ArticlePDF Available

An overview of full-waveform inversion in exploration geophysics

Authors:
  • University of Nice-Sophia Antipolis - Observatoire de la Cote d'Azur - CNRS - IRD

Abstract and Figures

Full-waveform inversion (FWI) is a challenging data-fitting procedure based on full-wavefield modeling to extract quantitative information from seismograms. High-resolution imaging at half the propagated wavelength is expected. Recent advances in high-performance computing and multifold/multicomponent wide-aperture and wide-azimuth acquisitions make 3D acoustic FWI feasible today. Key ingredients of FWI are an efficient forward-modeling engine and a local differential approach, in which the gradient and the Hessian operators are efficiently estimated. Local optimization does not, however, prevent convergence of the misfit function toward local minima because of the limited accuracy of the starting model, the lack of low frequencies, the presence of noise, and the approximate modeling of the wave-physics complexity. Different hierarchical multiscale strategies are designed to mitigate the nonlinearity and ill-posedness of FWI by incorporating progressively shorter wavelengths in the parameter space. Synthetic and real-data case studies address reconstructing various parameters, from V(P) and V(S) velocities to density, anisotropy, and attenuation. This review attempts to illuminate the state of the art of FWI. Crucial jumps, however, remain necessary to make it as popular as migration techniques. The challenges can be categorized as (1) building accurate starting models with automatic procedures and/or recording low frequencies, (2) defining new minimization criteria to mitigate the sensitivity of FWI to amplitude errors and increasing the robustness of FWI when multiple parameter classes are estimated, and (3) improving computational efficiency by data-compression techniques to make 3D elastic FWI feasible.
Content may be subject to copyright.
An overview of full-waveform inversion in exploration geophysics
J. Virieux1and S. Operto2
ABSTRACT
Full-waveform inversion FWIis a challenging data-fitting
procedure based on full-wavefield modeling to extract quantita-
tive information from seismograms. High-resolution imaging at
half the propagated wavelength is expected. Recent advances in
high-performance computing and multifold/multicomponent
wide-aperture and wide-azimuth acquisitions make 3D acoustic
FWI feasible today. Key ingredients of FWI are an efficient for-
ward-modeling engine and a local differential approach, in
which the gradient and the Hessian operators are efficiently esti-
mated. Local optimization does not, however, prevent conver-
gence of the misfit function toward local minima because of the
limited accuracy of the starting model, the lack of low frequen-
cies, the presence of noise, and the approximate modeling of the
wave-physics complexity. Different hierarchical multiscale
strategies are designed to mitigate the nonlinearity and ill-posed-
ness of FWI by incorporating progressively shorter wavelengths
in the parameter space. Synthetic and real-data case studies ad-
dress reconstructing various parameters, from VPand VSveloci-
ties to density, anisotropy, and attenuation. This review attempts
to illuminate the state of the art of FWI. Crucial jumps, however,
remain necessary to make it as popular as migration techniques.
The challenges can be categorized as 1building accurate start-
ing models with automatic procedures and/or recording low fre-
quencies, 2defining new minimization criteria to mitigate the
sensitivity of FWI to amplitude errors and increasing the robust-
ness of FWI when multiple parameter classes are estimated, and
3improving computational efficiency by data-compression
techniques to make 3D elastic FWI feasible.
INTRODUCTION
Seismic waves bring to the surface information gathered on the
physical properties of the earth. Since the discovery of modern seis-
mology at the end of the 19th century, the main discoveries have aris-
en from using traveltime information Oldham, 1906;Gutenberg,
1914;Lehmann, 1936. Then there was a hiatus until the 1980s for
amplitude interpretation, when global seismic networks could pro-
vide enough calibrated seismograms to compute accurate synthetic
seismograms using normal-mode summation. Differential seismo-
grams estimated through the Born approximation have been used
as perturbations for matching long-period seismograms, which can
provide high-resolution upper-mantle tomography Gilbert and Dz-
iewonski, 1975;Woodhouseand Dziewonski, 1984. The sensitivity
or Fréchet derivative matrix, i.e., the partial derivative of seismic
data with respect to the model parameters, is explicitly estimated be-
fore proceeding to inversion of the linearized system. The normal-
mode description allows a limited number of parameters to be in-
verted a few hundred parameters, which makes the optimization
procedure feasible through explicit sensitivity matrix estimation in
spite of the high number of seismograms.
Meanwhile, exploration seismology has taken up the challenge of
high-resolution imaging of the subsurface by designing dense, mul-
tifold acquisition systems. Construction of the sensitivity matrix is
too prohibitive because the number of parameters exceed 10,000. In-
stead, another road has been taken to perform high-resolution imag-
ing. Using the exploding-reflector concept, and after some kinemat-
ic corrections, amplitude summation has provided detailed images
of the subsurface for reservoir determination and characterization
Claerbout, 1971,1976. The sum of the traveltimes from a specific
point of the interface toward the source and the receiver should coin-
cide with the time of large amplitudes in the seismogram. The reflec-
tivity as an amplitude attribute of related seismic traces at the select-
ed point of the reflector provides the migrated image needed for seis-
mic stratigraphic interpretation. Although migration is more a con-
cept for converting seismic data recorded in the time-space domain
Manuscript received by the Editor 7 June 2009; revised manuscript received 8 July 2009; published online 3 December 2009.
1Université Joseph Fourier, Laboratoire de Géophysique Interne et Tectonophysique, CNRS, IRD, Grenoble, France. E-mail: Jean.Virieux@obs.ujf-
grenoble.fr.
2Université Nice-Sophia Antipolis, Géoazur, CNRS, IRD, Observatoire de la Côte d’Azur, Villefranche-sur-mer,France. E-mail: operto@geoazur.obs-vlfr.fr.
© 2009 Society of Exploration Geophysicists. All rights reserved.
GEOPHYSICS, VOL. 74, NO. 6 NOVEMBER-DECEMBER 2009; P. WCC127–WCC152, 15 FIGS., 1TABLE.
10.1190/1.3238367
WCC127
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
into images of physical properties, we often refer to it as the geomet-
ric description of the short wavelengths of the subsurface. Avelocity
macromodel or background model provides the kinematic informa-
tion required to focus waves inside the medium.
The limited offsets recorded by seismic reflection surveys and the
limited-frequency bandwidth of seismic sources make seismic im-
aging poorly sensitive to intermediate wavelengths Jannane et al.,
1989. This is the motivation behind a two-step workflow: construct
the macromodel using kinematic information, and then the ampli-
tude projection using different types of migrations Claerbout and
Doherty, 1972;Gazdag, 1978;Stolt, 1978;Baysal et al., 1983;Yil-
maz, 2001;Biondi and Symes, 2004. This procedure is efficient for
relatively simple geologic targets in shallow-water environments,
although more limited performances have been achieved for imag-
ing complex structures such as salt domes, subbasalt targets, thrust
belts, and foothills. In complex geologic environments, building an
accurate velocity background model for migration is challenging.
Various approaches for iterative updating of the macromodel recon-
struction have been proposed Snieder et al., 1989;Docherty et al.,
2003, but they remain limited by the poor sensitivity of the reflec-
tion seismic data to the large and intermediate wavelengths of the
subsurface.
Simultaneous with the global seismology inversion scheme,
Lailly 1983and Tarantola 1984recast the migration imaging
principle of Claerbout 1971,1976as a local optimization problem,
the aim of which is least-squares minimization of the misfit between
recorded and modeled data. They show that the gradient of the misfit
function along which the perturbation model is searched can be built
by crosscorrelating the incident wavefield emitted from the source
and the back-propagated residual wavefields. The perturbation mod-
el obtained after the first iteration of the local optimization looks like
a migrated image obtained by reverse-time migration. One differ-
ence is that the seismic wavefield recorded at the receiver is back
propagated in reverse time migration, whereas the data misfit is back
propagated in the waveform inversion of Lailly 1983and Tarantola
1984. When added to the initial velocity, the velocity perturbations
lead to an updated velocity model, which is used as a starting model
for the next iteration of minimizing the misfit function. The impres-
sive amount of data included in seismograms each sample of a time
series must be consideredis involved in gradient estimation. This
estimation is performed by summation over sources, receivers, and
time.
Waveform-fitting imaging was quite computer demanding at that
time, even for 2D geometries Gauthier et al., 1986. However, it has
been applied successfully in various studies using forward-model-
ing techniques such as reflectivity techniques in layered media Kor-
mendi and Dietrich, 1991, finite-difference techniques Kolb et al.,
1986;Ikelle et al., 1988;Crase et al., 1990;Pica et al., 1990;Djik-
péssé and Tarantola, 1999, finite-element methods Choi et al.,
2008, and extended ray theory Cary and Chapman, 1988;Koren et
al., 1991;Sambridge and Drijkoningen, 1992. A less computation-
ally intensive approach is achieved by Jin et al. 1992and Lambaré
et al. 1992, who establish the theoretical connection between ray-
based generalized Radon reconstruction techniques Beylkin, 1985;
Bleistein, 1987;Beylkin and Burridge, 1990and least-squares opti-
mization Tarantola, 1987. By defining a specific norm in the data
space, which varies from one focusing point to the next, they were
able to recast the asymptotic Radon transform as an iterative least-
squares optimization after diagonalizing the Hessian operator. Ap-
plications on 2D synthetic data and real data are provided Thierry et
al., 1999b;Operto et al., 2000and 3D extension is possible Thierry
et al., 1999a;Operto et al., 2003because of efficient asymptotic for-
ward modeling Lucio et al., 1996. Because the Green’s functions
are computed in smoothed media with the ray theory, the forward
problem is linearized with the Born approximation, and the optimi-
zation is iterated linearly, which means the background model re-
mains the same over the iterations. These imaging methods are gen-
erally called migration/inversion or true-amplitude prestack depth
migration PSDM. The main difference with the waveform-inver-
sion methods we describe is that the smooth background model does
not change over iterations and only the single scattered wavefield is
modeled by linearizing the forward problem.
Alternatively, the full information content in the seismogram can
be considered in the optimization. This leads us to full-waveform in-
version FWI, where full-wave equation modeling is performed at
each iteration of the optimization in the final model of the previous
iteration. All types of waves are involved in the optimization, includ-
ing diving waves, supercritical reflections, and multiscattered waves
such as multiples. The techniques used for the forward modeling
vary and include volumetric methods such as finite-element meth-
ods Marfurt, 1984;Min et al., 2003, finite-difference methods
Virieux, 1986, finite-volume methods Brossier et al., 2008, and
pseudospectral methods Danecek and Seriani, 2008; boundary in-
tegral methods such as reflectivity methods Kennett, 1983; gener-
alized screen methods Wu, 2003; discrete wavenumber methods
Bouchon et al., 1989; generalized ray methods such as WKBJ and
Maslov seismograms Chapman, 1985; full-wave theory de Hoop,
1960; and diffraction theory Klem-Musatov and Aizenberg, 1985.
FWI has not been recognized as an efficient seismic imaging tech-
nique because pioneering applications were restricted to seismic re-
flection data. For short-offset acquisition, the seismic wavefield is
rather insensitive to intermediate wavelengths; therefore, the opti-
mization cannot adequately reconstruct the true velocity structure
through iterative updates. Only when a sufficiently accurate initial
model is provided can waveform-fitting converge to the velocity
structure through such updates. For sampling the initial model, so-
phisticated investigations with global and semiglobal techniques
Koren et al., 1991;Jin and Madariaga, 1993,1994;Mosegaard and
Tarantola, 1995;Sambridge and Mosegaard, 2002have been per-
formed. The rather poor performance of these investigations that
arises from insensitivity to intermediate wavelengths has led many
researchers to believe that this optimization technique is not particu-
larly efficient.
Only with the benefit of long-offset and transmission data to re-
construct the large and intermediate wavelengths of the structure has
FWI reached its maturity as highlighted by Mora 1987,1988,Pratt
and Worthington 1990,Pratt et al. 1996, and Pratt 1999. FWI at-
tempts to characterize a broad and continuous wavenumber spec-
trum at each point of the model, reunifying macromodel building
and migration tasks into a single procedure. Historical crosshole and
wide-angle surface data examples illustrate the capacity of simulta-
neous reconstruction of the entire spatial spectrum e.g., Pratt, 1999;
Ravaut et al., 2004;Brenders and Pratt, 2007a. However, robust ap-
plication of FWI to long-offset data remains challenging because of
increasing nonlinearities introduced by wavefields propagated over
several tens of wavelengths and various incidence angles Sirgue,
2006.
Here, we consider the main aspects of FWI. First, we review the
forward-modeling problem that underlies FWI. Efficient numerical
WCC128 Virieux and Operto
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
modeling of the full seismic wavefield is a central issue in FWI, es-
pecially for 3D problems.
In the second part, we review the main theoretical aspects of FWI
based on a least-squares local optimization approach. We follow the
compact matrix formalism for its simplicity Pratt et al., 1998;Pratt,
1999, which leads to a clear interpretation of the gradient and the
Hessian of the objective function. Once the gradient is estimated, we
review different optimization algorithms used to compute the pertur-
bation model. We conclude the methodology section by the source
estimation problem in FWI.
In the third part, we review some key features of FWI. First, we
highlight the relationships between the experimental setup source
bandwidth, acquisition geometryand the spatial resolution of FWI.
The resolution analysis provides the necessary guidelines to design
the multiscale FWI algorithms required to mitigate the nonlinearity
of FWI. We discuss the pros and cons of the time and frequency do-
mains for efficient multiscale algorithms. We provide a few words
concerning the parallel implementation of FWI techniques because
these are computer demanding. Then we review some alternatives to
the least-squares criterion and the Born linearization. Akey issue of
FWI is the initial model from which the local optimization is started.
We also discuss several tomographic approaches to building a start-
ing model.
In the fourth part, we review the main case studies of FWI subdi-
vided into three categories of case studies: acoustic, multiparameter,
and three dimensional. Finally, we discuss the future challenges
raised by the revival of interest in FWI that has been shown by the
exploration and the earthquake-seismology communities.
THE FORWARD PROBLEM
Let us first introduce the notations for the forward problem, name-
ly, modeling the full seismic wavefield. The reader is referred to
Robertsson et al. 2007for an up-to-date series of publications on
modern seismic-modeling methods.
We use matrix notations to denote the partial-differential opera-
tors of the wave equation Marfurt, 1984;Carcione et al., 2002. The
most popular direct method to discretize the wave equation in the
time and frequency domains is the finite-difference method
Virieux, 1986;Levander, 1988;Graves, 1996;Operto et al., 2007,
although more sophisticated finite-element or finite-volume ap-
proaches can be considered. This is especially true when accurate
boundary conditions through unstructured meshes must be imple-
mented e.g., Komatitsch and Vilotte, 1998;Dumbser and Kaser,
2006.
In the time domain, we have
Mxd2ux,t
dt2Axux,tsx,t,1
where Mand Aare the mass and the stiffness matrices, respectively
Marfurt, 1984. The source term is denoted by sand the seismic
wavefield by u. In the acoustic approximation, ugenerally repre-
sents pressure, although in the elastic case ugenerally represents
horizontal and vertical particle velocities. The time is denoted by t
and the spatial coordinates by x. Equation 1generally is solved with
an explicit time-marching algorithm: The value of the wavefield at a
time step n1at a spatial position is inferred from the value of the
wavefields at previous time steps. Implicit time-marching algo-
rithms are avoided because they require solving a linear system
Marfurt, 1984. If both velocity and stress wavefields are helpful,
the system of second-order equations can be recast as a first-order
hyperbolic velocity-stress system by incorporating the necessary
auxiliary variables Virieux, 1986.
In the frequency domain, the wave equation reduces to a system of
linear equations; the right-hand side is the source and the solution is
the seismic wavefield. This system can be written compactly as
Bx,
ux,
sx,
,2
where Bis the impedance matrix Marfurt, 1984. The sparse com-
plex-valued matrix Bhas a symmetric pattern, although it is not sym-
metric because of absorbing boundary conditions Hustedt et al.,
2004;Operto et al., 2007.
Equation 2can be solved by a decomposition of Bsuch as lower
and upper LUtriangular decomposition, leading to direct-solver
techniques. The advantage of the direct-solver approach is that once
the decomposition is performed, equation 2is efficiently solved for
multiple sources using forward and backward substitutions Mar-
furt, 1984. The direct-solver approach is efficient for 2D forward
problems Jo et al., 1996;Stekl and Pratt, 1998;Hustedt et al., 2004.
However, the time and memory complexities of LU factorization
and its limited scalability on large-scale distributed memory plat-
forms prevent use of the approach for large-scale 3D problems i.e.,
problems involving more than 10 million unknowns; Operto et al.,
2007.
Iterative solvers provide an alternative approach for solving the
time-harmonic wave equation Riyanti et al., 2006,2007;Plessix,
2007;Erlangga and Herrmann, 2008. Iterative solvers currently are
implemented with Krylov subspace methods Saad, 2003that are
preconditioned by solving the dampened time-harmonic wave equa-
tion. The solution of the dampened wave equation is computed with
one cycle of a multigrid. The main advantage of the iterative ap-
proach is the low memory requirement, although the main drawback
results from a difficulty to design an efficient preconditioner because
the impedance matrix is indefinite. To our knowledge, the extension
to elastic wave equations still needs to be investigated. As for the
time-domain approach, the time complexity of the iterative ap-
proach increases linearly with the number of sources in contrast to
the direct-solver approach.
An intermediate approach between the direct and iterative meth-
ods consists of a hybrid direct-iterative approach based on a domain
decomposition method and the Schur complement system Saad,
2003;Sourbier et al., 2008. The iterative solver is used to solve the
reduced Schur complement system, the solution of which is the
wavefield at interface nodes between subdomains. The direct solver
is used to factorize local impedance matrices that are assembled on
each subdomain. Briefly, the hybrid approach provides a compro-
mise in terms of memory savings and multisource-simulation effi-
ciency between the direct and the iterative approaches.
The last possible approach to compute monochromatic wave-
fields is to perform the modeling in the time domain and extract the
frequency-domain solution either by discrete Fourier transform in
the loop over the time steps Sirgue et al., 2008or by phase-sensitiv-
ity detection once the steady-state regime is reached Nihei and Li,
2007. One advantage of the approach based on the discrete Fourier
transform is that an arbitrary number of frequencies can be extracted
within the loop over time steps at minimal extra cost. Second, time
windowing can be easily applied, which is not the case when the
modeling is performed in the frequency domain. Time windowing
allows the extraction of specific arrivals for FWI early arrivals, re-
flections, PS converted waves, which is often useful to mitigate the
Full-waveform inversion WCC129
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
nonlinearity of the inversion by judicious data preconditioning
Sears et al., 2008;Brossier et al., 2009a.
Among all of these possible approaches, the iterative-solver ap-
proach theoretically has the best time complexity here, “complexi-
ty” denotes how the computational cost of an algorithm grows with
the size of the computational domainif the number of iterations is
independent of the frequency Erlangga and Herrmann, 2008.In
practice, the number of iterations generally increases linearly with
frequency. In this case, the time complexities of the time-domain and
the iterative-solver approach are equivalent Plessix, 2007.
The reader is referred to Plessix 2007,2009and Virieux et al.
2009for more detailed complexity analyses of seismic modeling
based on different numerical approaches. A discussion on the pros
and cons of time-domain versus frequency-domain seismic model-
ing with application to FWI is also provided in Vigh and Starr
2008band Warner et al. 2008.
Source implementation is an important issue in FWI. The spatial
reciprocity of Green’s functions can be exploited in FWI to mitigate
the number of forward problems if the number of receivers is signifi-
cantly smaller than the number of sources Aki and Richards, 1980.
The reciprocity of Green’s functions also allows matching data emit-
ted by explosions and recorded by directional sensors, with pressure
synthetics computed for directional forces Operto et al., 2006.Of
note, the spatial reciprocity is satisfied theoretically for the unidirec-
tional sensor and the unidirectional impulse source. However, the
spatial reciprocity of the Green’s functions can also be used for ex-
plosive sources by virtue of the superposition principle. Indeed, ex-
plosions can be represented by double dipoles or, in other words, by
four unidirectional impulse sources.
A final comment concerns the relationship between the discretiza-
tion required to solve the forward problem and that required to re-
construct the physical parameters. Often during FWI, these two dis-
cretizations are identical, although it is recommended that the fin-
gerprint of the forward problem be kept minimal in FWI.
The properties of the subsurface that we want to quantify are em-
bedded in the coefficients of matrices M,A,orBof equations 1and
2. The relationship between the seismic wavefield and the parame-
ters is nonlinear and can be written compactly through the operator
G, defined as
uGm兲共3
in the time domain or in the frequency domain.
FWI AS A LEAST-SQUARES LOCAL
OPTIMIZATION
We follow the simplest view of FWI based on the so-called length
method Menke, 1984. For information on probabilistic maximum
likelihood or generalized inverse formulations, the reader is referred
to Menke 1984,Tarantola 1987,Scales and Smith 1994, and
Sen and Stoffa 1995.
We define the misfit vector ddobsdcalmof dimension N
by the differences at the receiver positions between the recorded
seismic data dobs and the modeled seismic data dcalmfor each
source-receiver pair of the seismic survey. Here, dcal can be related to
the modeled seismic wavefield uby a detection operator R, which
extracts the values of the wavefields computed in the full computa-
tional domain at the receiver positions for each source: dcal Ru.
The model mrepresents some physical parameters of the subsurface
discretized over the computational domain.
In the simplest case corresponding to the monoparameter acoustic
approximation, the model parameters are the P-wave velocities de-
fined at each node of the numerical mesh used to discretize the in-
verse problem. In the extreme case, the model parameters corre-
spond to the 21 elastic moduli that characterize linear triclinic elastic
media, the density, and some memory variables that characterize the
anelastic behavior of the subsurface Toksöz and Johnston, 1981.
The most common discretization consists of projection of the con-
tinuous model of the subsurface on a multidimensional Dirac comb,
although a more complex basis can be considered see Appendix A
in Pratt et al. 1998for a discussion on alternative parameteriza-
tions. We define a norm Cmof this misfit vector d, which is re-
ferred to as the misfit function or the objective function. We focus be-
low on the least-squares norm, which is easier to manipulate mathe-
matically Tarantola, 1987. Other norms are discussed later.
The Born approximation and the linearization of the
inverse problem
The least-squares norm is given by
Cm1
2dd,4
where denotes the adjoint operator transpose conjugate.
In the time domain, the implicit summation in equation 4is per-
formed over the number of source-channel pairs and the number of
time samples in the seismograms, where a channel is one component
of a multicomponent sensor. In the frequency domain, the summa-
tion over frequencies replaces that over time. In the time domain, the
misfit vector is real valued; in the frequency domain, it is complex
valued.
The minimum of the misfit function Cmis sought in the vicinity
of the starting model m0. The FWI is essentially a local optimization.
In the framework of the Born approximation, we assume that the up-
dated model mof dimension Mcan be written as the sum of the start-
ing model m0plus a perturbation model m:mm0m.Inthe
following, we assume that mis real valued.
A second-order Taylor-Lagrange development of the misfit func-
tion in the vicinity of m0gives the expression
Cm0mCm0
j1
M
Cm0
mj
mj
1
2
j1
M
k1
M
2Cm0
mj
mk
mjmkOm3.
5
Taking the derivative with respect to the model parameter mlresults
in
Cm
ml
Cm0
ml
j1
M
2Cm0
mj
ml
mj.6
The minimum of the misfit function in the vicinity of point m0is
reached when the first derivative of the misfit function vanishes. This
gives the perturbation model vector:
m⳱ⳮ
2Cm0
m2
1
Cm0
m.7
WCC130 Virieux and Operto
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
The perturbation model is searched in the opposite direction of the
steepest ascent i.e., the gradientof the misfit function at point m0.
The second derivative of the misfit function is the Hessian; it defines
the curvature of the misfit function at m0. Of note, the error term
Om3in equation 5is zero when the misfit function is a quadratic
function of m. This is the case for linear forward problems such as u
G.m. In this case, the expression of the perturbation model of
equation 7gives the minimum of the misfit function in one iteration.
In FWI, the relationship between the data and the model is nonlinear
and the inversion needs to be iterated several times to converge to-
ward the minimum of the misfit function.
Normal equations: The Newton, Gauss-Newton, and
steepest-descent methods
Basic equations
The derivative of Cmwith respect to the model parameter ml
gives
Cm
ml
⳱ⳮ
1
2
i1
N
dcali
ml
dobsi
dcali*
dobsi
dcali
dcali
*
ml
⳱ⳮ
i1
N
R
dcali
ml
*
dobsi
dcali
,8
where the real part and the conjugate of a complex number are denot-
ed by Rand *, respectively. In matrix form, equation 8translates to
Cm
Cm
m
⳱ⳮR
dcalm
m
dobsdcalm兲兲
⳱ⳮRJd,9
where Jis the sensitivity or the Fréchet derivative matrix. In equa-
tion 9,Cmis a vector of dimension M. Takingmm0in equation 9
provides the descent direction along which the perturbation model is
searched in equation 7.
Differentiation of the gradient expression 8, with respect to the
model parameters gives the following expression in matrix form for
the Hessian see Pratt et al. 1998for details:
2Cm0
m2RJ0
J0R
J0
t
mtd0
*...d0
*
.10
Inserting the expression of the gradient equation 9and the Hessian
equation 10into equation 7gives the following for the perturbation
model:
m⳱ⳮ
R
J0
J0
J0
t
mtd0
*...d0
*
1
RJ0
d0.
11
The method solving the normal equations, e.g., equation 11, general-
ly is referred to as the Newton method, which is locally quadratically
convergent.
For linear problems dG.m, the second term in the Hessian is
zero because the second-order derivative of the data with respect to
model parameters is zero. Most of the time, this second-order term is
neglected for nonlinear inverse problems. In the following, the re-
maining term in the Hessian, i.e., HaJ0
J0, is referred to as the ap-
proximate Hessian. The method which solves equation 11 when only
Hais estimated is referred to as the Gauss-Newton method.
Alternatively, the inverse of the Hessian in equation 11 can be re-
placed by a scalar
, the so-called step length, leading to the gradient
or steepest-descent method. The step length can be estimated by a
line-search method, for which a linear approximation of the forward
problem is used Gauthier et al., 1986;Tarantola, 1987. In the linear
approximation framework, the second-order Taylor-Lagrange de-
velopment of the misfit function gives
Cm
Cm0兲兲Cm
Cm兲兩Cm0兲典
1
2
2Ham兲具Cm0兲兩Cm0兲典,
12
where we assume a model perturbation of the form m
Cm0. In equation 12, we replace the second-order deriva-
tive of the misfit function by the approximate Hessian in the second
term on the right-hand side. Inserting the expression of the approxi-
mate Hessian Hainto the previous expression, zeroing the partial de-
rivative of the misfit function with respect to
, and using mm0
gives
Cm0兲兩Cm0兲典
Jtm0Cm0兲兩Jtm0Cm0兲兲典 .13
The term Jtm0Cm0is computed conventionally using a first-
order-accurate finite-difference approximation of the partial deriva-
tive of G,
Gm0
mCm01
Gm0Cm0兲兲Gm0兲兲,
14
with a small parameter . Estimation of
requires solving an extra
forward problem per shot for the perturbed model m0Cm0.
This line-search technique is extended to multiple-parameter classes
by Sambridge et al. 1991using a subspace approach. In this case,
one forward problem must be solved per parameter class, which can
be computationally expensive. Alternatively, the step length can be
estimated by parabolic interpolation through three points,
,Cm0
Cm0兲兲. The minimum of the parabola provides the desired
. In this case, two extra forward problems per shot must be solved
because we already have a third point corresponding to 0,Cm0兲兲
see Figure 1in Vigh et al. 2009for an illustration.
Pratt et al. 1998illustrate how quality and rate of convergence of
the inversion depend significantly on the Newton, Gauss-Newton, or
gradient method used. Importantly, they show how the gradient
method can fail to converge toward an acceptable model, however
many iterations, unlike the Newton and Gauss-Newton methods.
They interpret this failure as the result of the difficulty of estimating
a reliable step length. However, gradient methods can be significant-
ly improved by scaling i.e., dividingthe gradient by the diagonal
terms of Haor of the pseudo-Hessian Shin et al., 2001a.
Full-waveform inversion WCC131
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
Numerical algorithms: The conjugate-gradient method
Over the last decade, the most popular local optimization algo-
rithm for solving FWI problems was based on the conjugate-gradi-
ent method Mora, 1987;Tarantola, 1987;Crase et al., 1990. Here,
the model is updated at the iteration nin the direction of pn, which is
a linear combination of the gradient at iteration n,Cn, and the di-
rection pn1:
pnCn
npn1.15
The scalar
nis designed to guarantee that pnand pn1are con-
jugate. Among the different variants of the conjugate-gradient meth-
od to derive the expression of
n, the Polak-Ribière formula Polak
and Ribière, 1969is generally used for FWI:
nCnCn1Cn
Cn2.16
In FWI, the preconditioned gradient Wm
1Cnis used for pn, where
Wmis a weighting operator that is introduced in the next section
Mora, 1987. Only three vectors of dimension M, i.e., Cn,
Cn1, and pn1, are required to implement the conjugate-gradi-
ent method.
Numerical algorithms: Quasi-Newton algorithms
Finite approximations of the Hessian and its inverse can be com-
puted using quasi-Newton methods such as the BFGS algorithm
named after its discoverers Broyden, Fletcher, Goldfarb, and Sh-
anno; see Nocedal 1980for a review. The main idea is to update
the approximation of the Hessian or its inverse Hnat each iteration
of the inversion, taking into account the additional knowledge pro-
vided by Cnat iteration n. In these approaches, the approximation
of the Hessian or its inverse is explicitly formed.
For large-scale problems such as FWI in which the cost of storing
and working with the approximation of the Hessian matrix is prohib-
itive, a limited-memory variant of the quasi-Newton BFGS method
known as the L-BFGS algorithm allows computing in a recursive
manner HnCnwithout explicitly forming Hn. Only a few gradi-
ents of the previous nonlinear iterations typically, 3–20 iterations
need to be stored in L-BFGS, which represents a negligible storage
and computational cost compared to the conjugate-gradient algo-
rithm see Nocedal, 1980; p. 177–180. The L-BFGS algorithm re-
quires an initial guess H0, which can be provided by the inverse of
the diagonal Hessian Brossier et al., 2009a. For multiparameter
FWI, the L-BFGS algorithm provides a suitable scaling of the gradi-
ents computed for each parameter class and hence provides a com-
putationally efficient alternative to the subspace method of Sam-
bridge et al. 1991. A comparison between the conjugate-gradient
method and the L-BFGS method for a realistic onshore application
of multiparameter elastic FWI is shown in Brossier et al. 2009a.
Newton and Gauss-Newton algorithms
The more accurate, although more computationally intensive,
Gauss-Newton and Newton algorithms are described in Akcelik
2002,Askan et al. 2007,Askan and Bielak 2008, and Epano-
meritakis et al. 2008, with an application to a 2D synthetic model
of the San Fernando Valley using the SH-wave equation. At each
nonlinear FWI iteration, a matrix-free conjugate-gradient method is
used to solve the reduced Karush-Kuhn-Tucker KKToptimal sys-
tem, which turns out to be similar to the normal equation system
equation 11. Neither the full Hessian nor the sensitivity matrix is
formed explicitly; only the application of the Hessian to a vector
needs to be performed at each iteration of the conjugate-gradient al-
gorithm.
Application of the Hessian to a vector requires performing two
forward problems per shot for the incident and the adjoint wavefields
Akcelik, 2002. Because these two simulations are performed at
each iteration of the conjugate-gradient algorithm, an efficient pre-
conditioner must be used to mitigate the number of iterations of the
conjugate-gradient algorithm. Epanomeritakis et al. 2008use a
variant of the L-BFGS method for the preconditioner of the conju-
gate gradient, in which the curvature of the objective function is up-
dated at each iteration of the conjugate gradient using the Hessian-
vector products collected over the iterations.
Regularization and preconditioning of inversion
As widely stressed, FWI is an ill-posed problem, meaning that an
infinite number of models matches the data. Some regularizations
are conventionally applied to the inversion to make it better posed
Menke, 1984;Tarantola, 1987;Scales et al., 1990.The misfit func-
tion can be augmented as follows:
Cm1
2dWdd1
2mmpriorWmmmprior,
17
where WdSd
tSdand WmSm
tSm. Weighting operators are Wdand
Wm, the inverse of the data and model covariance operators in the
frame of the Bayesian formulation of FWI Tarantola, 1987. The
operator Sdcan be implemented as a diagonal weighting operator
that controls the respective weight of each element of the data-misfit
vector. For example, Operto et al. 2006use Sdas a power of the
source-receiver offset to strengthen the contribution of large-offset
data for crustal-scale imaging. In geophysical applications where the
smoothest model that fits the data is often sought, the aim of the
least-squares regularization term in the augmented misfit function
equation 17is to penalize the roughness of the model m, hence de-
fining the so-called Tikhonov regularization see Hansen 1998for
a review on regularization methods. The operator Smis generally a
roughness operator, such as the first- or second-difference matrices
Press et al., 1986, 1007.
For linear problems assuming the second term of the Hessian is
neglected, the minimization of the weighted misfit function gives
the perturbation model:
m⳱ⳮRJ0
WdJ0Wm1RJ0
Wdd0,18
where we use mprior m0. Of note, equation 18 is equivalent to
Tarantola 1987,p.70and Menke 1984,p.55:
m⳱ⳮWm
1RJ0Wm
1J0
Wd
11RJ0
d0.
19
Equation 19 can be more tractable from a computational viewpoint
when NM. Because Wmis a roughness operator, Wm
1is a
smoothing operator. It can be implemented, for example, with a mul-
tidimensional adaptive Gaussian smoother Ravaut et al., 2004;Op-
erto et al., 2006or with a low-pass filter in the wavenumber domain
Sirgue, 2003.
For the steepest-descent algorithm, the regularized solution for
the perturbation model is given by
WCC132 Virieux and Operto
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
m⳱ⳮ
Wm
1RJ0
Wdd0,20
where the scaling performed by the diagonal terms of the approxi-
mate Hessian can be embedded in the operator Wm
1in addition to the
smoothing operator.
A more complete and rigorous mathematical derivation of these
equations is presented in Tarantola 1987.
Alternative regularizations based on minimizing the total varia-
tion of the model have been developed mainly by the image-process-
ing and electromagnetic communities. The aim of the total variation
or edge-preserving regularization is to preserve both the blocky and
the smooth characteristics of the model Vogel and Oman, 1996;Vo-
gel, 2002. Total variation TVregularization is conventionally im-
plemented by minimizing the L1-norm of the model-misfit function
RTV mWmm1/2. Alternatively, van den Berg and Abubakar
2001implement TV regularization as a multiplicative constraint in
the original misfit function. In this framework, the original misfit
function can be seen as the weighting factor of the regularization
term, which is automatically updated by the optimization process
without the need for heuristic tuning. TV regularization is applied to
FWI in Askan and Bielak 2008. The weighted L2-norm regulariza-
tion applied to frequency-domain FWI is shown in Hu et al. 2009
and Abubakar et al. 2009.
The gradient and Hessian in FWI: Interpretation and
computation
A clear interpretation of the gradient and Hessian is given by Pratt
et al. 1998using the compact matrix formalism of frequency-do-
main FWI. Areview is given here. Let us consider the forward-prob-
lem equation given by equation 2for one source and one frequency.
In the following, we assume that the model is discretized in a finite-
difference sense using a uniform grid of nodes.
Differentiation of equation 2with respect to the model parameter
mlgives the expression of the partial derivative wavefield
u/
m
by solving the following system:
B
u
m
f,21
where
f⳱ⳮ
B
m
u.22
An analogy between the forward-problem equation 2and equa-
tion 21 shows that the partial-derivative wavefield can be computed
by solving one forward problem, the source of which is given by f.
This so-called virtual secondary source is formed by the product of
B/
mand the incident wavefield u. The matrix
B/
mis built by
differentiating each coefficient of the forward-problem operator B
with respect to m. Because the discretized differential operators in B
generally have local support, the matrix
B/
mis extremely
sparse.
The spatial support of the virtual secondary source is centered on
the position of m, whereas the temporal support of fis centered
around the arrival time of the incident wavefield at the position of m.
Therefore, the partial-derivative wavefield with respect to the model
parameter mcan be interpreted as the wavefield emitted by the seis-
mic source sand scattered by a point diffractor located at m. The ra-
diation pattern of the virtual secondary source is controlled by the
operator
B/
m. Analysis of this radiation pattern for different pa-
rameter classes allows us to assess to what extent parameters of dif-
ferent natures are uncoupled in the tomographic reconstruction as a
function of the diffraction angle and to what extent they can be reli-
ably reconstructed during FWI. Radiation patterns for the isotropic
acoustic, elastic, and viscoelastic wave equations are shown in Wu
and Aki 1985,Tarantola 1986,Ribodetti and Virieux 1996, and
Forgues and Lambaré 1997.
Because the gradient is formed by the zero-lag correlation be-
tween the partial-derivative wavefield and the data residual, these
have the same meaning: They represent perturbation wavefields
scattered by the missing heterogeneities in the starting model m0
Tarantola, 1984;Pratt et al., 1998.This interpretation draws some
connections between FWI and diffraction tomography; the perturba-
tion model can be represented by a series of closely spaced diffrac-
tors. By virtue of Huygens’ principle, the image of the model pertur-
bations is built by the superposition of the elementary image of each
diffractor, and the seismic wavefield perturbation is built by super-
position of the wavefields scattered by each point diffractor Mc-
Mechan and Fuis, 1987.
The approximate Hessian is formed by the zero-lag correlation
between the partial-derivative wavefields, e.g., equation 10. The di-
agonal terms of the approximate Hessian contain the zero-lag auto-
correlation and therefore represent the square of the amplitude of the
partial-derivative wavefield. Scaling the gradient by these diagonal
terms removes from the gradient the geometric amplitude of the par-
tial-derivative wavefields and the residuals. In the framework of sur-
face seismic experiments, the effects of the scaling performed by the
diagonal Hessian provide a good balance between shallow and deep
perturbations. A diagonal Hessian is shown in Ravaut et al. 2004,
their Figure 12. The off-diagonal terms of the Hessian are computed
by correlation between partial-derivative wavefields associated with
different model parameters. For 1D media, the approximate Hessian
is a band-diagonal matrix, and the numerical bandwidth decreases as
the frequency increases. The off-diagonal elements of the approxi-
mate Hessian account for the limited-bandwidth effects that result
from the experimental setup. Applying its inverse to the gradient can
be interpreted as a deconvolution of the gradient from these limited-
bandwidth effects.
An illustration of the scaling and deconvolution effects performed
by the diagonal Hessian on one hand and the approximate Hessian
on the other hand is provided in Figure 1. Asingle inclusion in a ho-
mogeneous background model Figure 1ais reconstructed by one
iteration of FWI using a gradient method preconditioned by the diag-
onal terms of the approximate Hessian Figure 1band by a Gauss-
Newton method Figure 1c. The image of the inclusion is sharper
when the Gauss-Newton algorithm is used. The corresponding ap-
proximate Hessian and its diagonal elements are shown in Figure 2.
An interpretation of the second term of the Hessian equation 10is
given in Pratt et al. 1998. This term accounts for multiscattering
events such as multiples in the reconstruction procedure. Through it-
erations, we might correct effects caused by this missing term as
long as convergence is achieved.
Although equation 21 gives some clear insight into the physical
sense of the gradient of the misfit function, it is impractical from a
computer-implementation point of view; with the computer explicit-
ly forming the sensitivity matrix with equation 21, it would require
performing as many forward problems as the number of model pa-
rameters m1,Mfor each source of the survey. To mitigate this
computational burden, the spatial reciprocity of Green’s functions
can be exploited as shown below.
Full-waveform inversion WCC133
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
Inserting the expression of the partial derivative of the wavefield
equation 21in the expression of the gradient of equation 9gives the
following expression of the gradient:
CR
ut
B
m
t
B1tPd*
,23
where Pdenotes an operator that augments the residual data vector
with zeroes in the full computational domain so that the dimension
of the augmented vector matches the dimension of the matrix B1t
Pratt et al., 1998. The column of B1corresponds to the Green’s
functions for unit impulse sources located at each node of the model.
By virtue of the spatial reciprocity of the Green’s functions, B1is
symmetric. Therefore, B1tcan be substituted by B1in equation 23,
which gives
CR
ut
B
m
t
B1Pd*
R
ut
B
m
t
rb
.
24
The wavefield rbcorresponds to the back-propagated residual wave-
field. All of the residuals associated with one seismic source are as-
sembled to form one residual source. The back propagation in time is
indicated by the conjugate operator in the frequency domain. The
number of forward seismic problems for computing the gradient is
reduced to two: one to compute the incident wavefield uand one to
back propagate the corresponding residuals. The underlying imag-
ing principle is reverse-time migration, which relies on the corre-
spondence of the arrival times of the incident wavefield and the
0123
0
1
2
3
Distance (km)
Depth (km)
0123
0
1
2
3
Depth (km)
0123
0
1
2
3
Depth (km)
Distance (km)
Distance (km)
4.0 4.1 4.2 4.3
0
1
2
3
4.0 4.1 4.2 4.3
0
1
2
3
VP(km/s)
VP(km/s)
a)
b)
c)
d)
e)
Figure 1. Reconstruction of an inclusion by frequency-domain FWI.
aTrue model and FWI models built bby a preconditioned gradi-
ent method and cby a Gauss-Newton method. Four frequencies 4,
5, 7, and 10 Hzwere inverted. One iteration per frequency was com-
puted. Fourteen shots were deployed along the top and left edges of
the model. Shots along the top edge were recorded by 14 receivers
along the bottom edge; shots along the left edges were recorded by
14 receivers along the right edge. The P-wave velocities in the back-
ground and in the inclusion are 4.0 and 4.2 km/s, respectively. Verti-
cal velocity profiles are extracted from the true model gray lineand
the FWI models black linefor dthe gradient and ethe Gauss-
Newton inversions.
0
200
400
600
800
Row num
b
er
Depth (km)
Column number
0 200 400 600 800
0
5
10
15
20
25
30
Distance (km)
0 5 10 15 20 25 30
a)
b)
Figure 2. Hessian operator. aApproximate Hessian corresponding
to the 31 31 model of Figure 1for a frequency of 4 Hz. Aclose-up
of the area delineated by the yellow square highlights the band-diag-
onal structure of the Hessian. bCorresponding diagonal terms of
the Hessian plotted in the distance-depth domain. The high-ampli-
tude coefficients indicate source and receiver positions. Scaling the
gradient by this map removes the geometric amplitude effects from
the wavefields.
WCC134 Virieux and Operto
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
back-propagated wavefield at the position of heterogeneity Claer-
bout, 1971;Lailly, 1983;Tarantola, 1984.
The approach that consists of computing the gradient of the misfit
function without explicitly building the sensitivity matrix is often re-
ferred to as the adjoint-wavefield approach by the geophysical com-
munity. The underlying mathematical theory is the adjoint-state
method of the optimization theory Lions, 1972;Chavent, 1974. In-
teresting links exist between optimization techniques used in FWI
and assimilation methods, widely used in fluid mechanics Tala-
grand and Courtier, 1987. A detailed review of the adjoint-state
method with illustrations from several seismic problems is given in
Tromp et al. 2005,Askan 2006,Plessix 2006, and Epanomer-
itakis et al. 2008. The expression of the gradient of the frequency-
domain FWI misfit function equation 24is derived from the ad-
joint-state method and the method of the Lagrange multiplier in Ap-
pendix A.
For multiple sources and multiple frequencies, the gradient is
formed by the summation over these sources and frequencies:
C
i1
N
s1
Ns
R
Bi
1sst
Bi
m
t
Bi
1Pdi,s
*兲兴
.
25
We also need to note that matrices Bi
1i1,N
do not depend on
shots; therefore, any speedup toward resolving systems that involve
these matrices with multiple sources should be considered Marfurt,
1984;Jo et al., 1996;Stekl and Pratt, 1998.
By comparing the expressions of the gradient in equations 9and
24, we can conclude that one element of the sensitivity matrix is giv-
en by
Jks,r,us
t
Bt
m
B1
r,26
where ks,rdenotes a source-receiver couple of the acquisition ge-
ometry, with sand rdenoting a shot and a receiver position, respec-
tively. An impulse source
ris located at receiver position r.Ifthe
sensitivity matrix must be built, one forward problem for the inci-
dent wavefield and one forward problem per receiver position must
be computed. Therefore, the number of simulations to build the sen-
sitivity matrix can be higher than that required by gradient estima-
tion if the number of nonredundant receiver positions significantly
exceeds the number of nonredundant shots, or vice versa. Comput-
ing each term of the sensitivity matrix is also required to compute the
diagonal terms of the approximate Hessian HaShin et al., 2001b.
To mitigate the resulting computational burden for coarse OBS sur-
veys, Operto et al. 2006suggest computing the diagonal terms of
Hafor a decimated shot acquisition. Alternatively, Shin et al.
2001apropose using an approximation of the diagonal Hessian,
which can be computed at the same cost as the gradient.
Although the matrix-free adjoint approach is widely used in ex-
ploration seismology, the earthquake-seismology community tends
to favor the scattering-integral method, which is based on the explic-
it building of the sensitivity matrix Chen et al., 2007. The linear
system relating the model perturbation to the data perturbation is
formed and solved with a conjugate-gradient algorithm such as
LSQR Paige and Saunders, 1982a. A comparative complexity
analysis of the adjoint approach and the scattering-integral approach
is presented in Chen et al. 2007, who conclude that the scattering-
integral approach outperforms the adjoint approach for a regional to-
mographic problem. Indeed, the superiority of one approach over the
other is highly dependent on the acquisition geometry the relative
number of sources and receiversand the number of model parame-
ters.
The formalism in equation 25 has been kept as general as possible
and can relate to the acoustic or the elastic wave equation. In the
acoustic case, the wavefield is the pressure scalar wavefield; in the
elastic case, the wavefield ideally is formed by the components of the
particle velocity and the pressure if the sensors have four compo-
nents. Equation 25 can be translated in the time domain using Parse-
val’s relation. The expression of the gradient in equation 25 can be
developed equivalently using a functional analysis Tarantola,
1984. The partial derivatives of the wavefield with respect to the
model parameters are provided by the kernel of the Born integral that
relates the model perturbations to the wavefield perturbations. Mul-
tiplying the transpose of the resulting operator by the conjugate of
the data residuals provides the expression of the gradient. The two
formalisms matrix and functionalgive the same expression, pro-
vided the discretization of the partial differential operators are per-
formed consistently in the two approaches. The derivation in the fre-
quency domain of the gradient of the misfit function using the two
formalisms is explicitly illustrated by Gelis et al. 2007.
Source estimation
Source excitation is generally unknown and must be considered
as an unknown of the problem Pratt, 1999. The source wavelet can
be estimated by solving a linear inverse problem because the rela-
tionship between the seismic wavefield and the source is linear
equation 2. The solution for the source is given by the expression
sgcalm0兲兩dobs
gcalm0兲兩gcalm0兲典,27
where gcalm0denotes the Green’s functions computed in the start-
ing model m0. The source function can be estimated directly in the
FWI algorithm once the incident wavefields have been modeled.
The source and the medium are updated alternatively over iterations
of the FWI. Note that it is possible to take advantage of source esti-
mation to design alternative misfit functions based on the differential
semblance optimization Pratt and Symes, 2002or to define more
heuristic criteria to stop the iteration of the inversion Jaiswal et al.,
2009.
Alternatively, new misfit functions have been designed so the in-
version becomes independent of the source function Lee and Kim,
2003;Zhou and Greenhalgh, 2003. The governing idea of the meth-
od is to normalize each seismogram of a shot gather by the sum of all
of the seismograms. This removes the dependency of the normalized
data with respect to the source and modifies the misfit function. The
drawback is that this approach requires an explicit estimate of the
sensitivity matrix; the normalized residuals cannot be back propa-
gated because they do not satisfy the wave equation.
SOME KEY FEATURES OF FWI
Resolution power of FWI and relationship to the
experimental setup
The interpretation of the partial-derivative wavefield as the wave-
field scattered by the missing heterogeneities provides some connec-
tions between FWI and generalized diffraction tomography Dev-
aney and Zhang, 1991;Gelius et al., 1991. Diffraction tomography
Full-waveform inversion WCC135
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
recasts the imaging as an inverse Fourier transform Devaney, 1982;
Wu and Toksoz, 1987;Sirgue and Pratt, 2004;Lecomte et al., 2005.
Let us consider a homogeneous background model of velocity c0,an
incident monochromatic plane wave propagating in the direction s
ˆ
and a scattered monochromatic plane wave in the far-field approxi-
mation propagating in the direction r
ˆFigure 3. If amplitude effects
are not taken into account, the incident and scattered Green’s func-
tions can be written compactly as
G0x,sexpik0s
ˆ.x,
G0x,rexpik0r
ˆ.x,28
with the relation k0
/c0. Inserting the expression of the incident
and scattered plane waves into the gradient of the misfit function of
equation 24 gives the expression Sirgue and Pratt, 2004
Cm⳱ⳮ
2
s
r
Rexpik0s
ˆr
ˆ.xd.
29
Equation 29 has the form of a truncated Fourier series where the
integration variable is the scattering wavenumber vector given by k
k0s
ˆr
ˆ. The coefficients of the series are the data residuals. The
summation is performed over sources, receivers, and frequencies
that control the truncation and sampling of the Fourier series.
We can express the scattering wavenumber vector k0s
ˆr
ˆas a
function of frequency, diffraction angle, or aperture to highlight the
relationship between the experimental setup and the spatial resolu-
tion of the reconstruction Figure 4:
k2f
c0
cos
2
n,30
where nis a unit vector in the direction of the slowness vector s
ˆ
r
ˆ. Equation 30 was also derived in the framework of the ray
Born migration/inversion, recast as the inverse of a generalized Ra-
don transform Miller et al., 1987or as a least-squares inverse prob-
lem Lambaré et al., 2003.
Several key conclusions can be derived from equation 30. First,
one frequency and one aperture in the data space map one wavenum-
ber in the model space. Therefore, frequency and aperture have re-
dundant control of the wavenumber coverage. This redundancy in-
creases with aperture bandwidth. Pratt and Worthington 1990,Sir-
gue and Pratt 2004,and Brenders and Pratt 2007apropose deci-
mating this wavenumber-coverage redundancy in frequency-do-
main FWI by limiting the inversion to a few discrete frequencies.
This data reduction leads to computationally efficient frequency-do-
main FWI and allows managing a compact volume of data, two clear
advantages with respect to time-domain FWI. The guideline for se-
lecting the frequencies to be used in the FWI is that the maximum
wavenumber imaged by a frequency matches the minimum vertical
wavenumber imaged by the next frequency Sirgue and Pratt, 2004,
their Figure 3. According to this guideline, the frequency interval
increases with the frequency.
Second, the low frequencies of the data and the wide apertures
help resolve the intermediate and large wavelengths of the medium.
At the other end of the spectrum, the maximum wavenumber, con-
strained by
0 and the highest frequency, leads to a maximum res-
olution of half a wavelength if normal-incidence reflections are re-
corded. Third, for surface acquisitions, long offsets are helpful for
sampling the small horizontal wavenumbers of dipping structures
such as flanks of salt domes.
A frequency-domain sensitivity kernel for point sources, referred
to as the wavepath by Woodward 1992, is shown in Figure 5. The
interference picture shows zones of equiphase over which the resid-
uals are back projected during FWI. The central zone of elliptical
shape is the first Fresnel zone of width
osr, where osr is the source-
receiver offset. Residuals that match the first arrival with an error
lower than half a period will be back projected constructively over
the first Fresnel zone, updating the large wavelengths of the struc-
ture. The outer fringes are isochrones on which residuals associated
with later-arriving reflection phases will be back projected, provid-
ing an update of the shorter wavelengths of the medium, just like
PSDM Lecomte, 2008. The width of the isochrones, which gives
some insight into the vertical resolution in Figure 4, is given by the
modulus of the wavenumber of equation 30.
To illustrate the relationship between FWI resolution and the ex-
perimental setup, we show the FWI reconstruction of an inclusion in
a homogeneous background for three acquisition geometries Figure
6. In the crosshole experiment Figure 6a, FWI has reconstructed a
low-pass-filtered smoothedversion of the vertical section of the in-
clusion and a band-pass-filtered version of the horizontal section of
the inclusion. This anisotropy of the imaging results from the trans-
mission-like reconstruction of the vertical wavenumbers and the re-
flection-like reconstruction of the horizontal wavenumbers of the in-
clusion. In the case of the double crosshole experiment Figure 6b,
the vertical and horizontal wavenumber spectra of the inclusion have
4
0
1
2
3
01234
Distance (km)
Dept
h
(
k
m)
0
1
2
3
01234
Distance (km)
0
1
2
3
0123
4
Distance (km)
44
s
^
r
^
k
s
^r
^
a) b) c
)
Figure 3. Resolution analysis of FWI. aIncident monochromatic
plane wave real part.bScattered monochromatic plane wave
real part.cGradient of FWI describing one wavenumber compo-
nent real partbuilt from the plane waves shown in aand b.
S
R
ps
pr
x
q
θ
k=fq
λ
=c/f
q=ps+p
r
ps=pr=1/c
Figure 4. Illustration of the main parameters in diffraction tomogra-
phy and their relationships. Key:
, wavelength;
, diffraction or ap-
erture angle; c, P-wave velocity; f, frequency; pS,pR,q, slowness
vectors; k, wavenumber vector; x, diffractor point; Sand R, source
and receiver positions.
WCC136 Virieux and Operto
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
been partly filled because of the combined use of transmission and
reflection wavepaths. Of note, the vertical section shows a lack of
low wavenumbers, whereas the horizontal section exhibits a deficit
of low wavenumbers because the maximum horizontal source-re-
ceiver offset is two times higher than the vertical one Figure 6b.
Therefore, the aperture illuminations of the horizontal and vertical
wavenumbers differ. For the surface acquisition Figure 6c, the ver-
tical section exhibits a strong deficit of low wavenumbers because of
the lack of large-aperture illumination. Of note, the pick-to-pick am-
plitude of the perturbation is fully recovered in Figure 6c. The hori-
zontal section of the inclusion is poorly recovered because of the
poor illumination of the horizontal wavenumbers from the surface.
The ability of the wide apertures to resolve the large wavelengths
of the medium has prompted some studies to consider long-offset ac-
quisitions as a promising approach to design well-posed FWI prob-
lems Pratt et al., 1996;Ravaut et al., 2004. For example, equation
30 can suggest that the long wavelengths of the medium can be re-
solved whatever the source bandwidth, provided that wide-aperture
data are recorded by the acquisition geometry. However, all of the
conclusions derived so far rely on the Born approximation. The Born
approximation requires that the starting model allows matching the
observed traveltimes with an error less than half the period Bey-
doun and Tarantola, 1988. If not, the so-called cycle-skipping arti-
facts will lead to convergence toward a local minimum Figure 7.
Pratt et al. 2008translates this condition in terms of relative time
error t/TLas a function of the number of propagated wavelengths
N
, expressed as
t
TL
1
N
,31
where TLdenotes the duration of the simulation. Condition 31 shows
that the traveltime error must be less than 1% for an offset involving
50 propagated wavelengths, a condition unlikely to be satisfied if
FWI is applied without data preconditioning. Therefore, some stud-
ies consider that recording low frequencies 1Hzis the best strat-
egy to design well-posed FWI Sirgue, 2006. Unfortunately, such
low frequencies cannot be recorded during controlled-source exper-
iments. As an alternative to low frequencies, multiscale layer-strip-
ping approaches where longer offsets, shorter apertures, and longer
recording times are progressively introduced in the inversion, have
been designed to mitigate the nonlinearity of the inversion.
Multiscale FWI: Time domain versus frequency domain
FWI can be implemented in the time domain or in the frequency
domain. FWI was originally developed in the time domain Taran-
X
X
X
θ
010 20 30 40 50 60 70 80 90 100
0
5
10
15
20
Distance (
k
m)
010 20 30 40 50 60 70 80 90 100
0
5
10
15
20
Depth (km)Depth (km)
sr
a)
b)
Figure 5. Wavepath. aMonochromatic Green’s function for a point
source. bWavepath for a receiver located at a horizontal distance
of 70 km from the source. The frequency is 5 Hz and the velocity in
the homogeneous background is 6 km/s. The dashed red lines delin-
eate the first Fresnel zone and an isochrone surface. The yellow line
is a vertical section across the wavepath. The blue lines represent dif-
fraction paths within the first Fresnel zone and from the isochrone.
Depth (km)
Distance (km)
012345678
0
1
2
3
4
V(km/s)
P
4.1
4.0
Distance (km)
012345678
0
1
2
3
4
4.2
4.1
4.0
VP(km/s)
Distance (km)
012345678
0
1
2
3
4
Depth (km)
4.0
3.9
VP(km/s)
Distance (km)
012345678
0
1
2
3
4
Depth (km)
a)
b)
c)
Figure 6. Imaging an inclusion by FWI. aCrosshole experiment.
Source and receiver lines are in red and blue, respectively. The con-
tour of the inclusion with a diameter of 400 m is delineated by the
blue circle. The true velocity in the inclusion is 4.2 km/s, whereas the
velocity in the background is 4 km/s. Six frequencies 4, 7, 9, 11, 12,
and 15 Hzwere inverted successively, and 20 iterations per frequen-
cy were computed. The black and gray curves along the right and
bottom sides of the model are velocity profiles across the center of
the inclusion extracted from the exact model and the reconstructed
model, respectively. bSame as afor a vertical and horizontal
crosshole experiment the shots along the red dashed line are record-
ed by only the receivers along the vertical blue dashed line.c
Same as afor a surface experiment.
Full-waveform inversion WCC137
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
tola, 1984;Gauthier et al., 1986;Mora, 1987;Crase et al., 1990
whereas the frequency-domain approach was proposed mainly in
the 1990s by Pratt and collaborators Pratt, 1990;Pratt and Wor-
thington, 1990;Pratt and Goulty, 1991, first with application to
crosshole data and later with application to wide-aperture surface
seismic data Pratt et al., 1996.
The nonlinearity of FWI has prompted many studies to develop
some hierarchical multiscale strategies to mitigate this nonlinearity.
Apart from computational efficiency, the flexibility offered by the
time domain or the frequency domain to implement efficient multi-
scale strategies is one of the main criteria that favors one domain
rather than the other. The multiscale strategy successively processes
data subsets of increasing resolution power to incorporate smaller
wavenumbers in the tomographic models. In the time domain,
Bunks et al. 1995propose successive inversion of subdata sets of
increasing high-frequency content because low frequencies are less
sensitive to cycle-skipping artifacts. The frequency domain provides
a more natural framework for this multiscale approach by perform-
ing successive inversions of increasing frequencies. In the frequency
domain, single or multiple frequencies i.e., frequency groupcan be
inverted at a time.
Although a few discrete frequencies theoretically are sufficient to
fill the wavenumber spectrum for wide-aperture acquisitions, simul-
taneous inversion of multiple frequencies improves the signal-to-
noise ratio and the robustness of FWI when complex wave phenom-
ena are observed i.e., guide waves, surface waves, dispersive
waves. Therefore, a trade-off between computational efficiency and
quality of imaging must be found. When simultaneous multifre-
quency inversion is performed, the bandwidth of the frequency
group ideally must be as large as possible to mitigate the nonlinearity
of FWI in terms of the nonunicity of the solution, whereas the maxi-
mum frequency of the group must be sufficiently low to prevent cy-
cle-skipping artifacts. An illustration of this tuning of FWI is given
in Brossier et al. 2009ain the framework of elastic seismic imaging
of complex onshore models from the joint inversion of surface
waves and body waves.
The regularization effects introduced by hierarchical inversion of
increasing frequencies might not be sufficient to provide reliable
FWI results for realistic frequencies and realistic starting models in
the case of complex structures. This has prompted some studies to
design additional regularization levels in FWI. One of these is to se-
lect a subset of arrivals as a function of time. An aim of this time win-
dowing is to remove arrivals that are not predicted by the physics of
the wave equation implemented in FWI for example, PS-converted
waves in the frame of acoustic FWI. A second aim is to perform a
heuristic selection of aperture angles in the data. Considering a nar-
row time window centered on the first arrival leads to so-called ear-
ly-arrival waveform tomography Sheng et al., 2006. Time win-
dowing the data around the first arrivals is equivalent to selecting the
large-aperture components of the data to image the large and inter-
mediate wavelengths of the medium. Alternatively, time windowing
can be applied to isolate later-arriving reflections or PS-converted
phases to focus on imaging a specific reflector or a specific parame-
ter class, such as the S-wave velocity Shipp and Singh, 2002;Sears
et al., 2008;Brossier et al., 2009a.
The frequency domain is the most appropriate to select one or a
few frequencies for FWI, but the time domain is the most appropriate
to select one type of arrival for FWI. Indeed, time windowing cannot
be applied in frequency-domain modeling, in which only one or few
frequencies are modeled at a time. Alast resort is the use of complex-
valued frequencies, which is equivalent to the exponential damping
of a signal ptin time from an arbitrary traveltime t0Sirgue, 2003;
Brenders and Pratt, 2007b:
P
i
et0/
ptett0/
ei
tdt,32
where P
denotes the Fourier transform of ptand
is the damp-
ing factor.
A last regularization level can be implemented by layer stripping,
in which the imaging proceeds hierarchically from the shallow part
to the deep part. Layer stripping in FWI can be applied by combined
offset and temporal windowing Shipp and Singh, 2002;Wang and
Rao, 2009.
These three levels of regularization frequency dependent, time
dependent, and offset dependent can be combined in one integrat-
ed multiloop FWI workflow.An example is provided in Shin and Ha
2009and Brossier et al. 2009a, in which the frequency- and time-
dependent regularizations are implemented into two nested loops
over frequency groups and time-damping factors. In this approach,
the frequencies increase in the outer loop and the damping factors
decrease in the inner loop. In Figure 8, the VPand VSmodels of the
overthrust model are inferred from the successive inversion of two
groups of five frequencies Brossier et al., 2009a. The frequencies
of the first group range from 1.7 to 3.5 Hz, whereas those of the sec-
ond group range from 3.5 to 7.2 Hz. Five damping factors of
be-
tween 0.67 and 30.0 s were applied hierarchically for data precondi-
tioning during the inversion of each frequency group. Without these
two regularization levels associated with frequency and aperture se-
lections, FWI fails to converge toward acceptable models.
In summary, the implementation of FWI in the frequency domain
allows the easy implementation of multiscale FWI based on the hier-
archical inversion of groups of frequencies of arbitrary bandwidth
and sampling intervals. Time-domain modeling provides the most
flexible framework to apply time windowing of arbitrary geometry.
n n +1
n1
n
Time
(
s
)
n+1
nn 1
T/2 T/2
n+1
n1
Figure 7. Schematic of cycle-skipping artifacts in FWI. The solid
black line represents a monochromatic seismogram of period Tas a
function of time. The upper dashed line represents the modeled
monochromatic seismograms with a time delay greater than T/2. In
this case, FWI will update the model such that the n1th cycle of
the modeled seismograms will match the nth cycle of the observed
seismogram, leading to an erroneous model. In the bottom example,
FWI will update the model such that the modeled and recorded nth
cycle are in-phase because the time delay is less than T/2.
WCC138 Virieux and Operto
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
This makes frequency-domain FWI based on time-domain model-
ing an attractive strategy to design robust FWI algorithms. This is es-
pecially true for 3D problems, for which time-domain modeling has
several advantages with respect to frequency-domain modeling Sir-
gue et al., 2008.
On the parallel implementation of FWI
FWI algorithms must be implemented in parallel to address large-
scale 3D problems. Depending on the numerical technique for solv-
ing the forward problem, different parallel strategies can be consid-
ered for FWI. If the forward problem is based on numerical methods
such as time-domain modeling or iterative solvers, which are not de-
manding in terms of memory, a coarse-grained parallelism that con-
sists of distributing sources over processors is generally used and the
forward problem is performed sequentially on each processor for
each source Plessix, 2009. If the number of processors significant-
ly exceeds the number of shots, which can be the case if source en-
coding techniques are used Krebs et al., 2009, a second level of
parallelism can be viewed by domain decomposition of the physical
computational domain. A comprehensive review of different algo-
rithms to efficiently compute the forward and the adjoint wavefields
in time-domain FWI is presented by Akcelik 2002.
In contract, if the forward problem is based on a method that em-
beds a memory-expensive preprocessing step, such as LU factoriza-
tion in the frequency-domain direct-solver approach, parallelism
must be based on a domain decomposition of the computational do-
main. Each processor computes a subdomain of the wavefields for
all sources. Examples of such algorithms are described in Ben Hadj
Alietal.2008a,Brossier et al. 2009a, and Sourbier et al. 2009a,
2009b. A contrast source inversion CSImethod is described by
Abubakar et al. 2009, which allows a decrease in the number of LU
factorization in frequency-domain FWI at the expense of the number
of iterations.
Variants of classic least-squares and
Born-approximation FWI
Although the most popular approach of FWI is based on minimiz-
ing the least-squares norm of the data misfit on the one hand and on
the Born approximation for estimating partial-derivative wavefields
on the other, several variants of FWI have been proposed over the
last decade. These variants relate to the definition of the minimiza-
tion criteria, the representation of the data amplitude, phase, loga-
rithm of the complex-valued data, envelopein the misfit, or the lin-
earization procedure of the inverse problem.
The choice of the minimization criterion
The least-squares norm approach assumes a Gaussian distribution
of the misfit Tarantola, 1987. Poor results can be obtained when
this assumption is violated, for example, when large-amplitude out-
liers affect the data. Therefore, careful quality control of the data
must be carried out before least-squares inversion. Crase et al.
1990investigate several norms such as the least-squares norm L2,
the least-absolute-values norm L1, the Cauchy criterion norm, and
the hyperbolic secant sechcriterion in FWI Figure 9. The
L1-norm specifically ignores the amplitude of the residuals during
back propagation of the residuals when gradient building, making
this criterion less sensitive to large errors in the data. The Cauchy and
sechcriteria can be considered a combination of the L1- and the
L2-norms with different transitions between the norms. Crase et al.
1990conclude that the most reliable FWI results have been ob-
tained with the Cauchy and the sechcriteria.
The L2and Cauchy criteria are also compared by Amundsen
1991in the framework of frequency-wavenumber-domain FWI
for stratified media described by velocity, density, and layer thick-
nesses Amundsen and Ursin, 1991. They consider random noise
and weather noise and conclude in both cases that the Cauchy criteri-
on leads to the more robust results.
The Huber norm also combines the L1- and the L2-norms; it is
combined with quasi-Newton L-BFGS by Guitton and Symes
2003and Bube and Nemeth 2007. The Huber norm is also used in
the framework of frequency-domain FWI by Ha et al. 2009and
shows an overall more robust behavior than the L2-norm.
The choice of the linearization method
The sensitivity matrix is generally computed with the Born ap-
proximation, which assumes a linear tangent relationship between
the model and wavefield perturbations Woodward, 1992. This lin-
ear relationship between the perturbations can be inferred from the
assumption that the wavefield computed in the updated model is the
wavefield computed in the starting model plus the perturbation
wavefield.
VP(km/s)
VS(km/s
)
04812
16 20
0
1
2
3
4
04812
16 20
0
1
2
3
4
048
12 16
0
2
4
6
8
10
12
14
Depth (km)
048
12 16
0
2
4
6
8
10
12
14
Depth (km)
048
12 16
0
2
4
6
8
10
12
14
Depth (km)
048
12 16
0
2
4
6
8
10
12
14
Depth (km)
Depth
(
km
)
Depth (km)
0
1
2
3
4
Depth (km)
0
1
2
3
4
Depth (km)
346
23
Offset (km) Offset (km)
Offset (km) Offset (km)
VS(km/s)
VP(km/s)
x= 7.5 km
x= 7.5 km
Distance (km)
a
)
b) c)
Figure 8. Multiscale strategy for elastic FWI with application to the
overthrust model. aFWI velocity models VPtopand VSbottom.
bComparison between the logs from the true model black, the
starting model dashed gray, and the final FWI model solid gray.
cSynthetic seismograms computed in the final FWI models for the
horizontal leftand vertical rightcomponents of particle velocity.
The bottom panels are the final residuals between seismograms
computed in the true and in the final FWI models image courtesy R.
Brossier.
Full-waveform inversion WCC139
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
The Rytov approach considers the generalized phase as the wave-
field Woodward, 1992.The Rytov approximation provides a linear
relationship between the complex-phase perturbation and the model
perturbation by assuming that the wavefield computed in the updat-
ed model is related to the wavefield computed in the starting model
through ux,
u0x,
exp
x,
兲兲, where
x,
denotes
the complex-phase perturbation. The sensitivity of the Rytov kernel
is zero on the Fermat raypath because the traveltime is stationary
along this path. A linear relationship between the model perturba-
tions and the logarithm of the amplitude ratio LnA
/A0
兲兴 is
also provided by the Rytov approximation by taking the real part of
the sensitivity kernel of the Rytov integral instead of the imaginary
part that provides the phase perturbation.
The Born approximation is valid in the case of weak and small
perturbations, but the Rytov approximation is supposed to be valid
for large-aperture angles and a small amount of scattering per wave-
length, i.e., smooth perturbations or smooth variation in the phase-
perturbation gradient Beydoun and Tarantola, 1988.Although sev-
eral analyses of the Rytov approximation have been carried out, it re-
mains unclear to what extent its domain of validity differs signifi-
cantly from that of the Born approximation. Acomparison between
the Born approximation and the Rytov approximation in the frame-
work of elastic frequency-domain FWI is presented in Gelis et al.
2007. The main advantage of the Rytov approximation might be to
provide a natural separation between phase and amplitude e.g.,
Woodward, 1992. This separation allows the implementation of
phase and amplitude inversions Bednar et al., 2007;Pyun et al.,
2007from a frequency-domain FWI code using a logarithmic norm
Shin and Min, 2006;Shin et al., 2007, the use of which leads to the
Rytov approximation in the framework of local optimization
problems.
Another approach in the time-frequency domain is developed by
Fichtner et al. 2008for continental- and global-scale FWI, in
which the misfit of the phase and the misfit of the envelopes are min-
imized in a least-squares sense. The expected benefit from this ap-
proach is to mitigate the nonlinearity of FWI by separating the phase
and amplitude in the inversion and by inverting the envelope instead
of the amplitudes, the former being more linearly related to the data.
Building starting models for FWI
The ultimate goal in seismic imaging is to be able to apply FWI re-
liably from scratch, i.e., without the need for sophisticated a priori
information. Unfortunately, because multidimensional FWI at
present can only be attacked through local optimization approaches,
building an accurate starting model for FWI remains one of the most
topical issues because very low frequencies 1Hzstill cannot be
recorded in the framework of controlled-source experiments.
A starting model for FWI can be built by reflection tomography
and migration-based velocity analysis such as those used in oil and
gas exploration. A review of the tomographic workflow is given in
Woodward et al. 2008. Other possible approaches for building ac-
curate starting models, which should tend toward a more automatic
procedure and might be more closely related to FWI, are first-arrival
traveltime tomography FATT, stereotomography, and Laplace-do-
main inversion.
FATT performs nonlinear inversions of first-arrival traveltimes to
produce smooth models of the subsurface e.g., Nolet, 1987;Hole,
1992;Zelt and Barton, 1998. Traveltime residuals are back project-
ed along the raypaths to compute the sensitivity matrix. The tomog-
raphic system, augmented with smoothing regularization, generally
is solved with a conjugate-gradient algorithm such as LSQR Paige
and Saunders, 1982b. Alternatively, the adjoint-state method can be
applied to FATT, which avoids the explicit building of the sensitivity
matrix, just as in FWI Taillandier et al., 2009. The spatial resolu-
tion of FATT is estimated to be the width of the first Fresnel zone
Williamson, 1991; Figure 5.
Examples of applications of FWI to real data using a starting mod-
el built by FATT are shown, for example, in Ravaut et al. 2004,Op-
erto et al. 2006,Jaiswal et al. 2008,2009, and Malinowsky and
Operto 2008for surface acquisitions; in Dessa and Pascal 2003in
the framework of ultrasonic experimental data; in Pratt and Goulty
1991for crosshole data; and in Gaoetal.2006bfor VSP data.
Several blind tests that correspond to surface acquisitions have
been tackled by joint FATT and FWI. Results at the oil-exploration
scale and at the lithospheric scale are presented in Brenders and Pratt
2007a,2007b,2007cand suggest that very low frequencies and
very large offsets are required to obtain reliable FWI results when
the starting model is built by FATT. For example, only the upper part
of the BP benchmark model was imaged successfully by Brenders
and Pratt 2007cusing a starting frequency as small as 0.5 Hz and a
maximum offset of 16 km. Another drawback of FATT is that the
method is not suitable when low-velocity zones exist because these
low-velocity zones create shadow zones.
Reliable picking of first-arrival times is also a difficult issue when
low-velocity zones exist. Fitting first-arrival traveltimes does not
guarantee that computed traveltimes of later-arriving phases, such as
0
1
2
3
4
5
6
7
8
4321 0 1 2 3 4
L2functional
L1functional
Huber norm functional
Hybrid norm functional
Cost function
1
2
3
4
0
1
2
3
4
4321 0 1 2 3
4
Residual source amplitude
L2residual source
L1residual source
Huber norm functional
Hybrid norm functional
D
ata
r
es
i
dua
l
a
)
b)
Figure 9. Different functionals for FWI. aValue of four functionals
as a function of real-valued data residual. The L1,L2, Huber, and
mixed L1L2functionals are plotted as indicated. bAmplitude of
the residual source used to compute the back-propagated adjoint
wavefield for the different functionals shown in a. Note that the
back-propagated source in the L1-norm is not sensitive to the residu-
al amplitude and therefore is less sensitive to large-amplitude residu-
als than the L2-norm image courtesy R. Brossier.
WCC140 Virieux and Operto
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
reflections, will match the true reflection traveltimes with an error
that does not exceed half a period, especially if anisotropy affects the
wavefield. We should stress that FATT can be recast as a phase inver-
sion of the first arrival using a frequency-domain waveform-inver-
sion algorithm within which complex-valued frequencies are imple-
mented Min and Shin, 2006;Ellefsen, 2009. Compared to FATT
based on the high-frequency approximation, this approach helps ac-
count for the finite-frequency effect of the data in the sensitivity ker-
nel of the tomography. Judicious selection of the real and imaginary
parts of the frequency allows extraction of the phase of the first arriv-
al. The principles and some applications of the method are presented
in Min and Shin 2006and Ellefsen 2009for near-surface applica-
tions. This is strongly related to finite-frequency tomography Mon-
telli et al., 2004.
Traveltime tomography methods that can manage refraction and
reflection traveltimes should provide more consistent starting mod-
els for FWI. Among these methods, stereotomography is probably
one of the most promising approaches because it exploits the slope
of locally coherent events and a reliable semiautomatic picking pro-
cedure has been developed Lambaré, 2008. Applications of ste-
reotomography to synthetic and real data sets are presented in Bil-
lette and Lambaré 1998,Alerini et al. 2002,Billette et al. 2003,
Lambaré and Alérini 2005, and Dummong et al. 2008.
To illustrate the sensitivity of FWI to the accuracy of different
starting models, Figure 10 shows FWI reconstructions of the syn-
thetic Valhall model when the starting model is built by FATTand re-
flection stereotomography Prieux et al., 2009. In the case of ste-
reotomography, the maximum offset is 9 km and only the reflection
traveltimes are used Lambaré and Alérini, 2005, whereas the maxi-
mum offset is 32 km for FATT Prieux et al., 2009. Stereotomogra-
phy successfully reconstructs the large wavelength within the gas
cloud down to a maximum depth of 2.5 km; FATT fails to reconstruct
the large wavelengths of the low-velocity zone associated with the
gas cloud. However, the FWI model inferred from the FATT starting
model shows an accurate reconstruction of the shallow part of the
model. These results suggest that joint inversion of refraction and re-
flection traveltimes by stereotomography can provide a promising
framework to build starting models for FWI.
A third approach to building a starting model for FWI can be pro-
vided by Laplace-domain and Laplace-Fourier-domain inversions
Shin and Cha, 2008,2009;Shin and Ha, 2008. The Laplace-do-
main inversion can be viewed as a frequency-domain waveform in-
version using complex-valued frequencies see equation 32, the
real part of which is zero and the imaginary part of which controls the
time damping of the seismic wavefield. In other words, the principle
is the inversion of the DC component of damped seismograms where
the Laplace variable scorresponds to 1 /
in equation 32. The DC of
the undamped data is zero, but the DC of the damped data is not and
is exploited in Laplace-domain waveform inversion. The informa-
tion contained in the data can be similar to the amplitude of the wave-
field Shin and Cha, 2009. Laplace-domain waveform inversion
provides a smooth reconstruction of the velocity model, which can
be used as a starting model for Laplace-Fourier and classical fre-
quency-domain waveform inversions.
The Laplace-Fourier domain is equivalent to performing an inver-
sion of seismograms damped in time. The results shown in Shin and
Cha 2009suggest that this method can be applied to frequencies
smaller than the minimum frequency of the source bandwidth. The
ability of the method to use frequencies smaller than the frequencies
effectively propagated by the seismic source is called a mirage resur-
rection of the low frequencies by Shin and Cha 2009. An applica-
tion to real data from the Gulf of Mexico is shown in Shin and Cha
2009. For the real-data application, frequencies between 0 and 2
Hz in combination with nine Laplace damping constants are used for
the Laplace-Fourier-domain inversion, the final model of which is
used as the starting model for standard frequency-domain FWI.
Joint application of Laplace-domain, Laplace-Fourier-domain
and Fourier-domain FWI to the BP benchmark model is illustrated in
Figure 11 Shin and Cha, 2009. The starting model is a simple ve-
locity-gradient model Figure 11b. A first velocity model of the
large wavelengths is obtained by Laplace-domain inversion Figure
11c, which is subsequently used as a starting model for inversion in
the Laplace-Fourier-domain inversion, the final model of which is
shown in Figure 11d. During this stage, the starting frequency used
in the inversion of the damped data is as low as 0.01 Hz. The final
model obtained after frequency-domain FWI is shown in Figure 11e.
All of the structures were successfully imaged, beginning with a
very crude starting model.
4
3
2
1
0
4567891011
4
3
2
1
0
456789101
1
Depth (km)
Distance (km) Distance (km)
4
3
2
1
0
4567891011
4
3
2
1
0
4567891011
Depth (km)
1.6 2.0 2.4 2.8 3.2
km/s
1.2 2.2 3.2
0
1
2
3
4
Depth (km)
VP(km/s)
1.2 2.2 3.2
0
1
2
3
4
Depth (km)
VP(km/s)
1.2 2.2 3.2
0
1
2
3
4
Depth (km)
VP(km/s)
a)
b)
c)
d)
e) f) g)
Figure 10. aClose-up of the synthetic Valhall velocity model cen-
tered on the gas layer. bFWI model built from a starting model ob-
tained by smoothing the true model with a Gaussian filter with hori-
zontal and vertical correlation lengths of 500 m. cFWI model from
a starting model built by FATT Prieux et al., 2009.dFWI model
from a starting model built by stereotomography Lambaré and
Alérini, 2005.eVelocity profiles at a distance of 7.5 km extracted
from the true model black line, from the starting model built by
smoothing the true model blue line, and from the FWI model of b
red line.fSame as efor the starting model built by FATT and c
the corresponding FWI model. gSame as efor the starting model
built by stereotomography and dthe corresponding FWI model.
The frequencies used in the inversion are between 4 and 15 Hz.
Full-waveform inversion WCC141
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
CASE STUDIES
Applications of FWI have been applied essentially to synthetic
examples for which high-resolution images have been constructed.
Because FWI is an attractive approach, the number of real-data ap-
plications has increased quite rapidly, from monoparameter recon-
structions of the VPparameter to multiparameter ones.
Monoparameter acoustic FWI
Most of the recent real-data case studies of FWI at different scales
and for different acquisition geometries have been performed in the
acoustic isotropic approximation, considering only VPas the model
parameter Dessa and Pascal, 2003;Ravaut et al., 2004;Chironi et
al., 2006;Gao et al., 2006a,2006b;Operto et al., 2006;Bleibinhaus
et al., 2007;Ernst et al., 2007;Jaiswal et al., 2008,2009;Shin and
Cha, 2009. Although the acoustic approximation can be questioned
in the framework of FWI because of unreliable amplitudes, one ad-
vantage of acoustic FWI is dealing with less computationally expen-
sive forward modeling than in the elastic case. Moreover, acoustic
FWI is better posed than elastic FWI because only the dominant pa-
rameter VPcan be involved in the inversion.
Specific waveform-inversion data processing generally is de-
signed to account for the amplitude errors introduced by acoustic
modeling Pratt, 1999;Ravaut et al., 2004;Brenders and Pratt,
2007b. The amplitude discrepancies in the P-wavefield result from
incorrectly modeling the amplitude-variation-with-offset AVOef-
fects and incorrectly modeling the directivity of the source and re-
ceiver Brenders and Pratt, 2007b. Acoustic-wave modeling gener-
ally is based on resolving the acoustic-wave equation in pressure;
therefore, the particle-velocity synthetic wavefields might not be
computed Hustedt et al., 2004. If the receivers are geophones, the
physical measurements collected in the field particle velocitiesare
not the same as those computed by the seismic modeling engine
pressure. Amatch between the vertical geophone data and the pres-
sure synthetics can, however, be performed by exploiting the reci-
procity of the Green’s functions if the sources are explosions Operto
et al., 2006.
In contrast, if the sources and receivers are directional, the pres-
sure wavefield cannot account for the directivity of the sources and
receivers, and heuristic amplitude corrections must be applied be-
fore inversion. Brenders and Pratt 2007bpropose optimizing an
empirical correction law for the decay of the rms amplitudes with
offset. Applying this correction law to the modeled data matches the
main AVO trend of the observed data before FWI. Using this data
preprocessing, Brenders and Pratt 2007bsuccessfully image the
onshore lithospheric model of the CCSS blind test Zelt et al., 2005
by acoustic FWI of synthetic elastic data. This strategy is also used
successfully by Jaiswal et al. 2008,2009in the framework of
acoustic FWI of real onshore data in the Naga thrust and fold belt in
India.
Successful application of acoustic FWI to synthetic elastic data
computed in the marine Valhall model from an OBC acquisition is
presented by Brossier et al. 2009b. An application of acoustic FWI
to real onshore long-offset data recorded in the southern Apennines
thrust belt is illustrated in Figure 12 Ravaut et al., 2004. The veloc-
ity model is validated locally by comparison with a VSP log.Appli-
cation of PSDM using the final FWI model as a starting model con-
tributes to the validation of the relevance of the velocity structure re-
constructed by FWI Operto et al., 2004. Some guidelines based on
numerical examples of the domain of validity of acoustic FWI ap-
plied to elastic data are also provided in Barnes and Charara 2008.
Multiparameter FWI
Because FWI accounts for the full wavefield, the seismic model-
ing embedded in the FWI algorithm theoretically should honor as far
as possible all of the physics of wave propagation. This is especially
required by FWI of wide-aperture data, in which significant AVO
and azimuthal anisotropic effects should be observed in the data. The
requirement of realistic seismic modeling has prompted some stud-
ies to extend monoparameter acoustic FWI to account for parameter
classes other than the P-wave velocity, such as density, attenuation,
Distance
km
ept
m
ept
m
ept
m
ept
m
ept
m
Velocit
km/s
Velocity (km/s)
Velocit
km/s
Velocit
km/s
Velocit
km/s
0 102030405060
0
4.5
24.0
43.5
63.0
2.5
8
2.0
10
1.5
0 102030405060
0
4.5
24.0
43.5
63.0
2.5
8
2.0
10
1.5
0 102030405060
0
4.5
24.0
43.5
63.0
2.5
8
2.0
10
1.5
0 102030405060
0
4.5
24.0
43.5
63.0
2.5
8
2.0
10
1.5
0 102030405060
0
4.5
24.0
43.5
63.0
2.5
8
2.0
10
1.5
)
)
)
)
)
Figure 11. Laplace-Fourier-domain waveform inversion. aBP
benchmark model. bStarting model. cVelocity model after
Laplace-domain inversion. dVelocity model after Laplace-Fouri-
er-domain inversion. eVelocity model after frequency-domain
FWI image courtesy C. Shin and Y. H. Cha.
WCC142 Virieux and Operto
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
shear-wave velocity or related parameters, and anisotropy. The fact
that additional parameter classes are taken into account in FWI in-
creases in the ill-posedness of the inverse problem because more de-
grees of freedom are considered in the parameterization and because
the sensitivity of the inversion can change significantly from one pa-
rameter class to the next.
Different parameter classes can be more or less coupled as a func-
tion of the aperture angle. This coupling can be assessed by plotting
the radiation pattern of each parameter class using some asymptotic
analyses. Diffraction patterns of the different combination of param-
eters for the acoustic, elastic, and viscoelastic wave equation are
shown in Wu and Aki 1985,Tarantola 1986,Ribodetti and
Virieux 1996, and Forgues and Lambaré 1997.An alternative is
to plot the sensitivity kernel, i.e., that obtained by summing all of the
rows of the sensitivity matrix for the full acquisition and for different
combinations of parameters and to qualitatively assess which com-
bination provides the best image Luo et al., 2009. Hierarchical
strategies that successively operate on different parameter classes
should be designed to mitigate the ill-posedness of FWI Tarantola,
1986;Kamei and Pratt, 2008;Sears et al., 2008;Brossier et al.,
2009b.
Density
Density is difficult to reconstruct Forgues and Lambaré, 1997.
As an illustration, acoustic radiation patterns are shown in Figure 13
for different combinations of parameters IP,
,IP,VP, and VP,
,
where IPdenotes P-wave impedance. The radiation pattern of VPis
isotropic because the operator
B/
mlreduces to a scalar for VPand
therefore represents an explosion. On the other hand, the density has
the same radiation pattern as VPat short apertures but does not scatter
energy at wide apertures because the secondary source fcorre-
sponds to a vertical force for the density. Because VPand
have the
same radiation pattern at short apertures, these two parameters are
difficult to reconstruct from short-offset data. For such data, the
P-wave impedance can be considered a reliable parameter for FWI.
If wide-aperture data are available, VPand IPmight provide the most
judicious parameterization because they scatter energy for different
aperture bands wide and short apertures, respectively; Figure 13b.
A successful reconstruction of the density parameter in the case of
the Marmousi case study is presented by Choi et al. 2008. Howev-
er, the use of an unrealistically low frequency 0.125 Hzbrings into
question the practical implication of these results.
Attenuation
The attenuation reconstruction can be implemented in frequency-
domain seismic-wave modeling using complex velocities Toksöz
and Johnston, 1981. The most commonly applied attenuation/dis-
persion model is referred to as the Kolsky-Futterman model Kol-
sky, 1956;Futterman, 1962. This model has linear frequency de-
pendence of the attenuation coefficient, whereas the deviation from
24
6
VP(km/s)
1
0
1
2
Elevation (km)
1
0
1
2
Elevation (km)
1
0
1
2
E
l
evation (
k
m)
1
0
1
2
0510
15
Elevation (km)
Distance (
k
m)
a)
b)
c)
d)
Figure 12. Two-dimensional seismic imaging of a thrust belt in the
southern Apennines, Italy, from long-offset data by frequency-do-
main FWI. Thirteen frequencies from 6 to 20 Hz were inverted suc-
cessively. aStarting model for FWI developed by FATT Improta
et al., 2002.b–dFWI model after inversion of bthe starting
6-Hz, c10-Hz, and d20-Hz frequencies Ravaut et al., 2004.
45°
90°
135°
180°
225°
270°
315°
45°
90°
135°
180°
225°
270°
315°
θ
=0°
45°
90°
135°
180°
225°
270°
315°
ρ
ρ
θ
=0°
θ
=0°
I
P
V
P
I
PV
P
a)
b)
c)
Figure 13. Radiation pattern of different parameter classes in acous-
tic FWI. aIP-
;bIP-VP;cVP-
.
Full-waveform inversion WCC143
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
constant-phase velocity is accounted for through a term that varies as
the logarithm of the frequency. This model is obtained by imposing
causality on the wave pulse and assuming the absorption coefficient
is strictly proportional to the frequency over a restricted range of fre-
quencies.
Using the Kolsky-Futterman model, the complex velocity c
¯
is giv-
en by
c
¯
c
11
Qlog
/
r兲兩
isgn
2Q
1
,33
where Qis the attenuation factor and
ris a reference frequency.
In frequency-domain FWI, the real and imaginary parts of the ve-
locity can be processed as two independent real-model parameters
without any particular implementation difficulty. However, the re-
construction of attenuation is an ill-posed problem. Ribodetti et al.
2000show that in the framework of the ray Born inversion,
P-wave velocity and Qare uncoupled i.e., the Hessian is nonsingu-
laronly if a reflector is illuminated from both sides with circular
acquisition as in medical imaging; see Figure 5b of Ribodetti et al.,
2000. In contrast, the Hessian becomes singular for a surface acqui-
sition. Mulder and Hak 2009also conclude that VPand Qcannot be
imaged simultaneously from short-aperture data because the Hilbert
transform with respect to depth of the complex-valued perturbation
model for VPand Qproduces the same wavefield perturbation in the
framework of the Born approximation as the original perturbation
model. Despite this theoretical limitation, preconditioning of the
Hessian is investigated by Hak and Mulder 2008to improve the
convergence of the joint inversion for VPand Q.
Assessment and application of viscoacoustic frequency-domain
FWI is presented at various scales by Liao and McMechan 1995,
1996,Song et al. 1995,Hicks and Pratt 2001,Pratt et al. 2005,
Malinowsky et al. 2007, and Smithyman et al. 2008.Kamei and
Pratt 2008recommend inversion for VPonly in a first step and then
joint inversion for VPand Qin a second step because the reliability of
the attenuation reconstruction strongly depends on the accuracy of
the starting VPmodel. Indeed, accurate VPmodels are required be-
fore reconstructing Qso that the inversion can discriminate the in-
trinsic attenuation from the extrinsic attenuation.
Elastic parameters
A limited number of applications of elastic FWI have been pro-
posed. Because VPis the dominant parameter in elastic FWI, Taran-
tola 1986recommends inversion first for VPand IP, second for VS
and IS, and finally for density. This strategy might be suitable if the
footprint of the S-wave velocity structure on the seismic wavefield is
sufficiently small. This hierarchical strategy over parameter classes
is illustrated by Sears et al. 2008, who assess time-domain FWI of
multicomponent OBC data with synthetic examples. They highlight
how the behavior of FWI becomes ill-posed for S-wave velocity re-
construction when the S-wave velocity contrast at the sea bottom is
small. In this case, the S-wave velocity structure has a minor foot-
print on the seismic wavefield because the amount of PS conversion
is small at the sea bottom. In this configuration, they recommend in-
version first for VP, using only the vertical component; second for VP
and VSfrom the vertical component; and finally for VS, using both
components. The aim of the second stage is to reconstruct the inter-
mediate wavelengths of the S-wave velocity structure by exploiting
the AVO behavior of the P-waves.
In contrast, Brossier et al. 2009aconclude that joint inversion
for VPand VSwith judicious hierarchical data preconditioning by
time damping is necessary for inversion of land data involving both
body waves and surface waves. The strong sensitivity of the high-
amplitude surface waves to the near-surface S-wave velocity struc-
ture requires inversion for VSduring the early stages of the inversion.
This makes the elastic inversion of onshore data highly nonlinear
when the surface waves are preserved in the data.
A recent application of elastic FWI to a gas field in China is pre-
sented by Shi et al. 2007. They invert for the Lamé parameters and
unambiguously image Poisson’s ratio anomalies associated with the
presence of gas. They accelerate the convergence of the inversion by
computing an efficient step length using an adaptive controller based
on the theory of model-reference nonlinear control. Several logs
available along the profile confirm the reliability of this gas-layer
reconstruction.
Anisotropy
Reconstruction of anisotropic parameters by FWI is probably one
of the most undeveloped and challenging fields of investigation. Ver-
tically transverse isotropy VTIor tilted transversely isotropic
TTImedia are generally considered a realistic representation of
geologic media in oil and gas exploration, although fractured media
require an orthorhombic description Tsvankin, 2001. The normal-
moveout NMOP-wave velocity in VTI media depends on only two
parameters: the NMO velocity for a horizontal reflector VNMO0
VP012
and the
/12
parameter Alkhali-
fah and Tsvankin, 1995, which is a combination of Thomsen’s pa-
rameters and
Thomsen, 1986. The dependency of NMO veloc-
ity in VTI media on a limited subset of anisotropic parameters sug-
gests that defining the parameter classes to be involved in FWI will
be a key task. Another issue will be to assess to what extent FWI can
be performed in the acoustic approximation knowing that acoustic
media are by definition isotropic Grechka et al., 2004. The kine-
matic and dynamic accuracy of an acoustic TTI wave equation for
FWI is discussed in Operto et al. 2009.
A feasibility study of FWI in VTI media for crosswell acquisitions
is presented in Barnes et al. 2008. They invert for five parameter
classes VP,VS, density,
, and and show reliable reconstruc-
tion of the classes, even with noisy data. Pratt et al. 2001,2008ap-
ply anisotropic FWI to crosshole real data. The results of Pratt et al.
2001highlight the difficulty in discriminating layer-induced an-
isotropy from intrinsic anisotropy in FWI.
Further investigations of anisotropic FWI in the case of surface
seismic data are required. In particular, the benefit of wide apertures
in resolving as many anisotropic parameters as possible needs to be
investigated Jones et al., 1999.
Three-dimensional FWI
Because of the continuous increase in computational power and
the evolution of acquisition systems toward wide-aperture and wide-
azimuth acquisition, 3D acoustic FWI is feasible today. In three di-
mensions, the computational burden of multisource seismic model-
ing is one of the main issues. The pros and cons of time-domain ver-
sus frequency-domain seismic modeling for FWI have been dis-
cussed. Another issue is assessing the impact of azimuth coverage on
FWI. Sirgue et al. 2007show the footprint of the azimuth coverage
in 3D surveys on the velocity model reconstructed by FWI. Their im-
aging confirms the importance of wide-azimuth surveys for FWI of
coarse acquisitions such as node surveys.
WCC144 Virieux and Operto
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
Most 3D FWI applications have been limited to low frequencies
7Hz. At these frequencies, FWI can be seen as a tool for veloci-
ty-model-building rather than a self-contained seismic-imaging tool
that continuously proceeds from the velocity-model-building task to
the migration task through the continuous sampling of wavenum-
bers Ben Hadj Ali et al., 2008b.
Ben Hadj Ali et al. 2008aapply a frequency-domain FWI algo-
rithm to a series of synthetic data sets. The forward problem is solved
in the frequency domain with a massively parallel direct solver.
Plessix 2009presents an application to real ocean-bottom-seis-
mometer OBSdata. Seismic modeling is performed in the frequen-
cy domain with an iterative solver. The inverted frequencies range
between 2 and 5 Hz. The inversion converges to a similar velocity
model down to the top of a salt body using two different starting ve-
locity models, with one a simple velocity-gradient model. An appli-
cation to a real 3D streamer data set is presented by Warner et al.
2008for the imaging of a shallow channel. They perform seismic
modeling in the frequency domain using an iterative solver Warner
et al., 2007.
Three-dimensional time-domain FWI is developed by Vigh and
Starr 2008a, in which the input data in FWI are plane-wave gathers
rather than shot gathers. The main motivation behind the use of
plane-wave shot gathers is to mitigate the computational burden by
decimating the volume of data. The computational cost is reduced by
one order of magnitude for 2D applications and by a factor 3 for 3D
applications when the plane-wave-based approach is used instead of
the shot-based approach.
Sirgue et al. 2009apply 3D frequency-domain FWI to the hy-
drophone component of OBC data from the shallow-water Valhall
field. Frequencies between 3.5 and 7 Hz are inverted successively,
using a starting model built by ray-based reflection tomography.
They successfully image a complex network of shallow channels at
150 m depth and a gas cloud between 1000 and 2500 m depth. Al-
though the spacing between the cables is as high as 300 m, a limited
footprint of the acquisition is visible in the reconstructed models.
Comparisons of depth-migrated sections computed from the reflec-
tion tomography model and the FWI velocity model show the im-
provements provided by FWI, both in the shallow structure and at
the reservoir level below the gas cloud. The step-change improve-
ment in the quality of the depth-migrated image results from the
high-resolution nature of the velocity model from FWI and the ac-
counting of the intrabed multiples by the two-way wave-equation
modeling engine. Comparisons between depth slices across the
channels and the gas cloud extracted from the reflection tomography
and the FWI models highlight the significant resolution improve-
ment provided by FWI Figure 14.
Solving large-scale 3D elastic problems is probably beyond our
current tools because of the computational burden of 3D elastic
modeling for many sources. This has prompted several studies to de-
velop strategies to mitigate the number of forward simulations re-
quired during migration or FWI of large data sets. One of these ap-
proaches stacks the seismic sources before modeling Capdeville et
al., 2005. Because the relationship between the seismic wavefield
and the source is linear, stacking sources is equivalent to emitting
each source simultaneously instead of independently. This assem-
3
4
5
6
7
8
9
10
11
3456789101112131415161718
1800
2000
2200
2400
3
4
5
6
7
8
9
10
11
3456789101112131415161718
1700
1800
1900
Dip
l
ine (
k
m)
Crossline (km)
Crossline (km)
V
P(km/s)
Dipline (km)
3
4
5
6
7
8
9
10
11
3456789101112131415161718
1
700
1
800
1
900
Crossline (km)
V
P(km/s)
V
P(km/s)
3
4
5
6
7
8
9
10
11
3456789101112131415161718
1800
2000
2200
2400
Crossline (km)
V
P(km/s)
a)
b)
c)
d)
Figure 14. Imaging of the Valhall field by 3D FWI. Depth slices extracted from the velocity models. a,cBuilt by ray-based reflection tomogra-
phy. b,dBuilt by 3D FWI. c,dThe shallow slice at 150 m depth shows a complex network of channels in the FWI model, although the deeper
slice at 1050 m depth shows a much higher resolution of the top of the gas cloud in the FWI model image courtesy L. Sirgue and O. I. Barkved,
BP.
Full-waveform inversion WCC145
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
blage generates artifacts in the imaging that arise from the undesired
correlation between each independent source wavefield with the
back-propagated residual wavefields associated with the other
sources of the stack. Applying some random-phase encoding to each
source before assemblage mitigates these artifacts. The method orig-
inally was applied to migration algorithms Romero et al., 2000.
Promising applications to time- and frequency-domain FWI are pre-
sented in Krebs et al. 2009and Ben Hadj Ali et al. 2009a,2009b.
Alternatively, Herrmann et al. 2009propose to recover the source-
separated wavefields from the simultaneous simulation before FWI.
An illustration of the source-encoding technique is provided in
Figure 15, in which a dip section of the overthrust model is built
three ways: by conventional frequency-domain FWI i.e., without
source assemblage; Figure 15b, by FWI with source assemblage but
without phase encoding Figure 15c, and by FWI with source as-
semblage and phase encoding Figure 15d. The models were ob-
tained by successive inversions of four groups of two frequencies
ranging between 3.5 and 20 Hz. The number of shots was 200, and
no noise was added to the data. The number of iterations per frequen-
cy group to obtain the final FWI models without and with the source
assemblage was 15 and 200, respectively. The time to compute the
model of Figure 15d was seven times less than the time to build the
model of Figure 15b. More details can be found in Ben Hadj Ali et al.
2009. If the phase-encoding technique is seen as sufficiently ro-
bust, especially in the presence of noise, it is likely that 3D elastic
FWI can be viewed in the near future using sophisticated modeling
engines.
DISCUSSION
Most of the FWI methods presented and assessed in the literature
are based on the local least-squares optimization formulation, in
which the misfit between the observed seismograms and the mod-
eled ones are minimized in the time domain or in the frequency do-
main. Without very low frequencies 1Hz, it remains very diffi-
cult to obtain reliable results from these approaches when consider-
ing real data, especially at high frequencies. Clearly, new formula-
tions of FWI are needed to proceed further.
Recent improvements relate to Hessian estimation, which has
been shown to be quite important for better convergence toward the
solution. A systematic strategy since the beginning of FWI investi-
gations has been the progressive introduction of the entire content of
seismograms through multiscale approaches to partially prevent the
convergence toward secondary minima. Transformingthe data at the
Table 1. Nomenclature, listed as introduced in the text.
Symbol Description
xSpatial coordinates m
t,f,
2
/fTime s, frequency Hz, circular frequency
rad/s
k,kc/fWavenumber vector, wavenumber 1/m,
where cis wavespeed
1/kWavelength m
M,A,BMass, stiffness, impedance matrices in
the wave equation
ux,t/
,sx,t/
Wavefield solution of the wave equation and
seismic source
VP,VS,
P-wave and S-wave velocities m/s, density
kg/m3
QAttenuation factor for P-waves
,Anisotropic Thomsen’s parameters
m,m0,mUpdated and starting FWI models,
perturbation model
dobs,dcalmRecorded and modeled data
ddobs dcalmData-misfit vector
Cm,CmMisfit function, gradient of the misfit
function
Jd/mSensitivity or Fréchet derivative matrix
Step length in gradient methods
nIteration number in FWI
HaJJApproximate Hessian
pn,
nDescent direction and Polak and Ribière
coefficient in conjugate
gradients
WdSd
tSdWeighting operators in the data space
WmSm
tSmWeighting operators in the model space
Damping in damped least-squares FWI
fVirtual source in FWI for diffractor m
s,rSource and receiver indices
Diffraction or aperture angle
Damping coefficient for time damping of
seismograms
02468101214161820
06000
15000
24000
33000
42000
Distance (
k
m)
Depth (km)
V
P
(m/s)
02468101214161820
06000
15000
24000
33000
42000
02468101214161820
06000
15000
24000
33000
42000
02468101214161820
06000
15000
24000
33000
42
000
V
P
(m/s)V
P
(m/s)V
P
(m/s)
Depth (km)Depth (km)Depth (km)
a)
b)
c)
d)
Figure 15. Application of source encoding in FWI; 200 sources and
receivers were on the surface. aDip section of the overthrust
model. bFWI model obtained without source assemblage and
phase encoding; 15 iterations per frequency were computed. cFWI
model after assemblage of all the sources in one super shot. No phase
encoding was applied. dFWI model obtained with source assem-
blage and phase encoding; 200 iterations per frequency were com-
puted image courtesy of H. Ben Hadj Ali.
WCC146 Virieux and Operto
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
different stages of the local inversion procedure particle displace-
ment, particle velocity, logarithms of these quantities, divergence
can also provide benefits for global convergence of the optimization
procedure.
Do we need to mitigate the nonlinearity of the inverse problem
through more sophisticated search strategies such as semiglobal ap-
proaches or even global sampling of the model space? Because this
search is quite computationally intensive, should we reduce the di-
mensionality of the parameter models? To perform such global
searches, should we speed up the forward model at the expense of its
precision? If so, can we consider these procedures as intermediate
strategies for approaching the global minimum?
A pragmatic workflow, from low to high spatial-frequency con-
tent, should be expected to move hierarchically from imaging the
large medium wavelengths to the short medium wavelengths.A step-
by-step procedure for extracting the information, starting from trav-
eltimes and proceeding to true amplitudes, might include many in-
termediate steps for interpreting new observables polarization, ap-
parent velocity, envelope.
Another aspect is the quality control of a reconstruction. An ob-
jective analysis for uncertainty estimation is important and might
rely on semiglobal investigations once the global minimum has been
found. Various strategies can be considered that would be manage-
able as statistical procedures bootstrapping or jackknifing tech-
niquesor as local nondifferential approaches simplex, simulation
annealing, genetic algorithms.
Finally, solving large-scale 3D elastic problems remains beyond
our present tools, although one must be aware that these aids are
right around the corner. Because massive data acquisition for 3D re-
construction has been achieved, we indeed expect an improvement
in our data-crunching for high-resolution imaging.
An appealing reconstruction is 4D imaging, which is based on
time-lapse evolution of targets inside the earth. Differential data are
available, providing us with new information for tracking the evolu-
tion of the subsurface parameters. Thus, fluid tracking and variations
in solid-rock matrices are possible challenges for FWI in the near fu-
ture.
CONCLUSION
FWI is the last-course procedure for extracting information from
seismograms. We have shown the conceptual efforts that have been
carried out over the last 30 years to provide FWI as a possible tool for
high-resolution imaging. These efforts have been focused on devel-
opment of large-scale numerical optimization techniques, efficient
resolution of the two-way wave equation, judicious model parame-
terization for multiparameter reconstruction, multiscale strategies to
mitigate the ill-posedness of FWI, and specific waveform-inversion
data preprocessing.
FWI is mature enough today for prototype application to 3D real
data sets. Although applications to 3D real data have shown promis-
ing results at low frequencies 7Hz, it is still unclear to what ex-
tent FWI can be applied efficiently at higher frequencies. To answer
this question, a more quantitative understanding of FWI sensitivity
to the accuracy of the starting model, to the noise, and to the ampli-
tude accuracies is probably required.
If FWI remains limited to low frequencies, it will remain a tool to
build background models for migration. In the opposite case, FWI
will tend toward a self-contained processing workflow that can re-
unify macromodel building and migration tasks.
The present is exciting because realistic applications are becom-
ing possible right now. However, new strategies must be found to
make this technique as attractive as the scientific issues require.
Fields of investigation should address the need to speed up the for-
ward problem by means of providing new hardware GPUsand
software compressive sensing, defining new minimization criteria
in the model and data spaces, and incorporating more sophisticated
wave phenomena attenuation, elasticity, anisotropyin modeling
and inversion.
ACKNOWLEDGMENTS
We thank L. Sirgue and O. I. Barkved from BP for providing Fig-
ure 14 and for permission to present it. L. Sirgue is acknowledged for
editing part of the manuscript. We thank C. Shin Seoul National
Universityand Y. H. Cha Seoul National Universityfor providing
Figure 11. We thank A. Ribodetti Géosciences Azur-IRDfor dis-
cussion on attenuation reconstruction. We are grateful to R. Brossier
GéosciencesAzur-UNSAand H. Ben HadjAli Géosciences Azur-
UNSAfor providing Figures 8,9, and 15 and for fruitful discus-
sions on many aspects of FWI discussed in this study, in particular
multiparameter FWI, norms, Rytov approximation, and source en-
coding. We would like to thank R. E. Plessix Shellfor his explana-
tions of the adjoint-state method. J. Virieux thanks E. Canet LGIT
and A. Sieminski LGITfor an interesting discussion on the inverse
problem and its relationship with assimilation techniques. We would
like to thank S. Buske University of Berlin, T. Nemeth Chevron,
I. Lecomte NORSAR, and V. Sallares CMIMA-CSIC Barcelona
for their encouragement during the preparation of the manuscript.
Finally, we thank I. Lecomte and T. Nemeth for the time they dedi-
cated to the review of the manuscript. We thank the sponsors of the
SEISCOPE consortium BP, CGGVeritas, ExxonMobil, Shell, and
Totalfor their support.
APPENDIX A
APPLICATION OF THE ADJOINT-STATE
METHOD TO FWI
In this appendix, we provide some guidelines for the derivation
of the gradient of the misfit function equation 24with the adjoint-
state method and Lagrange multipliers. The reader is referred to No-
cedal and Wright 1999for a review of constrained optimization
and to Plessix 2006for a review of the adjoint-state method and its
application to FWI.
First, we introduce the Lagrangian function Lcorresponding to
the misfit function Caugmented with equality constraints:
Lu
¯
,m,
1
2PdobsRu
¯
兲兩PdobsRu
¯
兲典
Bmu
¯
s典共A-1
The equality constraints correspond to the forward-problem equa-
tion, namely, the state equation, which must be satisfied by the seis-
mic wavefield. A realization uof the state equation is the so-called
state variable. In equation A-1, we introduce the variable u
¯
to distin-
guish any element of the state variable space from a realization of the
state equation Plessix, 2006.
The vector
, the dimension of which is that of the wavefield u,is
the Lagrange multiplier; it corresponds to the adjoint-state variables.
Full-waveform inversion WCC147
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
In the framework of the theory of constrained optimization, the first-
order necessary optimality conditions known as the Karush-Kuhn-
Tucker KKTconditions state that the solution of the optimization
problem is obtained at the stationary points of L.
The first condition,
L/
u
¯
cste,mcste 0, leads to the for-
ward-problem equation: Bmus. Resolving the state equation for
mm0provides the incident wavefield for FWI.
The second condition,
L/
u
¯
cste,mcste 0, leads to the so-
called adjoint-state equation, expressed as
Lu
¯
,m,
u
¯
PdobsRu
¯
Bm
0.A-2
For the derivation of equation A-2, we use the fact that
,Bmu
¯
Bm
,u
¯
and that the source does not depend on u
¯
. Choosing
mm0and u
¯
um0in equation A-2 leads to
Bm0
Pd0,A-3
which can be rewritten equivalently as
*B1m0Pd0
*,A-4
where we exploit the fact that B1is symmetric by virtue of the spa-
tial reciprocity of Green’s functions. The adjoint-state variables are
computed by solving a forward problem for a composite source
formed by the conjugate of the residuals, which is equivalent to back
propagation of the residuals in the model.
The third condition,
L/
mu
¯
cste,
cste 0, defines the mini-
mum of Lin a comparable way as for the unconstrained minimiza-
tion of the misfit function C. We have
Lu
¯
,m,
m
Bm
mu
¯
.A-5
For any realization of the forward problem u,Lu,m,
Cm.
Therefore, equation A-5 gives the expression of the desired gradient
of Cas a function of the state variable and adjoint-state variable
when u
¯
u:
C
m
Bm
mu
.A-6
Inserting the expression of
equation A-4into equation A-3 and
choosing mm0gives the expression of the gradient of Cat the
point m0in the opposite direction of which a minimum of Cis sought
for in FWI:
Cm0
m
utm0
Btm0
mB1m0Pd0
*.A-7
Equation A-7 is equivalent to equation 24 in the main text.
REFERENCES
Abubakar, A., W. Hu, T. Habashy, and P. van den Berg, 2009, A finite-differ-
ence contrast source inversion algorithm for full-waveform seismic appli-
cations: Geophysics, this issue.
Akcelik, V., 2002, Multiscale Newton-Krylov methods for inverse acoustic
wave propagation: Ph.D. thesis, Carnegie Mellon University.
Aki, K., and P. Richards, 1980, Quantitative seismology: Theory and meth-
ods: W.H. Freeman & Co.
Alerini, M., S. Le Bégat, G. Lambaré, and R. Baina, 2002, 2D PP- and PS-ste-
reotomography for a multicomponent datset: 72nd Annual International
Meeting, SEG, Expanded Abstracts, 838–841.
Alkhalifah, T., and I. Tsvankin, 1995, Velocity analysis for transversely iso-
tropic media: Geophysics, 60, 1550–1566.
Amundsen, L., 1991, Comparison of the least-squares criterion and the
Cauchy criterion in frequency-wavenumber inversion: Geophysics, 56,
2027–2035.
Amundsen, L., and B. Ursin, 1991, Frequency-wavenumber inversion of
acoustic data: Geophysics, 56, 1027–1039.
Askan, A., 2006, Full waveform inversion for seismic velocity and anelastic
losses in heterogeneous structures: Ph.D. thesis, Carnegie Mellon Univer-
sity.
Askan, A., V. Akcelik, J. Bielak, and O. Ghattas, 2007, Full waveform inver-
sion for seismic velocity and anelastic losses in heterogeneous structures:
Bulletin of the Seismological Society of America, 97, 1990–2008.
Askan, A., and J. Bielak, 2008, Full anelastic waveform tomography includ-
ing model uncertainty: Bulletin of the Seismological Society of America,
98, 2975–2989.
Barnes, C., and M. Charara, 2008, Full-waveform inversion results when us-
ing acoustic approximation instead of elastic medium: 78th Annual Inter-
national Meeting, SEG, Expanded Abstracts, 1895–1899.
Barnes, C., M. Charara, and T.Tsuchiya, 2008, Feasibility study for an aniso-
tropic full waveform inversion of cross-well data: Geophysical Prospect-
ing, 56, 897–906.
Baysal, E., D. Kosloff, and J. Sherwood, 1983, Reverse time migration: Geo-
physics, 48, 1514–1524.
Bednar, J. B., C. Shin, and S. Pyun, 2007, Comparison of waveform inver-
sion, Part 3: Amplitude approach: Geophysical Prospecting, 55, 477–485.
Ben Hadj Ali, H., S. Operto, and J. Virieux, 2008a, Velocity model building
by 3D frequency-domain full-waveform inversion of wide-aperture seis-
mic data: Geophysics, 73, no. 5, VE101–VE117.
Ben Hadj Ali, H., S. Operto, J. Virieux, and F. Sourbier, 2008b, 3D frequen-
cy-domain full-waveform tomography based on a domain decomposition
forward problem: 78th Annual International Meeting, SEG, ExpandedAb-
stracts, 1945–1949.
——–, 2009a, Efficient 3D frequency-domain full waveform inversion with
phase encoding: 71st Conference & TechnicalExhibition, EAGE, Extend-
ed Abstracts, 5812.
——–, 2009b, Three-dimensional frequency-domain full-waveform inver-
sion with phase encoding: 79th Annual International Meeting, SEG, Ex-
panded Abstracts, 2288–2292.
Beydoun, W. B., and A. Tarantola, 1988, First Born and Rytov approxima-
tion: Modeling and inversion conditions in a canonical example: Journal
of the Acoustical Society ofAmerica, 83, 1045–1055.
Beylkin, G., 1985, Imaging of discontinuities in the inverse scattering prob-
lem by inversion of a causal generalized Radon transform: Journal of
Mathematical Physics, 26, 99–108.
Beylkin, G., and R. Burridge, 1990, Linearized inverse scattering problems
in acoustics and elasticity: WaveMotion, 12, 15–52.
Billette, F., and G. Lambaré, 1998, Velocity macro-model estimation from
seismic reflection data by stereotomography: Geophysical Journal Inter-
national, 135, 671–680.
Billette, F., S. Le Bégat, P. Podvin, and G. Lambaré, 2003, Practical aspects
and applications of 2D stereotomography: Geophysics, 68, 1008–1021.
Biondi, B., and W. Symes, 2004, Angle-domain common-image gathers for
migration velocity analysis by wavefield-continuation imaging: Geophys-
ics, 69, 1283–1298.
Bleibinhaus, F., J. A. Hole, T. Ryberg, and G. S. Fuis, 2007, Structure of the
California Coast Ranges and San Andreas fault at SAFOD from seismic
waveform inversion and reflection imaging: Journal of Geophysical
Research, doi: 10.1029/2006JB004611.
Bleistein, N., 1987, On the imaging of reflectors in the earth: Geophysics, 52,
931–942.
Bouchon, M., M. Campillo, and S. Gaffet, 1989, A boundary integral equa-
tion Discrete wavenumber representation method to study wave propa-
gation in multilayered media having irregular interfaces: Geophysics, 54,
1134–1140.
Brenders, A. J., and R. G. Pratt, 2007a, Efficient waveform tomography for
lithospheric imaging: Implications for realistic 2D acquisition geometries
and low frequency data: Geophysical Journal International, 168, 152–170.
——–, 2007b, Full waveform tomography for lithospheric imaging: Results
from a blind test in a realistic crustal model: Geophysical Journal Interna-
tional, 168, 133–151.
——–, 2007c, Waveform tomography of marine seismic data: What can lim-
ited offset offer?: 75thAnnual International Meeting, SEG, Expanded Ab-
stracts, 3024–3029.
Brossier, R., S. Operto, and J. Virieux, 2009a, Seismic imaging of complex
structures by 2D elastic frequency-domain full-waveform inversion: Geo-
physics, this issue.
——–, 2009b, Two-dimensional seismic imaging of the Valhall model from
synthetic OBC data by frequency-domain elastic full-waveform inver-
sion: 79th Annual International Meeting, SEG, Expanded Abstracts,
2293–2297.
Brossier, R., J. Virieux, and S. Operto, 2008, Parsimonious finite-volume fre-
WCC148 Virieux and Operto
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
quency-domain method for 2D P-SV wave modelling: Geophysical Jour-
nal International, 175, 541–559.
Bube, K. P., and T.Nemeth, 2007, Fast line searches for the robust solution of
linear systems in the hybrid 1/2and Huber norms: Geophysics, 72, no. 2,
A13–A17.
Bunks, C., F. M. Salek, S. Zaleski, and G. Chavent, 1995, Multiscale seismic
waveform inversion: Geophysics, 60, 1457–1473.
Capdeville, Y., Y. Gung, and B. Romanowicz, 2005, Towards global earth to-
mography using the spectral element method: Atechnique based on source
stacking: Geophysical Journal International, 162, 541–554.
Carcione, J. M., G. C. Herman, and A. P. E. ten Kroode, 2002, Seismic mod-
eling: Geophysics, 67, 1304–1325.
Cary, P., and C. Chapman, 1988, Automatic 1-D waveform inversion of ma-
rine seismic refraction data: Geophysical Journal of the Royal Astronomi-
cal Society, 93, 527–546.
Chapman, C., 1985, Ray theory and its extension: WKBJ and Maslov seis-
mograms: Journal of Geophysics, 58, 27–43.
Chavent, G., 1974, Identification of parameter distributed systems, in R.
Goodson and M. Polis, eds., Identification of function parameters in par-
tial differential equations: American Society of Mechanical Engineers,
31–48.
Chen, P., T. Jordan, and L. Zhao, 2007, Full three-dimensional tomography:
A comparison between the scattering-integral and adjoint-wavefield meth-
ods: Geophysical Journal International, 170, 175–181.
Chironi, C., J. V. Morgan, and M. R. Warner, 2006, Imaging of intrabasalt and
subbasalt structure with full wavefield seismic tomography: Journal of
Geophysical Research, 11, B05313, doi: 10.1029/2004JB003595.
Choi, Y., D.-J. Min, and C. Shin, 2008, Two-dimensional waveform inver-
sion of multi-component data in acoustic-elastic coupled media: Geophys-
ical Prospecting, 56, 863–881.
Claerbout, J., 1971, Toward a unified theory of reflector mapping: Geophys-
ics, 36, 467–481.
——–, 1976, Fundamentals of geophysical data processing: McGraw-Hill
Book Co., Inc.
Claerbout, J., and S. Doherty, 1972, Downward continuation of moveout cor-
rected seismograms: Geophysics, 37, 741–768.
Crase, E., A. Pica, M. Noble, J. McDonald, and A. Tarantola, 1990, Robust
elastic non-linear waveform inversion: Application to real data: Geophys-
ics, 55, 527–538.
Danecek, P., and G. Seriani, 2008, An efficient parallel Chebyshev pseudo-
spectral method for large-scale 3D seismic forward modelling: 70th Con-
ference & TechnicalExhibition, EAGE, Extended Abstracts, P046.
de Hoop, A., 1960, A modification of Cagniard’s method for solving seismic
pulse problems: Applied Scientific Research, 8, 349–356.
Dessa, J. X., and G. Pascal, 2003, Combined traveltime and frequency-do-
main seismic waveform inversion: Acase study on multi-offset ultrasonic
data: Geophysical Journal International, 154, 117–133.
Devaney, A. J., 1982, A filtered backprojection algorithm for diffraction to-
mography: Ultrasonic Imaging, 4, 336–350.
Devaney, A. J., and D. Zhang, 1991, Geophysical diffraction tomography in
a layered background: WaveMotion, 14, 243–265.
Djikpéssé, H., and A. Tarantola, 1999, Multiparameter 1-norm waveform
fitting: Interpretation of Gulf of Mexico reflection seismograms: Geo-
physics, 64, 1023–1035.
Docherty, P., R. Silva, S. Singh, Z.-M. Song, and M. Wood, 2003, Migration
velocity analysis using a genetic algorithm: Geophysical Prospecting, 45,
865–878.
Dumbser, M., and M. Kaser, 2006, An arbitrary high-order discontinuous
Galerkin method for elastic waves on unstructured meshes II. The
three-dimensional isotropic case: Geophysical Journal International, 167,
319–336.
Dummong, S., K. Meier, D. Gajewski, and C. Hubscher, 2008, Comparison
of prestack stereotomography and NIP wave tomography for velocity
model building: Instances from the Messinian evaporites: Geophysics, 73,
no. 5, VE291–VE302.
Ellefsen, K., 2009, A comparison of phase inversion and traveltime tomogra-
phy for processing of near-surface refraction traveltimes: Geophysics, this
issue.
Epanomeritakis, I., V. Akçelik, O. Ghattas, and J. Bielak, 2008, A Newton-
CG method for large-scale three-dimensional elastic full waveform seis-
mic inversion: Inverse Problems, 24, 1–26.
Erlangga, Y. A., and F. J. Herrmann, 2008, An iterative multilevel method for
computing wavefields in frequency-domain seismic inversion: 78th Annu-
al International Meeting, SEG, Expanded Abstracts, 1956–1960.
Ernst, J. R., A. G. Green, H. Maurer, and K. Holliger, 2007, Application of a
new 2D time-domain full-waveform inversion scheme to crosshole radar
data: Geophysics, 72, no. 5, J53–J64.
Fichtner, A., B. L. N. Kennett, H. Igel, and H. P. Bunge, 2008, Theoretical
background for continental- and global-scale full-waveform inversion in
the time-frequency domain: Geophysical Journal International, 175,
665–685.
Forgues, E., and G. Lambaré, 1997, Parameterization study for acoustic and
elastic ray Born inversion: Journal of Seismic Exploration, 6, 253–278.
Futterman, W., 1962, Dispersive body waves: Journal of Geophysics Re-
search, 67, 5279–5291.
Gao, F., A. R. Levander, R. G. Pratt, C. A. Zelt, and G. L. Fradelizio, 2006a,
Waveform tomography at a groundwater contamination site: Surface re-
flection data: Geophysics, 72, no. 5, G45–G55.
——–, 2006b, Waveform tomography at a groundwater contamination site:
VSP-surface data set: Geophysics, 71, no. 1, H1–H11.
Gauthier, O., J. Virieux, and A. Tarantola, 1986, Two-dimensional nonlinear
inversion of seismic waveforms: Numerical results: Geophysics, 51,
1387–1403.
Gazdag, J., 1978, Wave equation migration with the phase-shift method:
Geophysics, 43, 1342–1351.
Gelis, C., J. Virieux, and G. Grandjean, 2007, 2D elastic waveform inversion
using Born and Rytov approximations in the frequency domain: Geophys-
ical Journal International, 168, 605–633.
Gelius, L. J., I. Johansen, N. Sponheim, and J. J. Stamnes, 1991, A general-
ized diffraction tomography algorithm: Journal of the Acoustical Society
of America, 89, 523–528.
Gilbert, F., and A. Dziewonski, 1975, An application of normal mode theory
to the retrieval of structural parameters and source mechanisms from seis-
mic spectra: Philosophical Transactions of the Royal Society of London,
278, 187–269.
Graves, R., 1996, Simulating seismic wave propagation in 3D elastic media
using staggered-grid finite differences: Bulletin of the Seismological Soci-
ety of America, 86, 1091–1106.
Grechka, V., L. Zhang, and J. W. Rector III, 2004, Shear waves in acoustic an-
isotropic media: Geophysics, 69, 576–582.
Guitton, A., and W. W. Symes, 2003, Robust inversion of seismic data using
the Huber norm: Geophysics, 68, 1310–1319.
Gutenberg, B., 1914, Über erdbenwellen viia. beobachtungen an registri-
erungen von fernbeben in göttingen und folgerungen über die konstitution
des erdkörpers: Nachrichten von der Könglichen Gesellschaft der Wissen-
schaften zu Göttinge, Mathematisch-Physikalische Klasse, 125–176.
Ha, T.,W. Chung, and C. Shin, 2009, Waveforminversion using a back-prop-
agation algorithm and a Huber function norm: Geophysics, 74, no. 3, R15–
R24.
Hak, B., and W.Mulder, 2008, Preconditioning for linearized inversion of at-
tenuation and velocity perturbations: 78th Annual International Meeting,
SEG, Expanded Abstracts, 2036–2040.
Hansen, C., ed., 1998, Rank-defficient and discrete ill-posed problems
Numerical aspects of linear inversion: Society for Industrial and Applied
Mathematics.
Herrmann, F.J., Y. Erlangga, and T. T. Y. Lin, 2009, Compressive sensing ap-
plied to full-wave form inversion: 71st Conference & Technical Exhibi-
tion, EAGE, Extended Abstracts, S016.
Hicks, G. J., and R. G. Pratt, 2001, Reflection waveform inversion using local
descent methods: Estimating attenuation and velocity over a gas-sand de-
posit: Geophysics, 66, 598–612.
Hole, J. A., 1992, Nonlinear high-resolution three-dimensional seismic trav-
el time tomography: Journal of Geophysical Research, 97, 6553–6562.
Hu, W.,A. Abubakar, and T. M. Habashy, 2009, Simultaneous multifrequen-
cy inversion of full-waveform seismic data: Geophysics, 74, no. 2, R1–
R14.
Hustedt, B., S. Operto, and J. Virieux, 2004, Mixed-grid and staggered-grid
finite difference methods for frequency domain acoustic wave modelling:
Geophysical Journal International, 157, 1269–1296.
Ikelle, L., J. P. Diet, and A. Tarantola,1988, Linearized inversion of multioff-
set seismic reflection data in the
-kdomain: Depth-dependent reference
medium: Geophysics, 53, 50–64.
Improta, L., A. Zollo, A. Herrero, R. Frattini, J. Virieux, and P.DellÁversana,
2002, Seismic imaging of complex structures by non-linear traveltimes in-
version of dense wide-angle data: Application to a thrust belt: Geophysical
Journal International, 151, 264–278.
Jaiswal, P., C. Zelt, A. W. Bally, and R. Dasgupta, 2008, 2-D traveltime and
waveform inversion for improved seismic imaging: Naga thrust and fold
belt, India: Geophysical Journal International, doi: 10.1111/j.1365-
246X.2007.03691.x.
Jaiswal, P., C. Zelt, R. Dasgupta, and K. Nath, 2009, Seismic imaging of the
Naga thrust using multiscale waveform inversion: Geophysics, this issue.
Jannane, M., W.Beydoun, E. Crase, D. Cao, Z. Koren, E. Landa, M. Mendes,
A. Pica, M. Noble, G. Roeth, S. Singh, R. Snieder, A. Tarantola, and D.
Trézéguet, 1989, Wavelengths of earth structures that can be resolved from
seismic reflection data: Geophysics, 54, 906–910.
Jin, S., and R. Madariaga, 1993, Background velocity inversion by a genetic
algorithm: Geophysical Research Letters, 20, 93–96.
——–, 1994, Non-linear velocity inversion by a two-step Monte Carlo meth-
od: Geophysics, 59, 577–590.
Jin, S., R. Madariaga, J. Virieux, and G. Lambaré, 1992, Two-dimensional
asymptotic iterative elastic inversion: Geophysical Journal International,
108, 575–588.
Jo, C. H., C. Shin, and J. H. Suh, 1996, An optimal 9-point, finite-difference,
Full-waveform inversion WCC149
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
frequency-space 2D scalar extrapolator: Geophysics, 61, 529–537.
Jones, K., M. Warner, and J. Brittan, 1999, Anisotropy in multi-offset deep-
crustal seismic experiments: Geophysical Journal International, 138, 300–
318.
Kamei, R., and R. G. Pratt, 2008, Waveformtomography strategies for imag-
ing attenuation structure for cross-hole data: 70th Conference & Technical
Exhibition, EAGE, Extended Abstracts, F019.
Kennett, B. L. N., 1983, Seismic wave propagation in stratified media: Cam-
bridge University Press.
Klem-Musatov, K. D., and A. M. Aizenberg, 1985, Seismic modelling by
methods of the theory of edge waves: Journal of Geophysics, 57, 90–105.
Kolb P., F. Collino, and P. Lailly, 1986, Prestack inversion of 1-D medium:
Proceedings of the IEEE, 498–508.
Kolsky, H., 1956, The propagation of stress pulses in viscoelastic solids:
Philosophical Magazine, 1, 693–710.
Komatitsch, D., and J. P. Vilotte,1998, The spectral element method: An effi-
cient tool to simulate the seismic response of 2D and 3D geological struc-
tures: Bulletin of the Seismological Society of America, 88, 368–392.
Koren, Z., K. Mosegaard, E. Landa, P. Thore, and A. Tarantola, 1991, Monte
Carlo estimation and resolution analysis of seismic background velocities:
Journal of Geophysical Research, 96, 20289–20299.
Kormendi, F., and M. Dietrich, 1991, Non-linear waveform inversion of
plane-wave seismograms in stratified elastic media: Geophysics, 56, 664–
674.
Krebs, J., J. Anderson, D. Hinkley, R. Neelamani, A. Baumstein, M. D.
Lacasse, and S. Lee, 2009, Fast full-wavefield seismic inversion using en-
coded sources: Geophysics, this issue.
Lailly, P., 1983, The seismic inverse problem as a sequence of before stack
migrations: Conference on Inverse Scattering, Theory and Application,
Society for Industrial and Applied Mathematics, Expanded Abstracts,
206–220.
Lambaré, G., 2008, Stereotomography: Geophysics, 73, no. 5, VE25–VE34.
Lambaré, G., and M. Alérini, 2005, Semi-automatic picking PP-PS stereoto-
mography: Application to the syntheticValhall data set: 75thAnnual Inter-
national Meeting, SEG, Expanded Abstracts, 943–946.
Lambaré, G., S. Operto, P. Podvin, P. Thierry, and M. Noble, 2003, 3-D ray
Born migration/inversion Part 1: Theory: Geophysics, 68, 1348–1356.
Lambaré, G., J. Virieux, R. Madariaga, and S. Jin, 1992, Iterative asymptotic
inversion in the acoustic approximation: Geophysics, 57, 1138–1154.
Lecomte, I., 2008, Resolution and illumination analyses in PSDM: A ray-
based approach: The Leading Edge, 27, 650–663.
Lecomte, I., S.-E. Hamran, and L.-J. Gelius, 2005, Improving Kirchhoff mi-
gration with repeated local plane-wave imaging? ASAR-inspired signal-
processing approach in prestack depth imaging: Geophysical Prospecting,
53, 767–785.
Lee, K. H., and H. J. Kim, 2003, Source-independent full-waveform inver-
sion of seismic data: Geophysics, 68, 2010–2015.
Lehmann, I., 1936, P: Publications du Bureau Central Séismologique Inter-
national, A14, 87–115.
Levander, A. R., 1988, Fourth-order finite-difference P-SV seismograms:
Geophysics, 53, 1425–1436.
Liao, Q., and G. A. McMechan, 1995, 2.5D full-wavefield viscoacoustic in-
version: Geophysical Prospecting, 43, 1043–1059.
——–, 1996, Multifrequency viscoacoustic modeling and inversion: Geo-
physics, 61, 1371–1378.
Lions, J., 1972, Nonhomogeneous boundary value problems and applica-
tions: Springer-Verlag, Berlin.
Lucio, P. S., G. Lambaré, and A. Hanyga, 1996, 3D multivalued travel time
and amplitude maps: Pure and Applied Geophysics, 148, 113–136.
Luo, Y., T. Nissen-Meyer, C. Morency, and J. Tromp, 2009, Seismic model-
ing and imaging based upon spectral-element and adjoint methods: The
Leading Edge, 28, 568–574.
Malinowsky, M., and S. Operto, 2008, Quantitative imaging of the permo-
mesozoic complex and its basement by frequency domain waveform to-
mography of wide-aperture seismic data from the polish basin: Geophysi-
cal Prospecting, 56, 805–825.
Malinowsky, M., A. Ribodetti, and S. Operto, 2007, Multiparameter full-
waveform inversion for velocity and attenuation Refining the imaging
of a sedimentary basin: 69th Conference & Technical Exhibition, EAGE,
Extended Abstracts, P276.
Marfurt, K., 1984, Accuracy of finite-difference and finite-elements model-
ing of the scalar and elastic wave equation: Geophysics, 49, 533–549.
McMechan, G. A., and G. S. Fuis, 1987, Ray equation migration of wide-an-
gle reflections from southern Alaska: Journal of Geophysical Research,
92, 407–420.
Menke, W., 1984, Geophysical data analysis: Discrete inverse theory:Aca-
demic Press, Inc.
Miller, D., M. Oristaglio, and G. Beylkin, 1987, A new slant on seismic imag-
ing: Migration and integral geometry: Geophysics, 52, 943–964.
Min, D.-J., and C. Shin, 2006, Refraction tomography using a waveform-in-
version back-propagation technique: Geophysics, 71, no. 3, R21–R30.
Min, D.-J., C. Shin, R. G. Pratt, and H. S. Yoo, 2003, Weighted-averaging fi-
nite-element method for 2D elastic wave equations in the frequency do-
main: Bulletin of the Seismological Society of America, 93, 904–921.
Montelli, R., G. Nolet, F.Dahlen, G. Masters, E. Engdahl, and S. Hung, 2004,
Finite-frequency tomography reveals a variety of plumes in the mantle:
Science, 303, 338–343.
Mora, P. R., 1987, Nonlinear two-dimensional elastic inversion of multi-off-
set seismic data: Geophysics, 52, 1211–1228.
——–, 1988, Elastic wavefield inversion of reflection and transmission data:
Geophysics, 53, 750–759.
Mosegaard, K., and A. Tarantola,1995, Monte Carlo sampling of solutions to
inverse problems: Journal of Geophysical Research, 100, 12431–12447.
Mulder, W. A., and B. Hak, 2009, Simultaneous imaging of velocity and at-
tenuation perturbations from seismic data is nearly impossible: 71st Con-
ference & TechnicalExhibition, EAGE, Extended Abstracts, S043.
Nihei, K. T., and X. Li, 2007, Frequency response modelling of seismic
waves using finite difference time domain with phase sensitive detection
TD-PSD: Geophysical Journal International, 169, 1069–1078.
Nocedal, J., 1980, Updating quasi-Newton matrices with limited storage:
Mathematics of Computation, 35, 773–782.
Nocedal, J., and S. J. Wright, 1999, Numerical optimization: Springer.
Nolet, G., 1987, Seismic tomography with applications in global seismology
and exploration geophysics: D. Reidel Publ. Co.
Oldham, R., 1906, The constitution of the earth: Quarterly Journal of the
Geological Society of London, 62, 456–475.
Operto, S., G. Lambaré, P. Podvin, and P. Thierry, 2003, 3-D ray-Born migra-
tion/inversion, Part 2: Application to the SEG/EAGE overthrust experi-
ment: Geophysics, 68, 1357–1370.
Operto, S., C. Ravaut, L. Improta, J. Virieux, A. Herrero, and P. Dell’
Aversana, 2004, Quantitative imaging of complex structures from multi-
fold wide aperture seismic data: A case study: Geophysical Prospecting,
52, 625–651.
Operto, S., J. Virieux, P. Amestoy, J. Y. L’Excellent, L. Giraud, and H. Ben-
Hadj-Ali, 2007, 3D finite-difference frequency-domain modeling of vis-
coacoustic wave propagation using a massively parallel direct solver: A
feasibility study: Geophysics, 72, no. 5, SM195–SM211.
Operto, S., J. Virieux, J. X. Dessa, and G. Pascal, 2006, Crustal imaging from
multifold ocean bottom seismometers data by frequency-domain full-
waveform tomography: Application to the eastern Nankai trough: Journal
of Geophysical Research, doi: 10.1029/2005JB003835.
Operto, S., J. Virieux, A. Ribodetti, and J. E. Anderson, 2009, Finite-differ-
ence frequency-domain modeling of viscoacoustic wave propagation in
two-dimensional tilted transversely isotropic media: Geophysics, 74,no.
5, T75–T95, doi: 10.1190/1.3157243.
Operto, S., S. Xu, and G. Lambaré, 2000, Can we image quantitatively com-
plex models with rays?: Geophysics, 65, 1223–1238.
Paige, C. C., and M. A. Saunders, 1982a, Algorithm 583 LSQR: Sparse linear
equations and least squares problems: Association for Computing Ma-
chinery Transactions on Mathematical Software, 8, 195–209.
——–, 1982b, LSQR: An algorithm for sparse linear equations and sparse
least squares: Association for Computing Machinery Transactions on
Mathematical Software, 8, 43–71.
Pica, A., J. Diet, and A. Tarantola, 1990, Nonlinear inversion of seismic re-
flection data in laterally invariant medium: Geophysics, 55, 284–292.
Plessix, R.-E., 2006, A review of the adjoint-state method for computing the
gradient of a functional with geophysical applications: Geophysical Jour-
nal International, 167, 495–503.
——–, 2007, AHelmholtz iterative solver for 3D seismic-imaging problems:
Geophysics, 72, no. 5, SM185–SM194.
——–, 2009, 3D frequency-domain full-waveform inversion with an itera-
tive solver: Geophysics, this issue.
Polak, E., and G. Ribière, 1969, Note sur la convergence de méthodes de di-
rections conjuguées: Revue Française d’Informatique et de Recherche
Opérationnelle, 16, 35–43.
Pratt, R. G., 1990, Inverse theory applied to multi-source cross-hole tomog-
raphy. Part II: Elastic wave-equation method: Geophysical Prospecting,
38, 311–330.
——–, 1999, Seismic waveform inversion in the frequency domain, Part I:
Theory and verification in a physical scale model: Geophysics, 64,
888–901.
Pratt, R. G., and N. R. Goulty, 1991, Combining wave-equation imaging with
traveltime tomography to form high-resolution images from crosshole
data: Geophysics, 56, 204–224.
Pratt, R. G., F. Hou, K. Bauer, and M. Weber, 2005, Waveform tomography
images of velocity and inelastic attenuation from the Mallik 2002 cross-
hole seismic surveys, in S. R. Dallimore, and T. S. Collett, eds., Scientific
results from the Mallik 2002 gas hydrate production research well pro-
gram, Mackenzie Delta, Northwest Territories, Canada: Geological Sur-
vey of Canada.
Pratt, R. G., R. E. Plessix, and W. A. Mulder, 2001, Seismic waveform to-
mography: The effect of layering and anisotropy: 63rd Conference &
Technical Exhibition, EAGE, ExtendedAbstracts, P092.
Pratt, R. G., C. Shin, and G. J. Hicks, 1998, Gauss-Newton and full Newton
WCC150 Virieux and Operto
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
methods in frequency-space seismic waveform inversion: Geophysical
Journal International, 133, 341–362.
Pratt, R. G., L. Sirgue, B. Hornby, and J. Wolfe, 2008, Cross-well waveform
tomography in fine-layered sediments Meeting the challenges of aniso-
tropy: 70th Conference & Technical Exhibition, EAGE, Extended Ab-
stracts, F020.
Pratt, R. G., Z. M. Song, P. R. Williamson, and M. Warner, 1996, Two-dimen-
sional velocity model from wide-angle seismic data by wavefield inver-
sion: Geophysical Journal International, 124, 323–340.
Pratt, R. G., and W.Symes, 2002, Semblance and differential semblance opti-
misation for waveform tomography: Afrequency domain implementation:
Sub-basalt imaging conference, Expanded Abstracts, 183–184.
Pratt, R. G., and M. H. Worthington, 1990, Inverse theory applied to multi-
source cross-hole tomography, Part 1: Acoustic wave-equation method:
Geophysical Prospecting, 38, 287–310.
Press, W. H., B. P. Flannery, S. A. Teukolsky, and W. T. Veltering, 1986, Nu-
merical recipes: The art of scientific computing: Cambridge University
Press.
Prieux, V., S. Operto, R. Brossier, and J. Virieux,2009, Application of acous-
tic full waveform inversion to the synthetic Valhall model: 77thAnnual In-
ternational Meeting, SEG, Expanded Abstracts, 2268–2272.
Pyun, S., C. Shin, and J. B. Bednar, 2007, Comparison of waveform inver-
sion, Part 3: Phase approach: Geophysical Prospecting, 55, 465–475.
Ravaut, C., S. Operto, L. Improta, J. Virieux, A. Herrero, and P. Dell’
Aversana, 2004, Multi-scale imaging of complex structures from multi-
fold wide-aperture seismic data by frequency-domain full-wavefield in-
versions: Application to a thrust belt: Geophysical Journal International,
159, 1032–1056.
Ribodetti, A., S. Operto, J. Virieux, G. Lambaré, H.-P. Valéro, and D. Gibert,
2000, Asymptotic viscoacoustic diffraction tomography of ultrasonic lab-
oratory data: Atool for rock properties analysis: Geophysical Journal In-
ternational, 140, 324–340.
Ribodetti, A., and J. Virieux, 1996, Asymptotic theory for imaging the atten-
uation factors Qpand Qs: Conference of Aixe-les-Bains, INRIA— French
National Institute for Research in Computer Science and Control, Expand-
ed Abstracts, 334–353.
Riyanti, C. D., Y. A. Erlangga, R. E. Plessix, W. A. Mulder, C. Vuik, and C.
Oosterlee, 2006, A new iterative solver for the time-harmonic wave equa-
tion: Geophysics, 71, no. 5, E57–E63.
Riyanti, C. D., A. Kononov, Y. A. Erlangga, C. Vuik, C. W. Oosterlee, R. E.
Plessix, and W. A. Mulder, 2007, A parallel multigrid-based precondi-
tioner for the 3D heterogeneous high-frequency Helmholtz equation: Jour-
nal of Computational Physics, 224, 431–448.
Robertsson, J. O., B. Bednar, J. Blanch, C. Kostov, and D. J. van Manen,
2007, Introduction to the supplement on seismic modeling with applica-
tions to acquisition, processing and interpretation: Geophysics, 72, no. 5,
SM1–SM4.
Romero, L. A., D. C. Ghiglia, C. C. Ober, and S. A. Morton, 2000, Phase en-
coding of shot records in prestack migration: Geophysics, 65, 426–436.
Saad, Y., 2003, Iterative methods for sparse linear systems, 2nd ed.: Society
for Industrial and Applied Mathematics.
Sambridge, M., and G. Drijkoningen, 1992, Genetic algorithms in seismic
waveform inversion: Geophysical Journal International, 109, 323–342.
Sambridge, M., and K. Mosegaard, 2002, Monte Carlo methods in geophysi-
cal inverse problems: Reviews of Geophysics, 40, 1–29.
Sambridge, M. S., A. Tarantola, and B. L. N. Kennett, 1991, An alternative
strategy for non linear inversion of seismic waveforms: Geophysical Pros-
pecting, 39, 457–491.
Scales, J. A., P. Docherty, and A. Gersztenkorn, 1990, Regularization of non-
linear inverse problems: Imaging the near-surface weathering layer: In-
verse Problems, 6, 115–131.
Scales, J. A., and M. L. Smith, 1994, Introductory geophysical inverse theo-
ry: Samizdat Press.
Sears, T. J., S. C. Singh, and P. J. Barton, 2008, Elastic full waveform inver-
sion of multi-component OBC seismic data: Geophysical Prospecting, 56,
843–862.
Sen, M. K., and P. L. Stoffa, 1995, Global optimization methods in geophysi-
cal inversion: Elsevier Science Publ. Co., Inc.
Sheng, J., A. Leeds, M. Buddensiek, and G. T. Schuster, 2006, Early arrival
waveform tomography on near-surface refraction data: Geophysics, 71,
no. 4, U47–U57.
Shi, Y., W. Zhao, and H. Cao, 2007, Nonlinear process control of wave-equa-
tion inversion and its application in the detection of gas: Geophysics, 72,
no. 1, R9–R18.
Shin, C., and W. Ha, 2008, Acomparison between the behavior of objective
functions for waveform inversion in the frequency and Laplace domains:
Geophysics, 73, no. 5, VE119–VE133.
Shin, C., and Y. H. Cha, 2008, Waveform inversion in the Laplace domain:
Geophysical Journal International, 173, 922–931.
——–, 2009, Waveform inversion in the Laplace-Fourier domain: Geophysi-
cal Journal International, 177, 1067–1079.
Shin, C., S. Jang, and D. J. Min, 2001a, Improved amplitude preservation for
prestack depth migration by inverse scattering theory: Geophysical Pros-
pecting, 49, 592–606.
Shin, C., and D.-J. Min, 2006, Waveform inversion using a logarithmic
wavefield: Geophysics, 71, no. 3, R31–R42.
Shin, C., S. Pyun, and J. B. Bednar, 2007, Comparison of waveform inver-
sion, Part 1: Conventional wavefield vs. logarithmic wavefield: Geophysi-
cal Prospecting, 55, 449–464.
Shin, C., K. Yoon, K. J. Marfurt, K. Park, D. Yang, H. Y. Lim, S. Chung, and
S. Shin, 2001b, Efficient calculation of a partial derivative wavefield using
reciprocity for seismic imaging and inversion: Geophysics, 66,1856
1863.
Shipp, R. M., and S. C. Singh, 2002, Two-dimensional full wavefield inver-
sion of wide-aperture marine seismic streamer data: Geophysical Journal
International, 151, 325–344.
Sirgue, L., 2003, Inversion de la forme d’onde dans le domaine fréquentiel de
données sismiques grand offset: Ph.D. thesis, Université Paris and
Queen’s University.
——–, 2006, The importance of low frequency and large offset in waveform
inversion: 68th Conference & TechnicalExhibition, EAGE, Extended Ab-
stracts,A037.
Sirgue, L., O. I. Barkved, J. P. V. Gestel, O. J. Askim, and J. H. Kommedal,
2009, 2D waveform inversion on Valhall wide-azimuth OBS: 71st Confer-
ence & TechnicalExhibition, EAGE, Extended Abstracts, U038.
Sirgue, L., J. Etgen, and U. Albertin, 2007, 3D full waveform inversion: Wide
versus narrow azimuth acquisitions: 77th Annual International Meeting,
SEG, Expanded Abstracts, 1760–1764.
——–, 2008, 3D frequency domain waveform inversion using time domain
finite difference methods: 70th Conference & Technical Exhibition,
EAGE, Extended Abstracts, F022.
Sirgue, L., and R. G. Pratt, 2004, Efficient waveform inversion and imaging:
A strategy for selecting temporal frequencies: Geophysics, 69, 231–248.
Smithyman, B. R., R. G. Pratt, J. G. Hayles, and R. J. Wittebolle, 2008, Near
surface void detection using seismic Q-factor waveform tomography:
70th Annual International Meeting, EAGE, ExpandedAbstracts, F018.
Snieder, R., M. Xie, A. Pica, and A. Tarantola, 1989, Retrieving both the im-
pedance contrast and background velocity: Aglobal strategy for the seis-
mic reflection problem: Geophysics, 54, 991–1000.
Song, Z., P. Williamson, and G. Pratt, 1995, Frequency-domain acoustic-
wave modeling and inversion of crosshole data, Part 2: Inversion method,
synthetic experiments and real-data results: Geophysics, 60,786–809.
Sourbier, F.,A. Haiddar, L. Giraud, S. Operto, and J. Virieux, 2008, Frequen-
cy-domain full-waveform modeling using a hybrid direct-iterative solver
based on a parallel domain decomposition method: A tool for 3D full-
waveform inversion?: 78th Annual International Meeting, SEG, Expand-
ed Abstracts, 2147–2151.
Sourbier, F., S. Operto, J. Virieux, P. Amestoy, and J. Y. L’Excellent, 2009a,
FWT2D: Amassively parallel program for frequency-domain full-wave-
form tomography of wide-aperture seismic data Part 1: Algorithm:
Computers and Geosciences, 35, 487–495.
——–, 2009b, FWT2D: Amassively parallel program for frequency-domain
full-waveform tomography of wide-aperture seismic data Part 2: Nu-
merical examples and scalability analysis: Computers & Geosciences, 35,
496–514.
Stekl, I., and R. G. Pratt, 1998, Accurate viscoelastic modeling by frequency-
domain finite difference using rotated operators: Geophysics, 63,
1779–1794.
Stolt, R. H., 1978, Migration by Fourier transform: Geophysics, 43, 23–48.
Taillandier, C., M. Noble, H. Chauris, and H. Calandra, 2009, First-arrival
travel time tomography based on the adjoint state method: Geophysics,
this issue.
Talagrand, O., and P. Courtier, 1987, Variational assimilation of meteorolog-
ical observations with the adjoint vorticity equation, I: Theory: Quarterly
Journal of the Royal Meteorological Society, 113, 1311–1328.
Tarantola, A., 1984, Inversion of seismic reflection data in the acoustic ap-
proximation: Geophysics, 49, 1259–1266.
——–, 1986, A strategy for nonlinear inversion of seismic reflection data:
Geophysics, 51, 1893–1903.
——–, 1987, Inverse problem theory: Methods for data fitting and model pa-
rameter estimation: Elsevier Science Publ. Co., Inc.
Thierry, P., G. Lambaré, P. Podvin, and M. Noble, 1999a, 3-D preserved am-
plitude prestack depth migration on a workstation: Geophysics, 64,
222–229.
Thierry, P., S. Operto, and G. Lambaré, 1999b, Fast 2D ray-Born inversion/
migration in complex media: Geophysics, 64, 162–181.
Thomsen, L. A., 1986, Weakelastic anisotropy: Geophysics, 51, 1954 –1966.
Toksöz, M. N., and D. H. Johnston, 1981, Seismic wave attenuation: SEG.
Tromp, J., C. Tape, and Q. Liu, 2005, Seismic tomography, adjoint methods,
time reversal, and banana-doughnut kernels: Geophysical Journal Interna-
tional, 160, 195–216.
Tsvankin, I., 2001, Seismic signature and analysis of reflection data in aniso-
tropic media: Elsevier Scientific Publ. Co., Inc.
van den Berg, P. M., and A. Abubakar, 2001, Contrast source inversion meth-
Full-waveform inversion WCC151
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
od: State of the art: Progress in Electromagnetics Research, 34, 189–218.
Vigh, D., and E. W. Starr, 2008a, 3D plane-wave full-waveform inversion:
Geophysics, 73, no. 5, VE135–VE144.
——–, 2008b, Comparisons for waveform inversion, time domain or fre-
quency domain?: 78th Annual International Meeting, SEG, ExpandedAb-
stracts, 1890–1894.
Vigh, D., E. W. Starr, and J. Kapoor, 2009, Developing earth model with full
waveform inversion: The Leading Edge, 28, 432–435.
Virieux, J., 1986, P-SV wave propagation in heterogeneous media, velocity-
stress finite difference method: Geophysics, 51, 889–901.
Virieux, J., S. Operto, H. B. H. Ali, R. Brossier, V. Etienne, F. Sourbier, L. Gi-
raud, and A. Haidar, 2009, Seismic wave modeling for seismic imaging:
The Leading Edge, 28, 538–544.
Vogel, C., 2002, Computational methods for inverse problems: Society of In-
dustrial and Applied Mathematics.
Vogel, C. R., and M. E. Oman, 1996, Iterative methods for total variation de-
noising: SIAM Journal on Scientific Computing, 17, 227–238.
Wang,Y., and Y.Rao, 2009, Reflection seismic waveform tomography: Jour-
nal of Geophysical Research, doi: 10.1029/2008JB005916.
Warner,M., I. Stekl, and A. Umpleby, 2007, Full wavefield seismic tomogra-
phy Iterative forward modeling in 3D: 69th Conference & Technical
Exhibition, EAGE, Extended Abstracts, C025.
——–, 2008, 3D wavefield tomography: synthetic and field data examples:
78th Annual International Meeting, SEG, Expanded Abstracts,
3330–3334.
Williamson, P., 1991, Aguide to the limits of resolution imposed by scatter-
ing in ray tomography: Geophysics, 56, 202–207.
Woodhouse, J., and A. Dziewonski, 1984, Mapping the upper mantle:Three
dimensional modelling of earth structure by inversion of seismic wave-
forms: Journal of Geophysical Research, 89, 5953–5986.
Woodward, M. J., 1992, Wave-equation tomography: Geophysics, 57,
15–26.
Woodward, M. J., D. Nichols, O. Zdraveva, P. Whitfield, and T. Johns, 2008,
A decade of tomography: Geophysics, 73, no. 5, VE5–VE11.
Wu, R. S., 2003, Wave propagation, scattering and imaging using dual-do-
main one-way and one-return propagation: Pure and Applied Geophysics,
160, 509–539.
Wu, R. S., and K. Aki, 1985, Scattering characteristics of elastic waves by an
elastic heterogeneity: Geophysics, 50, 582–595.
Wu, R.-S., and M. N. Toksöz, 1987, Diffraction tomography and multisource
holography applied to seismic imaging: Geophysics, 52, 11–25.
Yilmaz, O., 2001, Seismic data analysis: processing, inversion and interpre-
tation of seismic data: SEG.
Zelt, C., and P. J. Barton, 1998, Three-dimensional seismic refraction tomog-
raphy: Acomparison of two methods applied to data from the Faeroe ba-
sin: Journal of Geophysical Research, 103, 7187–7210.
Zelt, C. A., R. G. Pratt, A. J. Brenders, S. Hanson-Hedgecock, and J. A. Hole,
2005, Advancements in long-offset seismic imaging:A blind test of travel-
time and waveform tomography: Eos, Transactions of the American Geo-
physical Union, 86, Abstract S52A-04.
Zhou, B., and S. A. Greenhalgh, 2003, Crosshole seismic inversion with nor-
malized full-waveform amplitude data: Geophysics, 68, 1320–1330.
WCC152 Virieux and Operto
Downloaded 04 Dec 2009 to 193.50.85.151. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/
... The full-waveform inversion (FWI) method was originally introduced in the 1970s. Its basic principle is to iteratively update the Earth model parameters by continuously comparing simulated and observed waveforms (Tarantola, 1986;Virieux & Operto, 2009). In the past decades, the theory and practice of FWI have been a hot research topic in academia and industry. ...
... A. Alkhalifah, 2016). When the difference between the predicted and observed seismic waveforms is larger than a half cycle, the inversion often fails to converge to the true solution, especially with respect to the cycle skipped parts of the data (Virieux & Operto, 2009). Current theoretical and applied research on the FWI method mainly focuses on reducing the nonlinearity and nonuniqueness of the inverse problem to recover wavenumber information of the underground model from low to high (T. ...
... The inverse problem is constrained by the first-order elastic wave equation (Virieux, 1986;Virieux & Operto, 2009), which is given by where Ψ = (v 1 , v 2 , v 3 , σ 1 , σ 2 , σ 3 , σ 4 , σ 5 , σ 6 ) is a vector containing three particle velocities and six stresses, C is the stiffness matrix, which is a 6 × 6 matrix. E denotes space differentiation and its size is determined by the order of finite-difference schemes. ...
Article
Full-text available
Nowadays, the most successful applications of full‐waveform inversion (FWI) involve marine seismic data under acoustic approximations. Elastic FWI of land seismic data is still challenging in theory and practice. Here, we propose a full dispersion spectrum inversion method and apply it to seismic data acquired in West Antarctica. Inspired by the conventional surface wave dispersion curve inversion method, we propose to invert the surface wave dispersion spectrum instead of the complicated waveforms. We compare the frequency‐velocity, frequency‐slowness, and frequency‐wavenumber spectra in terms of their ability to resolve dispersion modes and the feasibility of their adjoint updates and conclude that the frequency‐slowness spectrum is the best for our inversion objectives. We test four objective functions, subtraction, zero‐lag crosscorrelation, optimal transport, and the local‐crosscorrelation to quantify the spectrum mismatch and provide the corresponding adjoint source. We then theoretically analyze the convexity of the proposed objective functions and examine their convergence behavior using numerical examples. We also compare the proposed method with the classic FWI method and the traditional surface wave dispersion curve inversion method and discuss the strengths and weaknesses of each method. This technique is employed to evaluate the shallow velocity structures beneath a seismic array stationed in West Antarctica. Our proposed inversion scheme is also useful for more general applications such as imaging the shallow subsurface of the critical zones, like geothermal reservoirs, and CO2 storage sites.
... At the basis of this effort is Full-Waveform Inversion (FWI), a modern technique that provides high-resolution images of the subsurface by exploiting information in the recorded seismic waveforms [Virieux andOperto, 2009, Chauris, 2019]. This approach requires solving a nonlinear and typically non-unique inverse problem. ...
... So far, two separate approaches are classified: deterministic and stochastic. Deterministic methods aim to reveal a sole optimal solution which is referred to as the maximum a posteriori (MAP), frequently achieved through optimization methods [Virieux and Operto, 2009]. However, these conventional strategies can often get stuck in a local minimum and do not fully address the uncertainties inherent in seismic inversion [Plessix, 2006]. ...
... In the context of seismic imaging, FWI is a common approach to reconstruct subsurface models by comparing observed seismic data with simulated data generated from a forward model [Virieux andOperto, 2009, Chauris, 2019]. This process involves solving the wave equation, which describes the propagation of seismic waves through the subsurface parameters such as velocity. ...
Preprint
Full-text available
To obtain high-resolution images of subsurface structures from seismic data, seismic imaging techniques such as Full Waveform Inversion (FWI) serve as crucial tools. However, FWI involves solving a nonlinear and often non-unique inverse problem, presenting challenges such as local minima trapping and inadequate handling of inherent uncertainties. In addressing these challenges, we propose leveraging deep generative models as the prior distribution of geophysical parameters for stochastic Bayesian inversion. This approach integrates the adjoint state gradient for efficient back-propagation from the numerical solution of partial differential equations. Additionally, we introduce explicit and implicit variational Bayesian inference methods. The explicit method computes variational distribution density using a normalizing flow-based neural network, enabling computation of the Bayesian posterior of parameters. Conversely, the implicit method employs an inference network attached to a pretrained generative model to estimate density, incorporating an entropy estimator. Furthermore, we also experimented with the Stein Variational Gradient Descent (SVGD) method as another variational inference technique, using particles. We compare these variational Bayesian inference methods with conventional Markov chain Monte Carlo (McMC) sampling. Each method is able to quantify uncertainties and to generate seismic data-conditioned realizations of subsurface geophysical parameters. This framework provides insights into subsurface structures while accounting for inherent uncertainties.
... Fullwave inversion techniques, in which the full scattering model (e.g. acoustic or elastic wave equation, Maxwell's equations, etc.) is used in a reconstruction method such as regularised least-squares or Bayesian inversion, have now been widely employed for many inverse scattering problems, in particular for geophysical imaging [47,49] and the related ground-penetrating radar (GPR) problem [25,32,33,48]. For through-wall radar, full-wave inversion approaches have been applied in conjunction with level-set techniques both for 2D [26,28] and 3D image formation [27, ch. ...
Article
Through-wall synthetic aperture radar (SAR) imaging is of significant interest for security purposes, in particular when using multi-static SAR systems consisting of multiple distributed radar transmitters and receivers to improve resolution and the ability to recognise objects. Yet there is a significant challenge in forming focused, useful images due to multiple scattering effects through walls, whereas standard SAR imaging has an inherent single scattering assumption. This may be exacerbated with multi-static collections, since different scattering events will be observed from each angle and the data may not coherently combine well in a naive manner. To overcome this, we propose an image formation method which resolves full-wave effects through an approximately known wall or other arbitrary obstacle, which itself has some unknown ``nuisance'' parameters that are determined as part of the reconstruction to provide well focused images. The method is more flexible and realistic than existing methods which treat a single wall as a flat layered medium, whilst being significantly computationally cheaper than full-wave methods, strongly motivated by practical considerations for through-wall SAR.
... The technique deploys transducer arrays around the region of interest, enabling the inversion of collected ultrasonic guided wave signals and the reconstruction of areas of material debonding [13]. Existing guided wave tomography algorithms primarily include traveltime tomography [14,15], diffraction tomography [16], HARBUT tomography [17,18], and full waveform inversion tomography [19]. ...
Article
Full-text available
This study aimed to address the challenges encountered in traditional bulk wave delamination detection methods characterized by low detection efficiency. Additionally, the limitations of guided wave delamination detection methods were addressed, particularly those utilizing reflected waves, which are susceptible to edge reflections, thus complicating effective defect extraction. Leveraging the full waveform inversion algorithm, an innovative approach was established for detecting delamination defects in multi-layered structures using ultrasonic guided wave arrays. First, finite element modeling was employed to simulate guided wave data acquisition by a circular array within an aluminum–epoxy bilayer structure with embedded delamination defects. Subsequently, the full waveform inversion algorithm was applied to reconstruct both regular and irregular delamination defects. Analysis results indicated the efficacy of the proposed approach in accurately identifying delamination defects of varying shapes. Furthermore, an experimental platform for guided wave delamination defect detection was established, and experiments were conducted on a steel–cement bilayer structure containing an irregular delamination defect. The experimental results validated the exceptional imaging precision of our proposed technique for identifying delamination defects in multi-layered boards. In summary, the proposed method can accurately determine both the positions and sizes of defects with higher detection efficiency than traditional pulse-echo delamination detection methods.
... Full waveform inversion (FWI) is an advanced iterative approach firmly grounded in the principles of wave propagation, as initially formulated by Taratola [1]. Its ability to leverage comprehensive wavefield information within geophysical observations positions it as the state-of-the-art approach for accurate reconstruction of physical parameters in complex subsurface conditions [2]. The successful applications of FWI with field seismic data in the past few decades demonstrate its significant potential in overcoming the limitations of traditional data processing for achieving high-precision imaging [3,4,5,6]. ...
Article
Full-text available
Full waveform inversion (FWI) stands as the fore-front geophysical inversion approach, however, its impediment in practical applications persists due to the absence of prior information and limitations in data acquisition. Addressing this challenge, we introduce implicit FWI (IFWI) which represents subsurface models as continuous and implicit functions, thereby reducing dependency on the initial model with frequency principle in deep learning optimization. Furthermore, through the design of a specialized deep learning model that emphasizes rigorous low-frequency learning, we present a robust IFWI algorithm exhibiting high-resolution reconstruction capabilities, even when low-frequency information is absent in observations, as demonstrated in numerical experiments. Moreover, experimental findings underscore the heightened robustness, reduced data requirements, and strong generalization ability of the proposed robust IFWI algorithm. This highlights its applicability to a variety of subsurface models with diverse acquisition settings, indicating promising potential for practical seismic inversion.
... Mathematically, FWI can be regarded as an optimization problem subject to the constraint of a partial differential equation (PDE). Under the constraint of the acoustic wave equation, sound speed is determined by minimizing the misfit between the observed and PDE-simulated data [17], i.e., ...
Preprint
Ultrasound computed tomography (USCT), as an emerging technology, can provide multiple quantitative parametric images of human tissue, such as sound speed and attenuation images, distinguishing it from conventional B-mode (reflection) ultrasound imaging. Full waveform inversion (FWI) is acknowledged as a technique with the greatest potential for reconstructing high-resolution sound speed images in USCT. However, traditional FWI for sound speed image reconstruction suffers from high sensitivity to the initial model caused by its strong non-convex nonlinearity, resulting in poor performance when ultrasound signals are at high frequencies. This limitation significantly restricts the application of FWI in the USCT imaging field. In this paper, we propose an untrained neural network (UNN) that can be integrated into the traditional iteration-based FWI framework as an implicit regularization prior. This integration allows for seamless deployment as a plug-and-play module within existing FWI algorithms or their variants. Notably, the proposed UNN method can be trained in an unsupervised fashion, a vital aspect in medical imaging where ground truth data is often unavailable. Evaluations of the numerical simulation and phantom experiment of the breast demonstrate that the proposed UNN improves the robustness of image reconstruction, reduces image artifacts, and achieves great image contrast. To the best of our knowledge, this study represents the first attempt to propose an implicit UNN for FWI in reconstructing sound speed images for USCT.
... Nonconvex Optimisation While many applications can be well modelled with a linear forward model, there are also many applications where a linear model is insufficient, e.g. electrical impedance tomography (EIT) [CIN99], phase retrieval for imaging [SEC + 15], photoacoustic tomography [WA11], full waveform inversion [Tar84] with applications in geophysics, [VO09], optical imaging generally [AH97] and diffuse optimal tomography specifically [BBM + 01], to name just a few. From an optimisation perspective this is challenging as the resulting minimisation problem is generally nonconvex, its regularity properties are heterogeneous and most algorithms are only guaranteed to converge to locally optimal solutions. ...
Preprint
Full-text available
Stochastic optimisation algorithms are the de facto standard for machine learning with large amounts of data. Handling only a subset of available data in each optimisation step dramatically reduces the per-iteration computational costs, while still ensuring significant progress towards the solution. Driven by the need to solve large-scale optimisation problems as efficiently as possible, the last decade has witnessed an explosion of research in this area. Leveraging the parallels between machine learning and inverse problems has allowed harnessing the power of this research wave for solving inverse problems. In this survey, we provide a comprehensive account of the state-of-the-art in stochastic optimisation from the viewpoint of inverse problems. We present algorithms with diverse modalities of problem randomisation and discuss the roles of variance reduction, acceleration, higher-order methods, and other algorithmic modifications, and compare theoretical results with practical behaviour. We focus on the potential and the challenges for stochastic optimisation that are unique to inverse imaging problems and are not commonly encountered in machine learning. We conclude the survey with illustrative examples from imaging problems to examine the advantages and disadvantages that this new generation of algorithms bring to the field of inverse problems.
Article
The propagation of seismic waves in viscous media, such as the loess plateau and shallow gas regions, alters their amplitude, frequency, and phase due to absorption attenuation, resulting in reductions in the resolution and fidelity of seismic profiles and the inaccurate identification of subtle structure and lithology. Q modeling and Q migration techniques proposed in this paper are used to compensate for the energy and frequency attenuation of seismic waves, obtain high-quality depth imaging results, and further enhance structural imaging to address the aforementioned problem. First, various prior information is utilized to construct an initial Q model. Q tomography techniques are employed to further optimize the precision of the initial Q model and build a high-precision Q model. Subsequently, Q prestack depth migration technology is employed to compensate for absorption and attenuation in the three-dimensional space along the seismic wave propagation path and correct the travel times, realizing the purposes of amplitude compensation, frequency recovery, and phase correction, which can help improve the wave group characteristics while enhancing the resolution. Model data and practical application results demonstrate that high-precision Q modeling and Q migration techniques can substantially improve the imaging quality of underground structures and formations in the loess plateau region with extremely complex surface and near-surface conditions. The resolution and fidelity of seismic data, as well as the capability to identify reservoirs, can be improved using these techniques.
Article
Full-text available
The need for reconstruction of the distribution of physical properties like dielectric permittivity and electrical conductivity of shallow subsurface sedimentary architecture leads to the development of an optimum strategy of GPR data inversion. In this paper, we present finite difference frequency domain (FDFD) full waveform inversion (FWI) method to get high-resolution subsurface model using GPR data. FWI is an optimization technique which involves in search of the minima between recorded and predicted data. The inversion process includes the quasi-Newton method and simultaneous frequency sampling strategy of irregular sampling. The Hessian term in quasi-Newton algorithm is approximated using preconditioned-LBFGS consideration and the search directions are also optimized after following the Wolfe conditions. At the end of each iteration during inversion, permittivity and conductivity models were updated and became ready to be the initial model for the next iteration. The goals of this research were to develop a robust framework for sedimentary-GPR data inversion and to evaluate the efficacy of the novel grid strategy introduced by Layek and Sengupta (2021) proposed for FWI. This paper presents a comparative analysis between conventional and newly proposed technique from Layek and Sengupta (2021), supported by numerical experiments conducted through our own MATLAB programming. Numerical tests conducted on a benchmark from previously published article, established the fact that new grid formulation produces a faster converging rate and required less computation time. This approach demonstrates remarkable efficacy when applied to a comprehensive sedimentary model comprising a lossy medium.
Chapter
In a followup to our participation in the 2004 BP velocity benchmark (Pratt and Brenders, 2004), we present a refined application of waveform tomography - traveltime tomography followed by frequency-domain waveform inversion - to estimate the (previously unknown) velocity model. Due to the importance of direct (refracted) energy, the limited offset of marine seismic streamers provides a fundamental limit to the depths tomographic techniques can accurately image. In addition, we present the results of waveform tomography using data simulating two-ship experiments, providing total offsets of up to 32 km. We use the 2004 BP velocity benchmark dataset to illustrate the ability of waveform tomography to delineate top-salt and image shallow, low-velocity anomalies.
Article
Traveltime tomography is an appropriate method for estimating seismic velocity structure from arrival times. However, tomography fails to resolve discontinuities in the velocities. Wave-equation techniques provide images using the full wave field that complement the results of traveltime tomography. We use the velocity estimates from tomography as a reference model for a numerical propagation of the time reversed data. These 'backpropagated' wave fields are used to provide images of the discontinuities in the velocity field. The combined use of traveltime tomography and wave-equation imaging is particularly suitable for forming high-resolution geologic images from multiple-source/multiple-receiver data acquired in the borehole-to-borehole seismic surveying. In the context of crosshole imaging, an effective implementation of wave-equation imaging is obtained by transforming the data and the algorithms into the frequency domain. This transformation allows the use of efficient frequency-domain numerical propagation methods. Experiments with computer-generated data demonstrate the quality of the images that can be obtained from only a single frequency component of the data.
Article
Modern wide-angle surveys are often multi-fold and multi-channel, with densely sampled source and receiver spacings. Such closely spaced data are potentially amenable to multi-channel techniques involving wavefield propagation methods, such as those commonly used in reflection data processing. However, the wide-angle configuration requires techniques capable of handling very general wave types, including those not commonly used in reflection seismology. This is a situation analogous to that faced in cross-borehole seismics, where similar wave types are also recorded. In a real cross-borehole example, we compare pre-stack migration, traveltime tomography and wavefield inversion. We find that wavefield inversion produces images that are quantitative in velocity (as are the tomograms) but are of significantly higher resolution; the wavefield inversion results have a resolution comparable to that of the (qualitative) pre-stack migration images. We seek to extend this novel development to the larger-scale problem of crustal imaging. An essential element of the approach we adopt is its formulation entirely within the temporal frequency domain. This has three principal advantages: (1) we can choose to ‘decimate’ the data, by selecting only a limited number of frequency components to invert, thus making inversion of data from large numbers of source positions feasible; (2) we can mitigate the notorious non-linearity of the seismic inverse problem by progressing from low-frequency components in the data to high-frequency components; and (3) we can include in the model any arbitrary frequency dependence of inelastic attenuation factors, Q(ω), and indeed solve for the spatial distribution of Q. An initial synthetic test with an anomaly located within the middle crust yields a velocity image with the correct structural features of the anomaly and the correct magnitude of velocity anomaly. This is related to the fact that the reconstruction is obtained from forward-scattered waves. Under these conditions, the method thus behaves much like tomography. A second test with a deeper, more extensive anomaly yields an image with the correct velocity polarity and the correct location, but with a deficiency in low and high wavenumbers. In this case, this is because the reconstruction is obtained from backscattered waves; under these conditions the method behaves not like tomography, but like migration. A more extensive test, based on a large wide-angle survey in south-eastern California and western Arizona, demonstrates a real potential for high-resolution imaging of crustal structures. Although our results are limited by the acoustic approximation and by the relatively low frequencies that we can model today, the images are sufficiently encouraging to warrant future research. The problem of local minima in the objective function is the most significant practical problem with our method, but we propose that appropriate ‘layer’ stripping methods can handle this problem.
Article
The aim of this paper is to present recent developments in the theory and practice of parameter estimation problems in distributed parameter systems. We first present two theoretical developments for the classical output least square method : the compact regularization method, and the curvature method for the investigation of non linear least square problems. Then we discuss the use of continuation or homotopy method for overcoming the problem of local minima in large size problems. Finally, a last paragraph, still missing in the present version of the paper, was intended for the presentation of competitors to the output least square method, as the Variational Method and the Augmented Lagrangian Method.
Article
The success of this book stems from its clear and concise, yet detailed summary of the advances in seismic source studies during the past two decades. Dr Kennett presents a mainly theoretical account of the passage of seismic waves from source to receiver, linking the theoretical development to the nature of seismograms observed across a wide range of distance scales - from a few kilometres, as in shallow reflection work for geophysical prospecting, to many thousands of kilometres for earthquakes. A unified framework is presented for all classes of seismic phenomena, for both body waves and surface waves. Each topic is taken up systematically, including many topics not normally covered in discussion of propagator theory, such as source representation theory, generalised ray theory, and the calculation of complete theoretical seismograms including all wave effects arising from the presence of the Earth's surface.