D.,2008. Wave-equation shear wave splitting tomogra-phy, Geophys. J. Int., 172(1), 311–330.

Long, M. D. & Silver, P. G., 2009. Shear wave splittingand mantle anisotropy: Measurements, interpreta-tions, and new directions, Surv Geophys, 30(4-5),407–461.

Masters, G., Jordan, T., Silver, P. G., & Gilbert, F.,1982. Aspherical Earth structure from fundamentalspheroidal mode data, Nature, 298, 609–613.

Masters, G., Laske, G., Bolton, H., & Dziewonski, A.,2000. The relative behavior of shear wave velocity,bulk sound speed, and compressional velocity in themantle: implications for chemical and thermal struc-ture, in Earth’s deep interior: mineral physics andtomography from the atomic to the global scale,pp. 63–87, ed. et al., S. K., American GeophysicalUnion, Washington.

Maupin, V., 1990. Modeling of three-component Lgwaves in anisotropic crustal models, Bull. Seismol.Soc. Am., 80(5), 1311–1325.

Maupin, V., Garnero, E. J., Lay, T., & Fouch, M. J., 2005.Azimuthal anisotropy in the D’’ layer beneath theCaribbean, J. Geophys. Res., 110(8), 1–20.

Maupin, V. & Park, J., 2007. Theory and observations- wave propagation in anisotropic media, in Treatiseon geophysics, Volume 1, pp. 289–321, ed. Schubert,G., Elsevier, Amsterdam.

McNamara, A., van Keken, P., & Karato, S.-I., 2002.Development of anisotropic structure by solid stateconvection in the Earth’s lower mantle, Nature, 416,310–314.

Megnin, C. & Romanowicz, B., 2000. The 3D shearvelocity structure of the mantle from the inversion ofbody, surface and higher modes wave forms, Geophys.J. Int., 143, 709–728.

Meier, U., Curtis, A., & Trampert, J., 2007. Fully non-linear inversion of fundamental mode surface wavesfor a global crustal model, Geophys. Res. Lett., 34,doi:10.1029/2007GL030989.

Meissner, R., Rabbel, W., & Kern, H., 2006. Seismiclamination and anisotropy of the lower continentalcrust, Tectonophysics, 416(1-2), 81–99.

Mitrovica, J. X. & Forte, A. M., 2004. A new inferenceof mantle viscosity based upon joint inversion of con-vection and glacial isostatic adjustment data, EarthPlanet. Sci. Lett., 225, 177–189.

Moczo, P., Kristek, J., Galis, M., Pazak, P., & Bala-zovjech, M., 2007. The finite-difference and finite-element modelling of seismic wave propagationand earthquake motion, Acta Physica Slovaca, 57,177–406.

Montagner, J.-P. & Tanimoto, T., 1991. Globalupper mantle tomography of seismic veloci-ties and anisotropies, J. Geophys. Res., 96(B12),20,337–20,351.

Morelli, A., Dziewonski, A.M., & Woodhouse, J.H.,1986. Anisotropy of the inner core inferred fromPKIKP travel times, Geophys. Res. Lett., 13,1545–1548.

Mosca, I., Cobden, L., Deuss, A., Ritsema, J., &Trampert, J., 2012. Seismic and mineralogicalstructures of the lower mantle from probabilis-tic tomography, J. Geophys. Res., 117, B06304,doi:10.1029/2011JB008851.

Nolet, G., 2008. A breviary of seismic tomography:Imaging the interior of the Earth and the Sun, Cam-bridge University Press, Cambridge.

Okal, E. A. & Jo, B.-G., 1990. Q measurements for phaseX overtones, Pure Appl. Geophys., 132, 331–362.

Page 25: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

348 jeannot trampert and andreas fichtner

Panning, M. P. & Nolet, G., 2008. Surface wave tomog-raphy for azimuthal anisotropy in a strongly reducedparameter space, Geophys. J. Int., 174(2), 629–648.

Panning, M. P. & Romanowicz, B., 2006. A three-dimensional radially anisotropic model of shear ve-locity in the whole mantle, Geophys. J. Int., 167,361–379.

Parker, R. L., 1994. Geophysical inverse theory, Prince-ton University Press.

Paulssen, H., 2004. Crustal anisotropy in southern Cal-ifornia from local earthquake data, Geophys. Res.Lett., 31(1), L01601 1–4.

Peter, D., Komatitsch, D., Luo, Y., Martin, R., Le Goff,N., Casarotti, E., Le Loher, P., Magnoni, F., Liu, Q.,Plitz, C., Nissen-Meyer, T., Basini, P., & Tromp, J.,2011. Forward and adjoint simulations of seismicwave propagation on fully unstructured hexahedralmeshes, Geophys. J. Int., 186, 721–739.

Rawlinson, N., Pogay, S., & Fishwick, S., 2010. Seismictomography: A window to the deep Earth, Phys. EarthPlanet. Int., 178, 101–135.

Reid, F. J. L., Woodhouse, J. H., & van Heijst, H. J.,2001. Upper mantle attenuation and velocity struc-ture from measurements of differential S phases,Geophys. J. Int., 145, 615–630.

Resovsky, J. S. & Ritzwoller, M. H., 1999. Regularisationuncertainty in density models estimated from normalmode data, Geophys. Res. Lett., 26, 2319–2322.

Resovsky, J. & Trampert, J., 2003. Using probabilisticseismic tomography to test mantle velocity-densityrelationships, Earth Planet. Sci. Lett., 215, 121–134.

Resovsky, J. S., Trampert, J., & van der Hilst, R. D.,2005. Error bars for the global seismic Q profile,Earth Planet. Sci. Lett., 230, 413–423.

Ritsema, J., Deuss, A., van Heijst, H. J., & Wood-house, J. H., 2011. S40RTS: a degree-40 shear-velocitymodel for the mantle from new Rayleigh wave disper-sion, teleseismic traveltime and normal-mode split-ting function measurements, Geophys. J. Int., 184,1223–1236.

Ritsema, J., McNamara, A. K., & Bull, A. L., 2007.Tomographic filtering of geodynamic models: Im-plications for models interpretation and large-scale mantle structure, J. Geophys. Res., 112(1),doi:10.1029/2006JB004566.

Ritzwoller, M. H. & Lavely, E. M., 1995. Three-dimensional seismic models of the Earth’s mantle,Rev. Geophys., 33, 1–66.

Romanowicz, B., 1990. The upper mantle degree 2:Constraints and inferences from global mantle wave

attenuation measurements, J. Geophys. Res., 95,11051–11071.

Romanowicz, B., 1994. On the measurement of anelas-tic attenuation using amplitudes of low-frequencysurface waves, Phys. Earth Planet. Int., 84, 179–191.

Romanowicz, B., 1995. A global tomographic model ofshear attenuation in the upper mantle, J. Geophys.Res., 100, 12375–12394.

Romanowicz, B., 2001. Can we resolve 3D density het-erogeneity in the lower mantle?, Geophys. Res. Lett.,28, 1107–1110.

Romanowicz, B., 2003. Global mantle tomogra-phy:progress status in the past 10 years, Ann. Rev.Earth Planet. Sci., 31, 303–328.

Sailor, R. V. & Dziewonski, A. M., 1978. Measurementsand interpretation of normal mode attenuation, Geo-phys. J. Roy. Astr. Soc., 53, 559–581.

Sambridge, M., 1999a. Geophysical inversion with aneighbourhood algorithm - I. Searching a parameterspace, Geophys. J. Int., 138, 479–494.

Sambridge, M., 1999b. Geophysical inversion with aneighbourhood algorithm - I. Appraising the ensem-ble, Geophys. J. Int., 138, 727–746.

Sambridge, M. & Mosegaard, K., 2002. Monte Carlomethods in geophysical inverse problems, Rev. Geo-phys., 40, doi: 10.1029/2000RG000089.

Schulte-Pelkum, V., Masters, G., & Shearer, P. M., 2001.Upper mantle anisotropy from long-period P polari-sation, J. Geophys. Res., 106(B10), 21917–21934.

Selby, N. D. & Woodhouse, J. H., 2000. Controls onRayleigh wave amplitudes: Attenuation and focusing,Geophys. J. Int., 142, 933–940.

Selby, N. D. & Woodhouse, J. H., 2002. The Qstructure of the upper mantle: Constraints fromRayleigh wave amplitudes, J. Geophys. Res., 107,doi:10.1029/2001JB000257.

Sheehan, A. F. & Solomon, S. C., 1992. Differentialshear-wave attenuation and its lateral variation inthe North Atlantic region, J. Geophys. Res., 97,15339–15350.

Sieminski, A., Trampert, J., & Tromp, J., 2009.Principal component analysis of anisotropic finite-frequency sensitivity kernels, Geophys. J. Int., 179(2),1186–1198.

Sigloch, K., 2008. Multiple-frequency body wave to-mography, Phd thesis, Princeton University.

Simmons, N. A., Forte, A. M., Boschi, L., & Grand, S. P.,2010. GyPSuM: A joint tomographic model of mantledensity and seismic wave speeds, J. Geophys. Res.,115, doi:10.1029/2010JB007631.

Page 26: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

Global Imaging of the Earth’s Deep Interior 349

Sipkin, S. A. & Jordan, T. H., 1979. Frequency-dependence of QScS, Bull. Seis. Soc. Am., 69,1055–1079.

Sipkin, S. A. & Revenaugh, J., 1994. Regional variationof attenuation and traveltime in China from anal-ysis of multiple ScS phases, J. Geophys. Res., 99,2687–2699.

Smith, G. P. & Ekstrom, G., 1999. A global study ofPn anisotropy beneath continents, J. Geophys. Res.,104(B1), 963–980.

Soldati, G., Boschi, L., & Piersanti, A., 2006. Globalseismic tomography and modern parallel computers,Annals of Geophysics, 49(4-5), 977–986.

Stacey, F. D. & Davis, P. M., 2004. High pressureequations of state with applications to the lowermantle and core, Phys. Earth Planet. Inter., 142(3-4),137–184.

Stachnik, J. C., Abers, G. A., & Christensen, D. H., 2004.Seismic attenuation and mantle wedge temperaturesin the Alaska subduction zone, J. Geophys. Res., 109,doi:10.1029/2004JB003018.

Stixrude, L. & Lithgow-Bertelloni, C., 2005. Thermo-dynamics of mantle minerals–I. Physical properties,GJI, 162(2), 610–632.

Stixrude, L. & Lithgow-Bertelloni, C., 2011. Thermo-dynamics of mantle minerals–II. Phase equilibria,Geophys. J. Int., 184(3), 1180–1213.

Tanaka, H., Nagamine, K., Kawamura, N., Nakamura, S.N., Ishida, K., & Shimomura, K., 2003. Developmentof a two-fold segmented detection system for nearhorizontally cosmic-ray muons to probe the internalstructure of a volcano, Nucl. Instr. Meth. A, 507,657–669.

Tanaka, H., Taira, H., Uchida, T., Tanaka, M., Takeo,M., Ohminato, T., Aoki, Y., Nishimata, R., Shoji, D.,& Tsuiji, H., 2010. Three-dimensional computationalaxial tomography scan of a volcano with cosmic raymuon radiography, J. Geophys. Res., 115, B12332,doi:10.1019/2010JB007677.

Tape, C., Liu, Q., Maggi, A., & Tromp, J., 2010. Seismictomography of the southern California crust basedupon spectral-element and adjoint methods., Geo-phys. J. Int., 180, 433–462.

Tarantola, A., 1986. A strategy for nonlinear elasticinversion of seismic reflection data, Geophysics, 51,1893–1903.

Tarantola, A., 1988. Theoretical background for theinversion of seismic waveforms, including elasticityand attenuation, Pure Appl. Geophys., 128, 365–399.

Tarantola, A., 2005. Inverse problem theory and meth-ods for model parameter estimation, Society forIndustrial and Applied Mathematics, Philadelphia.

Tian, Y., Sigloch, K., & Nolet, G., 2009. Multiple-frequency SH-wave tomography for the western USupper mantle, Geophys. J. Int., 178, 1384–1402.

Tondi, R., Achauer, U., Landes, M., davi, R., & Besutiu,L., 2009. Unveiling seismic and density structurebeneath the Vrancea seismogenic zone, Romania, J.Geophys. Res., 114, doi:10.1029/2008JB005992.

Tondi, R., de Franco, R., & Barzaghi, R., 2000. Sequentialinversion of refraction and wide-angle reflection trav-eltimes and gravity data for two-dimensional velocitystructures, Geophys. J. Int., 141, 679–698.

Trampert, J., Deschamps, F., Resovsky, J., & Yuen, D.,2004. Probabilistic tomography maps chemical het-erogeneities throughout the lower mantle., Science,306, 853–856.

Trampert, J., Vacher, P., & Vlaar, N., 2001. Sensitivitiesof seismic velocities to temperature, pressure andcomposition in the lower mantle, Phys. Earth Planet.Inter., 124(3-4), 255–267.

Trampert, J. & van der Hilst, R., 2005. Towards a quan-titative interpretation of global seismic tomography,in Earth’s deep mantle: Structure, composition andevolution, pp. 47–62, eds van der Hilst, R., Bass,J., Matas, J., & Trampert, J., American GeophysicalUnion, Washington.

Trampert, J. & van Heijst, H. J., 2002. Global az-imuthal anisotropy in the transition zone, Science,296, 1297–1299.

Trampert, J. & Woodhouse, J. H., 2003. Globalanisotropic phase velocity maps for fundamentalmode surface waves between 40 and 150 s, Geophys.J. Int., 154(1), 154–165.

van den Berg, J., Curtis, A., & Trampert, J., 2003. Op-timal nonlinear bayesian experimental design: Anapplication to amplitude versus offset experiments,Geophys. J. Int., 155(2), 411–421.

van der Hilst, R. D., Widiyantoro, S., & Engdahl, E.R., 1997. Evidence for deep mantle circulation fromglobal tomography, Nature, 386, 578–584.

Visser, K., Trampert, J., Lebedev, S., & Kennett, B. L.N., 2008. Probability of radial anisotropy in the deepmantle, Earth Planet. Sci. Lett., 270(3-4) 241–250.

Warren, L. M. & Shearer, P. M., 2002. Mappinglateral variations in upper mantle attenuation bystacking P and PP spectra, J. Geophys. Res., 107,doi:10.1029/2001JB001195.

Page 27: 11 Global Imaging of the Earth’s Deep Interior: Seismic ...

350 jeannot trampert and andreas fichtner

Wenk, H., Cottaar, S., Tome, C. N., McNamara, A.,& Romanowicz, B., 2011. Deformation in the lower-most mantle: From polycrystal plasticity to seismicanisotropy, Earth Planet. Sci. Lett., 306(1-2), 33–45.

Widmer, R., Masters, G., & Gilbert, F., 1991. Sphericallysymmetric attenuation within the Earth from normalmode data, Geophys. J. Int., 104, 541–553.

Woodhouse, J. H. & Dziewonski, A. M., 1984. Mappingthe upper mantle: Three-dimensional modeling ofEarth structure by inversion of seismic waveforms, J.Geophys. Res., 89, 5953–5986.

Woodhouse, J. H. & Dziewonski, A. M., 1989. Seismicmodelling of the Earth’s large-scale three dimen-sional structure, Philos. Trans. R. Soc. Lond. A, 328,291–308.

Woodhouse, J. H., Giardini, D., & Li, X. D., 1986. Evi-dence for inner core anisotropy from free oscillations,Geophys. Res. Lett., 13, 1549–1552.

Woodhouse, J. H. & Wong, Y. K., 1986. Amplitude,phase and path anomalies of mantle waves, Geophys.J. Roy. Astr. Soc., 87, 753–773.

Wu, R. & Aki, K., 1985. Scattering characteristics ofelastic waves by an elastic heterogeneity, Geophysics,50, 582–595.

Zener, C., 1948. Elasticity and anelasticity of metals,University of Chicago Press, Chicago.

Zhou, Y., 2009. Surface-wave sensitivity to 3-D anelas-ticity, Geophys. J. Int., 178, 1403–1410.