ADVANCES IN GEOPHYSICS, VOL. 50, CHAPTER 1
COHERENT BACK-SCATTERING AND WEAK
LOCALIZATION OF SEISMIC WAVES
Ludovic Margerin
Abstract
I present a review of the weak localization effect in seismology. To understand this multiple
scattering phenomenon, I begin with an intuitive approach illustrated by experiments performed
in the laboratory. The importance of reciprocity and interference in scattering media is emphasized.
I then consider the role of source mechanism, again starting with experimental evidence. Important
theoretical results for elastic waves are summarized, that take into account the full vectorial
character of elastic waves. Applications to the characterization of heterogeneous elastic media
are discussed.
Key Words:
Multiple scattering, interference, reciprocity, elastic waves.
! 2008 Elsevier Inc.
1. Introduction
In strongly scattering media, the propagation of multiply-scattered waves is best
described by considering the transport of the energy. An elastic scattering medium is
an inhomogeneous medium where the wavespeed and the density vary laterally. It can
also contain embedded obstacles such as cracks or cavities. Upon propagation, an
incoming plane wave with well-defined wavevector k will transfer energy to all possible
space directions, a phenomenon known as scattering. The energy transport approach has
been developed by astrophysicists at the beginning of the twentieth century and has given
birth to the theory of radiative transfer (Chandrasekhar, 1960; Apresyan and Kravtov,
1996). Phenomenologically, the transfer equation for acoustic, electromagnetic, and
elastic waves can be derived from a detailed local balance of energy that neglects the
possible interference between wave packets (see, e.g., Sato, 1994; Sato and Fehler, 1998;
and Margerin, 2005, for seismic applications). This important assumption is justified by
the fact that the phase of the wave is randomized by the scattering events. Thus at a given
point, the field can be written as a sum of random phasors and on average, intensities can
be added, rather than amplitudes. This seemingly convincing argument can actually be
shown to be wrong and one of the goals of this paper will be to put forward the role of
interferences in scattering media in specific cases. To begin with, we present the results
of an experiment of ultrasound propagation in a granular material, which can be defined
as a material containing many individual solid particles with arbitrary sizes. An array of
128 acoustic transducers has been placed at the surface of a box with lateral dimensions
0.15 m ! 0.15 m ! 0.15 m containing commercial sand for aquariums. The sand does not
have a well-defined granulometry but the typical size of a grain is "2 mm. The central
transducer emits a short pulse in the 1#1.5 MHz frequency range and the waves are recorded
1
#
2008 Elsevier Inc. All rights reserved.
ISSN: 0065-2687
DOI: 10.1016/S0065-2687(08)00001-0
2
MARGERIN
by the whole array. The wavespeed of the dominant ballistic pulse is about 1000 m/s,
which gives a dominant wavelength of "0.8 mm, which is of the same order as
the size of one transducer of the array ("0.55 mm). The logarithm of the energy of
the wavefield along the array after time averaging over four cycles is shown in
Fig. 1, as a function of distance from the center of the array in millimeters
(horizontal axis) and time in microseconds (vertical axis).
In this figure, one can identify direct waves that propagate along the array and decay
exponentially with distance because of the energy losses due to scattering. They are
followed by a diffuse coda, which can be thought of as the result of the random walk of
the energy p
inffiffiffiffithe
ffi scattering medium. As a rule of thumb, the multiple scattering halo
grows like Dt, where D = ul*/3 is the diffusion constant of the waves in the medium and
u is the wave velocity. The transport mean free path l* (see, e.g., Sheng, 1995, for a
rigorous definition) represents the typical step length of the random walk of the energy in
the scattering medium and is much larger than the wavelength. In the case of sand
samples, the transport mean free path is roughly 10 times larger than the wavelength.
According to diffusion theory (Akkermans and Montambaux, 2005), at fixed time t = t0,
the energy distribution in the scattering medium is approximately proportional to
2
%
e#3r =ð4ut0 l Þ , where r is the distance from the source. Therefore, at fixed time, the energy
in the diffuse halo is expected to vary significantly on the scale of l*. Yet, at the center of
Fig. 1, the reader will notice a clear, but highly localized increase of intensity. This is not
20
−4
30
−5
40
Time (microsec)
Diffuse
halo
−3
−6
50
−7
60
70
−8
80
Weak
localization 90
−9
−10
100
−11
Log (energy)
Ballistic
waves
10
110
−12
−60
−40
−20
0
20
Distance (mm)
40
60
FIG. 1. Energy of the wavefield recorded at the surface of a granular material as a function of
time. A short pulse with a central frequency of 1.2 MHz is shot at the central transducer. Direct
waves propagating along the array rapidly attenuate. They are followed by coda waves that form a
diffuse halo in the medium. Note the sharp increase of intensity at the center of the array, where
energy was initially released. Experiment performed at the Laboratoire Ondes et Acoustique, Paris,
by R. Hennino and A. Derode.
COHERENT BACK-SCATTERING AND WEAK LOCALIZATION
3
an artifact. In this particular experiment, the typical width of the zone of enhanced
intensity is roughly equal to the size of one transducer. Other experiments, to be described
later, have demonstrated that the zone of enhancement actually coincides with the central
wavelength of the waves, which is the clear signature of an interference effect that takes
place around the source in the multiple scattering medium: this is known as weak
localization. In the next section, I provide a simple explanation for this observation.
2. Weak Localization Effect: A Heuristic View
In what follows, I represent a scalar partial wave as a complex number c = Aeif, where
A and f are real numbers denoting the amplitude and phase, respectively. Each partial
wave follows an arbitrarily complicated scattering path from source to receiver in the
medium. At a given point, the measured field u is a superposition of a large number of
partial waves that have propagated along different scattering paths
u¼
X
Aj eifj ;
ð1Þ
j
where Aj and fj are random and uncorrelated because of the multiple scattering events,
and j can be understood as a “label” for the different paths. The representation (1) is
strictly valid for point scatterers and will suffice for the present purposes. Typical
examples of scattering paths are shown in Fig. 2.
In Eq. (1), I now pair direct and reciprocal scattering paths as shown in Fig. 2. The
direct and reciprocal paths are characterized by the fact that the same scatterers are
visited, but the sequences of scattering events are opposite. To illustrate this definition, in
Fig. 2, the sequence S, A, B, C, D, R (solid line) and S, D, C, B, A, R (dashed line)
represent the direct and reciprocal paths, respectively. One obtains
u¼
X"
j0
#
cdj0 þ crj0 ;
ð2Þ
where c denotes the complex partial waves, the superscripts d and r stand for “direct”
and “reciprocal,” and a new label j0 has been introduced to emphasize the new representation of the field. The intensity I is proportional to |u|2 and reads
I¼
X"
j0 ;k0
cdj0 þ crj0
#$
%
cdk0 þ crk0 ;
ð3Þ
where the overbar denotes complex conjugation. In Eq. (3), it is reasonable to assume
that the waves visiting different scatterers will have random phase differences and after
averaging over scatterer positions will have no contribution. Thus, we can restrict the
summation to the case j0 ¼ k0 to obtain
I¼
X'&& &&2 && &&2 ( X
ðcdj0 crj0 þ cdj0 crj0 Þ:
&cdj0 & þ &crj0 & þ
j0
j0
ð4Þ
4
MARGERIN
Configuration 1
S
S R
R
D
D
A
A
C
C
B
B
Configuration 2
S R
R
S
C
C
D
D
B
B
A
A
FIG. 2. Examples of multiple scattering paths from source S to receiver R. Scattering events are
labeled with letters A, B, C, and D. Solid and dashed lines represent direct and reciprocal paths,
respectively. The two configurations differ by the position of the scattering events. On the left:
source and receiver coincide. On the right, source and receiver are typically a few wavelengths
apart. Reprinted figure with permission from Margerin et al. (2001). Blackwell Publishing.
The first sum in Eq. (4) represents the usual incoherent contribution to the measured
intensity, which can be calculated with radiative transfer theory (see Wegler et al., 2006
for recent applications). The second sum can be interpreted as the interference term
between the direct and reciprocal paths in the scattering medium. In a reciprocal medium,
that is, a medium where the reciprocity principle is verified, the amplitude and phase of
the direct and reciprocal wave paths are exactly the same, that is, Adj ¼ Arj and fdj ¼ frj ,
provided that source and receiver are located at the same place. Therefore, the total
intensity which includes the interference term is exactly double of the classical incoherent term. This is the interference term which causes the intensity to be higher in the
experiment shown in Fig. 1. Reciprocity is a general property of wave equations such as
the acoustic and elastic wave equation. In a simple scalar case, it means that the response
measured at r2 due to a source at r1 is the same as the response measured at r1 due to a
source at r2. This remarkable property can be broken when an external field acts on the
system. An interesting seismic example of broken reciprocity is the effect of the Coriolis
force on the seismic wave motion at long period where the effect of the rotation of the
Earth is important. A generalized reciprocity relation can still be given upon exchange of
COHERENT BACK-SCATTERING AND WEAK LOCALIZATION
5
source and receiver, which involves the inversion of the instantaneous rotation vector of
the Earth (see Dahlen and Tromp, 1998, for a thorough discussion).
Although it had been theoretically predicted in several pioneering papers published
around 1970 (Watson, 1969; de Wolf, 1971; Barabanenkov, 1973), it is only in the mideighties that the role of interference in multiple scattering has been appreciated with the
discovery of the coherent backscattering of light (Kuga and Ishimaru, 1984; van Albada
and Lagendijk, 1985; Wolf and Maret, 1985; Kaveh et al., 1986). Later the coherent
backscattering effect has been predicted and observed for acoustic and elastic waves in
both stationary and dynamic experiments (Bayer and Niederdrank, 1993; Sakai et al.,
1997; Tourin et al., 1997; de Rosny et al., 2000). Today, coherent backscattering or weak
localization is still a very active topic of research. The coherent backscattering for
moving scatterers has been studied by Snieder (2006) and Lesaffre et al. (2006).
Derode et al. (2005) have used the coherent backscattering effect to measure the
heterogeneity of human bones. Aubry and Derode (2007) have devised an ingenious
technique to measure lateral variations of the diffusion constant of strongly scattering
media based upon the separation of the incoherent and coherent intensities.
Note that the term “coherent backscattering” refers to the intensity enhancement
observed in a small cone of direction in the far-field of a disordered sample for plane
^ Although the basic physics of coherent
wave sources with fixed incident direction k.
backscattering and weak localization are identical, the latter term indicates that the loops
of interference occur inside the disordered sample, and should therefore be preferred to
describe the seismic experiments. These interference loops result in a deviation from the
diffusive behavior (Haney and Snieder, 2003). When the wavelength is of the same order
as the mean free path, the interference effects can completely block the transport of
energy away from the source, a phenomenon known as strong localization (see, e.g.,
Sheng, 1995; Akkermans and Montambaux, 2005). Weak localization is therefore a basic
phenomenon to explain the transition from the diffuse to the localized propagation
regime. Note that there exists a number of other mechanisms of intensity enhancement.
One of the most famous is the “opposition effect” in astrophysics, which manifests itself
as an increase of the reflectance of celestial bodies such as the Moon when the light of the
Sun reflected from the regolith is observed close to the backscattering direction. Hapke
et al. (1993) have shown that the opposition effect is partly explained by the coherent
backscattering of light. I refer the interested reader to the paper by Barabanenkov et al.
(1991) for an extensive review of backscattering enhancement phenomena in optics. In
particular, these authors discuss the case of backscattering enhancement by several
deterministic scatterers which can also be of interest in seismology. Let me finally
point out that weak localization is only one manifestation of the role of the phase in
the seismic coda (see Campillo, 2006, for a review).
I have shown with a very simple argument that interference effects have to be
incorporated in the usual transport theory, but for the moment, it has not been explained
why the enhancement due to interference is so highly localized. In Fig. 2, I schematically
represent the more usual case where source and detection are not collocated. In that case
there is a phase shift between the two wavepaths which is acquired during the propagation from the source to the first scattering event and from the last scattering event to the
receiver. Clearly, if the distance between source and detection is “large enough,” the
phase shift will fluctuate randomly from one configuration of the scatterers to the other.
Therefore, the interference term is expected to vanish upon averaging when source and
receiver are sufficiently far apart. One of the goals of the paper is to demonstrate and
6
MARGERIN
illustrate the fact that the enhancement zone is actually narrow and about the size of the
wavelength. A final comment on the role of multiple scattering is in order. It is clear that
the representation of the wavefield [Eq. (3)] makes sense only if one can pair the direct
and reciprocal propagation paths. If there is a single scattering event, there is only one
possible path from source to receiver and therefore no interference is possible. This basic
observation proves that weak localization is indeed a genuine multiple scattering effect.
In what follows, I pursue the experimental approach of weak localization by considering
the role of the source mechanism.
3. The Role of Source Mechanism and Wavefield Polarization
Because of their vectorial nature, the weak localization of elastic waves cannot be fully
explained by the simple intuitive approach presented above. As will be shown shortly,
the reciprocity principle of elastic waves in its full extent has to be obeyed in order to
preserve the factor of 2 enhancement at the source position. This subtle effect is first
examined through a laboratory experiment with ultrasound.
3.1. Effect of Source Mechanism
The most common seismic sources, that is, explosions and earthquakes are combinations of dipoles and/or couples of forces. We must therefore consider more complex
sources than simple isotropic point sources. In the case of earthquakes, the radiation is
strongly anisotropic and the radiation pattern displays nodal planes with reversal of the
polarity of first motions. de Rosny et al. (2001) have studied the weak localization of
elastic waves propagating in a chaotic reverberant cavity. The nature of the disorder is
different from the scattering medium, but until a time known as the Heisenberg time, the
mechanisms of enhancement are similar and can be based on an intuitive ray description.
Beyond the Heisenberg time, the eigenmodes of the system can be resolved and the statistical
properties of the eigenfunctions lead to an enhancement of intensity by a factor 3
around the source, as demonstrated experimentally and theoretically by Weaver
and Burkhardt (1994) and Weaver and Lobkis (2000). This result is valid in chaotic
cavities only. Using an interferometer, de Rosny et al. (2001) have recorded the
vertical motions of Lamb waves generated by vertical monopole and dipole sources
in a thin (0.5 mm thickness) chaotic plate of total area 2335 mm2, with the shape of
a quarter stadium. The dominant frequency of the signal is 1.0 MHz and the typical
wavelength is "2.5 mm. After time averaging between lapse time t = 200 ms and
t = 500 ms, they have measured the intensity patterns shown in Fig. 3. The
beginning of the signal is excluded in order to avoid the first reflection on the
boundary of the plate and the choice of the end of the time window is dictated by
the signal to noise ratio. The distribution of energy is perfectly homogeneous in the
plate, except in a small area centered around the source where, in the case of a
monopole, it is the double of the background intensity. An important result of this
study is the confirmation of the typical wavelength size of the zone of intensity
enhancement, which shows that weak localization is a near-field effect. In the 3-D
case, the increase of intensity would occur inside a sphere centered at the source.
1.5
1
7
2
2
Normalized intensity
Normalized intensity
COHERENT BACK-SCATTERING AND WEAK LOCALIZATION
1
0.5
0
-5
0
y(mm)
5 -5
0
x(mm)
1.5
1
0.5
0
5
2
1
y(mm)
0
-1
-2
-1
-2
1
0
x(mm)
2
FIG. 3. Enhanced backscattering of elastic waves in a 2-D chaotic cavity. The central frequency is
1 MHz and the dominant wavelength is 2.5 mm. The integrated intensity between lapse time
t = 200 ms and t = 5 ms is represented as a function of position around the source. Left: Monopolar
source. Right: Dipolar source. In the dipolar case, the enhancement disappears on the line going
through the source and perpendicular to the dipole axis. Along the dipole, the intensity enhancement
presents two maxima located about half a wavelength away from the source. Reprinted figure with
permission from de Rosny et al. (2001). Copyright (2001) by the American Physical Society.
In addition, Fig. 3 (right) highlights the importance of the source mechanism. In the
dipolar case, there are actually two zones of enhancement with an enhancement factor
of about 1.6, separated by a line of zero enhancement. To understand this puzzling
observation, one must consider the full reciprocity principle for elastic waves. In the
present experiment, there is a lack of symmetry between the dipolar source and the
monopolar detection. To restore reciprocity one would like to measure not the field
itself, but its directional derivative along the dipole axis. When this operation is
performed and intensity is redefined as the square of the partial derivative, de
Rosny et al. (2001) have demonstrated both experimentally and theoretically that the
factor of 2 enhancement at the source is restored. Being motivated by this result,
I now give an asymptotic but rigorous theory of weak localization for vector waves.
3.2. Review of Multiple Scattering Formalism
To obtain a satisfactory theory of weak localization, one needs to develop a transport
theory of the energy that keeps track of all polarization indices at both source and
receiver. As shown by Weaver (1990), the necessary information is contained in the
fourth rank coherence tensor Gij! kl of the elastic wavefield defined as
)
*
Gij!kl ðt; r1 ; r2 ! r3 ; r4 Þ ¼ Gaki ðt; r3 ; r1 ÞGljaðt; r4 ; r2 Þ ;
ð5Þ
where Gaki ðt; r3 ; r1 Þ is the element of the Green matrix corresponding to a point force
applied at r1 in direction i, and measured displacements in direction k at r3. The
superscript a is introduced to label the realizations of the random medium. To each a
there corresponds exactly one medium of the statistical ensemble. t denotes the time
elapsed since energy has been released by sources with a common origin time. Note that
in the analysis that follows, the signal is assumed to be band-pass filtered in a narrow
frequency band with central angular frequency o. In order to simplify notation, all tensor
8
MARGERIN
quantities are assumed to depend implicitly on o. The tensor Gij!kl (t, r1, r2!r3, r4)
describes the transfer of the displacement correlation function from source (displacement
indices i, j and positions r1, r2) to receiver (displacement indices k, l and positions r3, r4).
The brackets denote an average over a, that is, over an ensemble of random media with
prescribed statistics. In what follows we assume that the property of statistical homogeneity holds, in which case the ensemble-averaged Green tensor depends on the difference
of the position vectors of the source and detector only
hGaki ðt; r3 ; r1 Þi ¼ Gki ðt; r3 # r1 Þ;
ð6Þ
where for notational simplicity Gki (without superscript) denotes the ensemble averaged
Green tensor.
The complete evolution of the tensor Gij!kl is described by the Bethe-Salpeter
equation which contains all correlations among all possible scattering paths in the
medium (Sheng, 1995). This is far too detailed for the present purposes, and one usually
contents oneself with the approximate calculation of two terms: the classical contribution—also termed “diffuson”—and the interference term between reciprocal paths—also
termed “cooperon”—(see Akkermans and Montambaux, 2005, for the origin of this
terminology). In the radiative transfer equation, information on source is usually
integrated out and the cooperon term is neglected (see Margerin, 2005, for details).
Apresyan and Kravtov (1996) have suggested a modification of the radiative transfer
equation, which includes contribution of the cooperon and thereby is able to describe
coherent phenomena such as weak localization. The cooperon and diffuson contributions
are conveniently represented by Feynman diagrams which are both computationally
efficient and physically meaningful. Typical diagrams are shown in Fig. 4. In the ladder
diagrams, the Green function (upper line) and its complex conjugate (lower line) visit the
same scatterers in the same order. In the crossed diagrams, first introduced for multiple
scattering of classical waves in the pioneering papers of Barabanenkov (1973, 1975), the
upper line is unchanged but in the lower line the sequence of scattering is reversed. The
ladder and crossed diagrams correspond to the classical (incoherent) and interference
(coherent) contributions, respectively. To make the link with the elementary treatment
given in Section I, the reader can think of the ladder diagrams alone, as the result of
summing the intensities of the direct path (solid line in Fig. 2) and reciprocal path
(dashed line in Fig. 2). The ladder and crossed diagrams altogether are the result of
first summing and then squaring the fields of the direct and reciprocal paths.
Below, I give a long-time asymptotic formula for the ladder term. To calculate the
crossed term, one can make use of the following reciprocity argument: the field produced
by a force in direction k at r2 and recorded in direction l at r4 after scattering at A, B,
C, D, . . . is equal to the field produced by a force in direction l at r4 and recorded in
direction k at r2 after scattering at . . ., D, C, B, A. This is equivalent to saying that every
crossed diagram can be turned into a ladder diagram after suitable exchange of the
polarization indices, and positions on the lower line. We now decompose the tensor
Gij!kl into the fundamental diffuson and cooperon contributions and write
Gij!kl ðt; r1 ; r2 ! r3 ; r4 Þ ¼ Lij!kl ðt; r1 ; r2 ! r3 ; r4 Þ þ Cij!kl ðt; r1 ; r2 ! r3 ; r4 Þ:
ð7Þ
The previous discussion implies the following fundamental reciprocity relation between
Cij!kl and Lij!kl
COHERENT BACK-SCATTERING AND WEAK LOCALIZATION
rs + x/2
i
Source
j
rD + y/2
x
x
x
x
x
x
x
x
k
Detector
l
rs - x/2
rD - y/2
rs + x/2
rD + y/2
i
Source
j
x
x
x
x
x
x
x
x
k
Detector
l
rs - x/2
rD - y/2
p + q/2
p! + q/2
i
Source
j
p - q/2
9
x
x
x
x
x
x
x
x
L
C
k
Detector Lpp!(q)
l
p! - q/2
FIG. 4. Typical Feynman diagrams representing the classical (top) and interference (middle)
contribution to the intensity pattern. Crosses connected by dashed lines represent the same scatterer.
The crosses are also connected by solid lines which represent Green functions describing the
propagation between different scatterers. Note that in the lower line, the Green functions are
complex conjugated. The upper diagram L, often pictorially termed “ladder diagram,” gives the
classical contribution to the measured intensity. The middle diagram termed “crossed diagram”
represents the interference term between reciprocal wavepaths. Bottom: scattering diagram in the
wavenumber domain. Reprinted figure with permission from van Tiggelen et al. (2001). Copyright
(2001) by the American Institute of Physics.
Cij!kl ðt; r1 ; r2 ! r3 ; r4 Þ ¼ Lil!kj ðt; r1 ; r4 ! r3 ; r2 Þ:
ð8Þ
As above, I draw the attention of the reader to the fact that relation (8) is true in multiple
scattering only. Fortunately, in nonabsorbing media, the single scattering contribution
vanishes exponentially (see, e.g., Sato and Fehler, 1998) while, at large lapse time, the
ladder contribution can be shown to be the solution of a diffusion equation
(Barabanenkov and Ozrin, 1991, 1995; Sheng, 1995; Akkermans and Montambaux,
2005) and therefore decays only algebraically. In the presence of absorption, both the
single scattering and ladder contribution exhibit an algebro-exponential decay but the
single scattering term still vanishes faster because it suffers from scattering losses. These
facts have been illustrated in papers by Gusev and Abubakirov (1987), Hoshiba (1991),
and Zeng et al. (1991), where solutions of the single scattering and full multiple
scattering problem are presented. Therefore, relation (8) applies after one mean free
time, which is the average time between two scattering events. In addition, from Eq. (8),
one can easily infer that when source and detection coincide r1 ¼ r2 ¼ r3 ¼ r4, and the
polarization of source and detection are identical i ¼ j ¼ k ¼ l, the interference term
equals exactly the diffuson term. Our next task is to provide a general formula to predict
the exact shape of the weak localization effect in the long time limit. This requires an
asymptotic solution of the Bethe-Salpeter equation, which is presented in what follows.
10
MARGERIN
3.3. Theoretical Results for Acoustic and Elastic Waves
The Bethe-Salpeter equation is an exact equation for the full coherence tensor of the
wavefield. The complete solution of this equation seems out of reach in general, but
asymptotic solutions for the diffuson contribution in the long time limit have been
presented by Barabanenkov and Ozrin (1991, 1995), which can be applied to classical
scalar and vectorial linear waves, independent of the underlying equation of motion. In
the theoretical approach, one considers a narrowly band-passed signal with central period
T which is much smaller than the typical duration of the coda. This is known in the
literature as the slowly varying envelope approximation (Sheng, 1995). The results of
Barabanenkov and Ozrin are valid in this limit and I refer the interested reader to the
original publications for further technical details. Asymptotically t ! 1, the ladder or
diffuson or classical contribution takes the form
Lij!kl ðt; r1 ; r2 ! r3 ; r4 Þ ¼
e#t=ta
ðDtÞ3=2
ImGij ðr1 # r2 ÞImGlk ðr4 # r3 Þ;
ð9Þ
where Im denotes the imaginary part of a complex number, D is the diffusion constant of
the waves in the random medium, and ta denotes a phenomenological absorption term.
Note that in the last equation, the tensor Gij stands for the spatial part of the ensembleaveraged elastic Green tensor at angular frequency o ¼ 2p/T.
Because of the underlying assumption of statistical homogeneity, the tensor Gij
depends on the separation vector between source and station only and is an implicit
function of o. The tensor Gij!kl depends on both central frequency of the waves, and
lapse time in the coda. Equation (9) is valid in the slowly varying envelope approximation. The reader is referred to Sheng (1995) and Apresyan and Kravtov (1996) for more
mathematical details. According to Eq. (8), the cooperon term reads
Cij!kl ðt; r1 ; r2 ! r3 ; r4 Þ ¼
e#t=ta
ðDtÞ3=2
ImGil ðr1 # r4 ÞImGjk ðr2 # r3 Þ:
ð10Þ
Let us investigate some consequences of Eq. (10) for a given source station configuration: r1 ¼ r2, r3 ¼ r4. To illustrate the fact that interference effects are important only in a
region of the size of one wavelength around the source, we consider the scalar case and
introduce the enhancement factor defined as the total intensity normalized by the
diffuson contribution: E ¼ 1 þ C/L. For scalar waves in 3-D, the enhancement profile
E is given by
E¼1þ
'
(
sinðkr Þ 2
;
kr
ð11Þ
where k ¼ o/c is the central wavenumber of the scalar waves with velocity c and r is the
source receiver distance. For scalar waves in 2-D, the corresponding formula is
E ¼ 1 þ J0 ðkr Þ2
ð12Þ
COHERENT BACK-SCATTERING AND WEAK LOCALIZATION
11
where J0 denotes the Bessel function of the first kind of order 0. These results are valid in
the long lapse time limit and when the mean free path is much larger than the wavelength,
which is the most common situation in practice. When scattering is extremely strong, the
wavelength can be of the order of the mean free path, and strong localization can set in, a
regime which is very difficult to reach (Akkermans and Montambaux, 2005). The
cardinal sine and Bessel function in Eqs. (11) and (12) are proportional to the imaginary
part of the Green function of the Helmoltz equation in 2-D and 3-D, respectively.
Equation (11) has been verified by numerical simulations (Margerin et al., 2001),
while Eq. (12) has been checked experimentally by de Rosny et al. (2000) (see also
the left part of Fig. 3). It is important to note that the enhancement profile is independent
from absorption. This is not surprising since absorption does not affect the reciprocity
principle. This is often illustrated by the following sentence (van Tiggelen and Maynard,
1997): “If you can see me, I can see you!” that applies even in the fog. This independence
of the weak localization effect on absorption is used later in the paper to give estimates of
the scattering properties of heterogeneous media.
I now explore theoretically the role of the source mechanism. As an example, I will
give a simple formula that explains all the features of the experiments described in Fig. 3
with a dipole source. The radiation of the dipole is obtained by taking the directional
partial derivative of the Green function with respect to source coordinates. For a dipole
source along the x axis of the coordinate system, the coherence function G of the scalar
field is given by
Gðt; r1 ; r2 ! r3 ; r4 Þ ¼ h@x1 Ga ðt; r3 ; r1 Þ@x2 Gaðt; r4 ; r2 Þi:
ð13Þ
Because the operation of taking derivatives is linear, the @x1 and @x2 symbols can be taken
outside of the ensemble average brackets. For the diffuson and cooperon contributions,
one obtains respectively
L¼
C¼
e#t=ta
ðDtÞ3=2
e#t=ta
ðDtÞ3=2
ImGð0Þ@x1 @x2 ImGðr2 # r1 Þjr2 ¼r1
ð14Þ
@x1 ImGðr1 # r4 Þjr4 ¼r3 @x2 ImGðr3 # r2 Þjr2 ¼r1 :
ð15Þ
In the 2-D case, after some algebra, one obtains the following enhancement pattern for
the dipole source
E ¼ 1 þ ðJ1 ðkr ÞcosðyÞÞ2 ;
ð16Þ
where J1 denotes the Bessel function of the first kind of order 1. This theoretical
prediction matches the observations of Fig. 3 (left) very closely.
The calculation of the weak localization effect has thus been illustrated in simple
cases. The calculations in the full elastic case with arbitrary moment tensor sources have
been performed by van Tiggelen et al. (2001). The principle is the same as above but the
calculations are much more tedious. To illustrate the effect of broken symmetry between
source and receiver, I show in Fig. 5 the calculation of the weak localization profile for
explosion and dislocation sources in an infinite elastic medium. As usual, an explosion
Explosion and dislocation: Potential energy cone
(b)
12
(a)
Explosion: Kinetic cone
2.0
1.015
Enhancement
Enhancement
1.8
1.6
1.4
1.010
1.005
1.2
−2
(c)
0
Distance [lambda S]
1.000
2
Dislocation: Kinetic cone between seismic axes
−2
(d)
0
Distance [lambda S]
2
Dislocation: Kinetic cone along seismic axes
1.05
1.08
Enhancement
Enhancement
1.04
1.06
1.04
1.02
1.03
1.02
1.01
1.00
−2
0
Distance [lambda S]
2
1.00
−2
0
Distance [lambda S]
2
MARGERIN
1.0
COHERENT BACK-SCATTERING AND WEAK LOCALIZATION
13
and a seismic dislocation will be represented by three mutually orthogonal dipole of
forces, and by a double couple of forces, both with zero net applied linear and angular
momentum, respectively. The enhancement profiles in Fig. 5 are calculated in the plane
perpendicular to the axis of the double couple. In this plane, there exist two perpendicular
directions, termed seismic axes, with respect to which the moment tensor is diagonal. For
the simple dislocation source, the seismic axes define the direction of maximum radiation of P waves and make an angle 45) with the applied forces. Note that the direction of
application of the forces also coincide with the maximum radiation of S waves. Figure 5
demonstrates that the backscattering enhancement can be totally destroyed if the operations carried out at the source and at the detection are different. For instance, the
enhancement of kinetic energy due to a dislocation is typically less than 10% and
vanishes exactly at the source position. The angular dependence of the enhancement
profile in the plane containing the double couple is illustrated in Fig. 5c and d. It is
seen that the enhancement is slightly higher along the direction of maximum radiation of
S waves. Although I shall not prove it, it is always possible to recover the factor 2
enhancement by measuring the appropriate quantity. For the case of a dipole source, one
needs to measure the derivative of the field along the dipole axis. This of course
complicates the experimental setup but it can be done in practice (see, e.g., Hennino
et al., 2001). For an explosion, one needs to evaluate the divergence of the wavefield
(i.e., the compressional energy of the waves), which is even more demanding. Although
this seems discouraging, I present below some experiments with seismic waves that
demonstrate the potential usefulness of weak localization.
4. Geophysical Applications
At this point, I would like to convince the reader that the weak localization effect is not
only a theoretical or mathematical curiosity but a useful interference effect which can be
used to probe the medium properties when multiple scattering hampers the detection of
ballistic waves. I give two examples of applications of weak localization: the first one
illustrates the measurement of elastic properties of a concrete slab (Larose et al., 2006); the
second one reports the measurement of the mean free path in a volcano (Larose et al., 2004).
4.1. Measurement of the Dispersion Relation of Surface Waves
Let me first consider an application at a scale which is intermediate between the
laboratory and the field. An array of vertical accelerometers with typical bandwidth 0.1–
5 kHz has been set up on a steel reinforced concrete slab. Upon propagation, waves
undergo strong multiple scattering and reflections. In the time domain, the ballistic
signals are difficult to separate from the multiple reflections on the sides of the slab and
FIG. 5 Enhancement profiles as a function of the distance from the source, in units of the shear
wavelength, for an infinite Poissonian medium. (a) The enhancement of the potential energy for an
explosion (solid) and a dislocation (dashed). (b) Enhancement profile for the kinetic energy near an
explosion. (c) and (d) Enhancement profiles for the kinetic energy between (at an angle 45) to) and
along the seismic axes of a dislocation source. The term “cone” shown on top of each plot refers to
the zone of intensity enhancement around the source. Reprinted figure with permission from van
Tiggelen et al. (2001). Copyright (2001) by the American Institute of Physics.
14
MARGERIN
are also rapidly attenuating. In such a situation, an alternative to classical signal processing
techniques is welcome. The elastic waves are generated by an approximately vertical
hammer strike, thus ensuring that source and detection verify the general reciprocity
condition previously discussed. In the frequency range of interest (from 500 to 1500 Hz),
the elastic energy is transported through the slab by the fundamental antisymmetric Lamb
mode, which is known to be strongly dispersive. The phase speed of the fundamental Lamb
mode at 1 kHz is "1000 m/s. The geometry of the experiment can thus be considered quasi
2-D and the shape of the weak localization can be accurately predicted using the theoretical
expression (12). I refer the reader to Larose et al. (2006) for further details.
To illustrate the practical measurement of the weak localization effect, I show in Fig. 6
(top) a schematic view of the spatial distribution of the field intensity at a given instant of
time. The upper plot in Fig. 6 is not the outcome of the experiment but serves as an
Energy ratio S(Dr)
2
3 ms−10 ms
10 ms−100 ms
Theoretical fit
1.5
1
[500 Hz−1500 Hz ]
0.5
−1.5
−1
−0.5
0
0.5
Distance Dr(m)
1
1.5
FIG. 6. (Top) Schematic snapshot of random spatial intensity fluctuations measured in the coda.
One can imagine that the typical side length is "2 m. The weak localization effect is hidden in this
speckle pattern. (Bottom) Reprinted figure with permission from Larose et al. (2006).
Copyright (2006) by the American Physical Society: Typical profile of intensity enhancement
around the source obtained after averaging over a large number of speckle patterns.
COHERENT BACK-SCATTERING AND WEAK LOCALIZATION
15
illustration of the experimental process involved in the measurement of the weak
localization effect. This complex interference pattern shows rapid spatial fluctuations
and is known in optics as a speckle pattern. In the speckle, the maxima can be viewed as
random constructive interferences between multiply-scattered waves that mask the weak
localization spot (at the center of Fig. 6 (top)). At another time instant, the speckle pattern
will have completely changed, except for the deterministic constructive interference
between reciprocal waves around the source. By averaging speckle patterns over sufficiently long time windows, one suppresses unwanted fluctuations thus revealing the
weak localization intensity pattern as shown in Fig. 6 (bottom). In the present example,
the averaging is performed in a lapse time window ranging from 10 to 100 ms. This
simple procedure makes the link between averaging in theory and in practice. The reader
will notice that the enhancement factor is lower than 2 in this experiment. Presumably,
this is caused by the application of a force with a significant horizontal component, thus
breaking the symmetry between source and detection. Despite the imperfection of the
reconstruction, it is still possible to determine with reasonable accuracy the width of the
weak localization zone as a function of frequency using the theoretical relation (12), thus
providing the dispersion of the fundamental antisymmetric Lamb mode. In Fig. 7, the
results of the dispersion measurements are plotted together with a theoretical fit. Indeed,
the phase velocity c of the fundamental Lamb mode obeys the following rather complicated dispersion relation (see, e.g., Royer and Dieulesaint, 2000)
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
vffiffiffiffiffiffiffiffiffiffiffiffiffiffis
u
u
b2
t 2p
c ¼ pffiffi fhb 1 # 2 ;
a
3
ð17Þ
where f is the frequency, h is the slab thickness, and a and b denote the P and S wave
speeds, respectively. In the present experiment, h is known and the values of the
wavespeeds can be adjusted to fit the experimental data. This technique agrees with
independent measurements to within a few percents, which is highly satisfactory.
Phase velocity (m/s)
2000
1500
1000
From WL width
Theory
500
0
0
500
1000
Frequency (Hz)
1500
2000
FIG. 7. Dispersion relation of the fundamental antisymmetric Lamb mode in a concrete slab.
(Squares) Experimental data obtained from the width of the weak localization zone. The theoretical
relation between the wavelength and the width of the weak localization zone is given by Eq. (12).
(Solid line) Theoretical fit of the experimental data with Eq. (17). Reprinted figure with permission
from Larose et al. (2006). Copyright (2006) by the American Physical Society.
16
MARGERIN
4.2. Measurement of the Diffusion Constant in Strongly Scattering Media
Recent active experiments have shown the extreme character of the propagation in
volcanic regions (see, e.g., Wegler and Lühr, 2001; Wegler, 2004). In the short period
band, it is hardly possible and even sometimes impossible to extract the coherent ballistic
waves from the complex recorded waveform. In such extreme cases, a meaningful
measurement of the medium heterogeneity is the diffusion constant of the waves.
However, in practice, its measurement can be affected by a number of factors, such as
anelastic dissipation or local boundary conditions (see Friedrich and Wegler, 2005 for
details). As demonstrated above, absorption does not influence the measurement of the
weak localization effect. Unfortunately, the weak localization intensity profile does not
depend on the scattering properties of the medium and therefore does not offer direct
access to the scattering properties of the medium. However, we know that weak localization is a multiple scattering process which sets in only after the single scattering
intensity has become negligible. Such a time dependence has been confirmed in a
numerical study by Margerin et al. (2001), which concludes that after a time of the
order of the mean free time, the weak localization pattern is measurable. This has also
been verified experimentally by Larose et al. (2004). They measured the weak localization effect on a volcano in the 10–20 Hz frequency band and found that there exists a
characteristic rise time of the intensity enhancement. Their result is shown in Fig. 8. In
the case of the Puy des Goules, one can infer a mean free time of the order of 1 s which
gives a mean free path of a few hundred meters. Note that in their experiment, the width
of the weak localization depends in a nonlinear way on frequency. This is explained by
the fact that according to the equipartition principle (Weaver, 1990; Hennino et al.,
2001), the energy at the surface is largely dominated by the fundamental mode Rayleigh
wave, whose dispersion is caused by the complex layering at the surface. Although the
S(Dr)
2
1.7 s
0.7 s
0.3 s
1
0
−20
−10
0
Meters
10
20
FIG. 8. Emergence of the weak localization spot in the coda. The normalized intensity is
represented as a function of distance around the source. The dominant frequency of the waves is
15 Hz and the dominant signal is the Rayleigh wave with a wavespeed "250 m/s. The wavelength
is of the order of 15 m. The different curves correspond to the averaged intensity profiles obtained
at different lapse time in the coda (as indicated in the inset). Reprinted figure with permission from
Larose et al. (2004). Copyright (2004) by the American Physical Society.
COHERENT BACK-SCATTERING AND WEAK LOCALIZATION
17
observation of weak localization on a volcano requires active sources, it is free of the
effect of absorption and therefore offers direct access to the scattering properties of the
medium.
5. Conclusion
In this paper, I have given a brief introduction to the weak localization effect in
seismology. I have summarized general formulas that enable the calculation of the weak
localization effect for a wide range of practical cases. The fundamental role of reciprocity
between source and detection has been emphasized and illustrated with experimental
results. In practice, the control of the source mechanism is crucial. The simplest solution
is to measure the vertical displacements generated by vertical forces. Applications to the
characterization of scattering or bulk elastic properties have been presented. In particular, I have shown that the emergence time of weak localization yields an estimate of the
scattering mean free path, independent of absorption effects. Thus, weak localization
combined with other measurements such as time and space dependence of the coda could
be used to discriminate anelastic and scattering attenuation.
Acknowledgments
I would like to thank E. Larose, A Derode, J. de Rosny, and R. Hennino for their invaluable help
with the figures and for many discussions on the weak localization effect. I would like to thank M.
Campillo and B. van Tiggelen for their constant input. Comments by H. Sato, Yu. Kravtsov, and an
anonymous reviewer greatly helped to improve the presentation and the readability of this chapter.
References
Akkermans, E., Montambaux, G. (2005). Physique mésoscopique des électrons et des photons
CNRS Editions, Paris: EDP Sciences, Paris.
Apresyan, L.A., Kravstov, Yu. A. (1996). Radiation Transfer: Statistical and Wave Aspects.
Amsterdam: Gordon and Breach, Amsterdam.
Aubry, A., Derode, A. (2007). Ultrasonic imaging of highly scattering media from local measurements of the diffusion constant: Separation of coherent and incoherent intensities. Phys. Rev. E
75, 026602.
Barabanenkov, Yu.N. (1973). Wave corrections to the transfer equation for “back” scattering.
Radiophysics and Quantum Electronics 16, 65–71.
Barabanenkov, Yu.N. (1975). Multiple scattering of waves by ensembles of particles and the theory
of radiation transport. Sov. Phys. Usp 18, 673–689.
Barabanenkov, Yu.N., Ozrin, V.D. (1991). Asymptotic solution of the Bethe-Salpeter equation and
the Green-Kubo formula for the diffusion constant for wave propagation in random media.
Phys. Lett. A 154, 38–42.
Barabanenkov, Yu.N., Ozrin, V.D. (1995). Diffusion asymptotics of the Bethe-Salpeter equation
for electromagnetic waves in discrete random media. Phys. Lett. A 206, 116–122.
Barabanenkov, Yu.N., Kravtsov, Yu.A., Ozrin, V.D., Saichev, A.I. (1991). Enhanced backscattering in optics. Prog. Opt. 29, 65–197.
Bayer, G., Niederdrank, T. (1993). Weak localization of acoustic waves in strongly scattering
media. Phys. Rev. Lett. 70, 3884–3887.
18
MARGERIN
Campillo, M. (2006). Phase and correlation in ‘random’ seismic fields and the reconstruction of the
Green function. Pure Appl. Geophys. 163, 475–502.
Chandrasekhar, S. (1960). Radiative Transfer. New York: Dover, New York.
Dahlen, F.A., Tromp, J. (1998). Theoretical Global Seismology. Princeton, New Jersey: Princeton
University Press, Princeton, New Jersey.
Derode, A., Mamou, V., Padilla, F., Jenson, F., Laugier, P. (2005). Dynamic coherent backscattering in an absorbing heterogeneous medium: Application to human trabecular bone characterization. Appl. Phys. Lett. 87, 114101.
de Rosny, J., Tourin, A., Fink, M. (2000). Coherent backscattering of elastic wave in a chaotic
cavity. Phys. Rev. Lett. 84, 1693–1695.
de Rosny, J., Tourin, A., Fink, M. (2001). Observation of a coherent backscattering effect with a
dipolar source for elastic waves: Highlight of the role played by the source. Phys. Rev. E 64,
066604, doi:10.1103/PhysRevE.64.066604.
de Wolf, D.A. (1971). Electromagnetic reflection from an extended turbulent medium: Cumulative
forward-scatter single-backscatter approximation. IEEE Trans Antennas Propagation 19,
254–262.
Friedrich, C., Wegler, U. (2005). Localization of seismic coda at Merapi volcano (Indonesia).
Geophys. Res. Lett. 32, L14312, doi: 10.1029/2005GL023111.
Gusev, A., Abubakirov, I.R. (1987). Monte-Carlo simulation of record envelope of a near earthquake. Phys. Earth Planet. Inter. 49, 30–36.
Haney, M., Snieder, R. (2003). Breakdown of wave diffusion in 2D due to loops. Phys. Rev. Lett.
91, 093902.
Hapke, B.W., Nelson, R.M., Smythe, W.D. (1993). The opposition effect of the moon—
The contribution of coherent backscatter. Science 260, 509–511.
Hennino, R., Tregoures, N., Shapiro, N.M., Margerin, L., Campillo, M., van Tiggelen, B.A.,
Weaver, R.L. (2001). Observation of equipartition of seismic waves. Phys. Rev. Lett. 86,
3447–3450.
Hoshiba, M. (1991). Simulation of multiple-scattered coda wave excitation based on the energy
conservation law. Phys. Earth Planet. Inter. 67, 123–136.
Kaveh, M., Rosenbluh, M., Edrei, I., Freund, I. (1986). Weak localization and light scattering from
disordered solids. Phys. Rev. Lett. 57, 2049–2052.
Kuga, Y., Ishimaru, A. (1984). Retroreflection from a dense distribution of spherical particles.
J. Opt. Soc. Am. A 8, 831–835.
Larose, E., Margerin, L., Campillo, M., van Tiggelen, B.A. (2004). Weak localization of seismic
waves. Phys. Rev. Lett. 93, 048501.
Larose, E., de Rosny, J., Margerin, L., Anache, D., Gouedard, P., Campillo, M., van Tiggelen, B.
(2006). Observation of multiple scattering of kHz vibrations in a concrete structure and
application to monitoring weak changes. Phys. Rev. E 73, 016609.
Lesaffre, M., Atlan, M., Gross, M. (2006). Effect of the photon’s Brownian Doppler shift on the
weak-localization coherent-backscattering cone. Phys. Rev. Lett. 97, 033901.
Margerin, L. (2005). Introduction to radiative transfer of seismic waves. In: Levander, A.,
Nolet, G., (Eds.), Seismic Earth: Array Analysis of Broadband Seismograms: Geophysical
Monograph Series, Vol. 157, Chapter 14. AGU, Washington, pp. 229–252.
Margerin, L., Campillo, M., van Tiggelen, B.A. (2001). Coherent backscattering of acoustic waves
in the near field. Geophys. J. Int. 145, 593–603.
Royer, D., Dieulesaint, E. (2000). Elastic Waves in Solids, Part I. Free and Guided Propagation.
Berlin Heidleberg: Springer-Verlag, Berlin Heidleberg.
Sakai, K., Yamamoto, K., Takagi, K. (1997). Observation of acoustic coherent backscattering.
Phys. Rev. B 56, 10930–10933.
Sato, H. (1994). Multiple isotropic scattering model including P-S conversions for the seismogram
envelope formation. Geophys. J. Int. 117, 487–494.
COHERENT BACK-SCATTERING AND WEAK LOCALIZATION
19
Sato, H., Fehler, M. (1998). Seismic Wave Propagation in the Heterogeneous Earth. Heidelberg:
Springer, Heidelberg.
Sheng, P. (1995). Introduction to Wave Scattering, Localization and Mesoscopic Phenomena San
Diego: Academic Press, San Diego.
Snieder, R. (2006). The coherent backscattering effect for moving scatterers. Europhys. Lett. 74,
630–636.
Tourin, A., Roux, P., Derode, A., van Tiggelen, B.A., Fink, M. (1997). Time dependent coherent
backscattering of acoustic waves. Phys. Rev. Lett. 79, 3637–3639.
van Albada, M.P., Lagendijk, A. (1985). Observation of weak localization of light in a random
medium. Phys. Rev. Lett. 55, 2692–2695.
van Tiggelen, B.A., Maynard, R. (1997). Reciprocity and coherent backscattering of light.
In: Papanicolaou, G.R., (Ed.), Wave Propagation in Complex Media (IMA Volumes in Mathematics and Its Applications, Vol. 96). Springer, New York, pp. 247–271.
van Tiggelen, B.A., Margerin, L., Campillo, M. (2001). Coherent backscattering of elastic waves:
Specific role of source, polarization, and near field. J. Acoust. Soc. Am. 110, 1291–1298.
Watson, K. (1969). Multiple scattering of electromagnetic waves in an underdense plasma. J. Math.
Phys. 10, 688–702.
Weaver, R.L. (1990). Diffusivity of ultrasound in polycrystals. J. Mech. Phys. Solids 38, 55–86.
Weaver, R., Burkhardt, J. (1994). Weak Anderson Localization and enhanced backscatter in
reverberation rooms and quantum dots. J. Acoust. Soc. Am. 96, 3186–3190.
Weaver, R.L., Lobkis, O.I. (2000). Enhanced backscattering and modal echo of reverberant elastic
waves. Phys. Rev. Lett. 84, 4942–4945.
Wegler, U. (2004). Diffusion of seismic waves in a thick layer: Theory and application to Vesuvius
volcano. J. Geophys. Res. 109, doi: 10.1029/2004JB003048.
Wegler, U., Lühr, B.G. (2001). Scattering behavior at Merapi volcano (Java) revealed from an
active seismic experiment. Geophys. J. Int. 145, 579–592.
Wegler, U., Korn, M., Przybilla, J. (2006). Modelling full seismogram envelopes using radiative
transfer theory with Born scattering coefficients. Pure Appl. Geophys. 163, 503–531.
Wolf, P.E., Maret, G. (1985). Weak localization and coherent backscattering of photon. Phys. Rev.
Lett. 55, 2696–2699.
Zeng, Y., Su, F., Aki, K. (1991). Scattering wave energy propagation in a random isotropic
scattering medium. 1. Theory. J. Geophys. Res. 96, 607–619.