scispace - formally typeset
Open AccessJournal ArticleDOI

Tutorial on seismic interferometry: Part 1 — Basic principles and applications

Reads0
Chats0
TLDR
In this article, it was shown that the autocorrelation of the transmission response at two receivers along the x-axis gives the Green's function of the direct wave between these receivers.
Abstract
Seismic interferometry involves the crosscorrelation of responses at different receivers to obtain the Green’s function between these receivers. For the simple situation of an impulsive plane wave propagating along the x-axis, the crosscorrelation of the responses at two receivers along the x-axis gives the Green’s function of the direct wave between these receivers. When the source function of the plane wave is a transientas in exploration seismology or a noise signalas in passive seismology, then the crosscorrelation gives the Green’s function, convolved with the autocorrelation of the source function. Direct-wave interferometry also holds for 2D and 3D situations, assuming the receivers are surrounded by a uniform distribution of sources. In this case, the main contributions to the retrieved direct wave between the receivers come from sources in Fresnel zones around stationary points. The main application of direct-wave interferometry is the retrieval of seismic surface-wave responses from ambient noise and the subsequent tomographic determination of the surfacewave velocity distribution of the subsurface. Seismic interferometry is not restricted to retrieving direct waves between receivers. In a classic paper, Claerbout shows that the autocorrelation of the transmission response of a layered medium gives the plane-wave reflection response of that medium. This is essentially 1D reflected-wave interferometry. Similarly, the crosscorrelation of the transmission responses, observed at two receivers, of an arbitrary inhomogeneous medium gives the 3D reflection response of that medium. One of the main applications of reflected-wave interferometry is retrieving the seismic reflection response from ambient noise and imaging of the reflectors in the subsurface. A common aspect of direct- and reflected-wave interferometry is that virtual sources are created at positions where there are only receivers without requiring knowledge of the subsurface medium parameters or of the positions of the actual sources.

read more

Content maybe subject to copyright    Report

Tutorial on seismic interferometry:
Part 1 Basic principles and applications
Kees Wapenaar
1
, Deyan Draganov
1
, Roel Snieder
2
, Xander Campman
3
, and Arie Verdel
3
ABSTRACT
Seismic interferometry involves the crosscorrelation of re-
sponses at different receivers to obtain the Green’s function be-
tween these receivers. For the simple situation of an impulsive
plane wave propagating along the x-axis, the crosscorrelation of
the responses at two receivers along the x-axis gives the Green’s
function of the direct wave between these receivers. When the
source function of the plane wave is a transient as in exploration
seismology or a noise signal as in passive seismology, then the
crosscorrelation gives the Green’s function, convolved with the
autocorrelation of the source function. Direct-wave interferome-
try also holds for 2D and 3D situations, assuming the receivers
are surrounded by a uniform distribution of sources. In this case,
the main contributions to the retrieved direct wave between the
receivers come from sources in Fresnel zones around stationary
points. The main application of direct-wave interferometry is the
retrieval of seismic surface-wave responses from ambient noise
and the subsequent tomographic determination of the surface-
wave velocity distribution of the subsurface. Seismic interferom-
etry is not restricted to retrieving direct waves between receivers.
In a classic paper, Claerbout shows that the autocorrelation of the
transmission response of a layered medium gives the plane-wave
reflection response of that medium. This is essentially 1D reflect-
ed-wave interferometry. Similarly, the crosscorrelation of the
transmission responses, observed at two receivers, of an arbitrary
inhomogeneous medium gives the 3D reflection response of that
medium. One of the main applications of reflected-wave interfer-
ometry is retrieving the seismic reflection response from ambient
noise and imaging of the reflectors in the subsurface. A common
aspect of direct- and reflected-wave interferometry is that virtual
sources are created at positions where there are only receivers
without requiring knowledge of the subsurface medium parame-
ters or of the positions of the actual sources.
INTRODUCTION
In this two-part tutorial, we give an overview of the basic princi-
ples and the underlying theory of seismic interferometry and discuss
applications and new advances. The term seismic interferometry re-
fers to the principle of generating new seismic responses of virtual
sources
4
by crosscorrelating seismic observations at different re-
ceiver locations. One can distinguish between controlled-source and
passive seismic interferometry. Controlled-source seismic interfer-
ometry, pioneered by Schuster 2001, Bakulin and Calvert 2004,
and others, comprises a new processing methodology for seismic ex-
ploration data. Apart from crosscorrelation, controlled-source inter-
ferometry also involves summation of correlations over different
source positions. Passive seismic interferometry, on the other hand,
is a methodology for turning passive seismic measurements ambi-
ent seismic noise or microearthquake responses into deterministic
seismic responses. Here, we further distinguish between retrieving
surface-wave transmission responses Campillo and Paul, 2003;
Shapiro and Campillo, 2004; Sabra, Gerstoft, et al., 2005a and ex-
ploration reflection responses Claerbout, 1968; Scherbaum, 1987b;
Draganov et al., 2007, 2009. In passive interferometry of ambient
noise, no explicit summation of correlations over different source
positions is required because the correlated responses are a superpo-
sition of simultaneously acting uncorrelated sources.
In all cases, the response that is retrieved by crosscorrelating two
Manuscript received by the Editor 30 November 2009; published online 14 September 2010.
1
Delft University of Technology, Department of Geotechnology, Delft, The Netherlands. E-mail: c.p.a.wapenaar@tudelft.nl; d.s.draganov@tudelft.nl.
2
Colorado School of Mines, Center for Wave Phenomena, Golden, Colorado, U.S.A. E-mail: rsnieder@mines.edu.
3
Shell International Exploration and Production, Rijswijk, The Netherlands. E-mail: xander.campman@shell.com; arie.verdel@gmail.com.
© 2010 Society of Exploration Geophysicists.All rights reserved.
4
In the literature on seismic interferometry, the term virtual source often refers to the method of Bakulin and Calvert 2004, 2006, which is discussed extensive-
ly in Part 2. However, creating a virtual source is the essence of nearly all seismic interferometry methods see e.g., Schuster 2001, who already used this term.
In this paper Parts 1 and 2 we use the term virtual source whenever appropriate. When it refers to Bakulin and Calvert’s method, we will mention this explicitly.
GEOPHYSICS, VOL. 75, NO. 5 SEPTEMBER-OCTOBER 2010; P. 75A195–75A209, 15 FIGS.
10.1190/1.3457445
75A195
Downloaded 28 Sep 2012 to 131.180.130.198. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/

receiver recordings and summing over different sources can be in-
terpreted as the response that would be measured at one of the re-
ceiver locations as if there were a source at the other. Because such a
point-source response is equal to a Green’s function convolved with
a wavelet, seismic interferometry is also often called Green’s func-
tion retrieval. Both terms are used in this paper. The term interferom-
etry is borrowed from radio astronomy, where it refers to crosscorre-
lation methods applied to radio signals from distant objects Thomp-
son et al., 2001. The name Green’s function honors George Green
who, in a privately published essay, introduced the use of impulse re-
sponses in field representations Green, 1828. Challis and Sheard
2003 give a brief history of Green’s life and theorem. Ramírez and
Weglein 2009 review applications of Green’s theorem in seismic
processing.
Early successful results of Green’s function retrieval from noise
correlations were obtained in the field of ultrasonics Weaver and
Lobkis, 2001, 2002. The experiments were done with diffuse fields
in a closed system. Here diffuse means that the amplitudes of the nor-
mal modes are uncorrelated but have equal expected energies.
Hence, the crosscorrelation of the field at two receiver positions does
not contain cross-terms of unequal normal modes. The sum of the re-
maining terms is proportional to the modal representation of the
Green’s function of the closed system Lobkis and Weaver, 2001.
This means that the crosscorrelation of a diffuse field in a closed sys-
tem converges to its impulse response. Later, it was recognized e.g.,
Godin, 2007 that this theoretical explanation is akin to the fluctua-
tion-dissipation theorem Callen and Welton, 1951; Rytov, 1956;
Rytov et al., 1989; Le Bellac et al., 2004.
The earth is a closed system; but at the scale of global seismology,
the wavefield is far from diffuse.At the scale of exploration seismol-
ogy, an ambient-noise field may have a diffuse character, but the en-
compassing system is not closed. Hence, for seismic interferometry,
the normal-mode approach breaks down. Throughout this paper, we
consider seismic interferometry or Green’s function retrieval in
open systems, including half-spaces below a free surface. Instead of
a treatment per field of application or a chronological discussion, we
have chosen a setup in which we explain the principles of seismic in-
terferometry step by step. In Part 1, we start with the basic principles
of 1D direct-wave interferometry and conclude with a discussion of
the principles of 3D reflected-wave interferometry. We present ap-
plications in controlled-source as well as passive interferometry and,
where appropriate, review the historical background. To stay fo-
cused on seismic applications, we refrain from a further discussion
of the normal-mode approach, nor do we address the many interest-
ing applications of Green’s function retrieval in underwater acous-
tics e.g., Roux and Fink, 2003; Sabra et al., 2005; Brooks and Ger-
stoft, 2007.
DIRECT-WAVE INTERFEROMETRY
1D analysis of direct-wave interferometry
We start our explanation of seismic interferometry by considering
an illustrative 1D analysis of direct-wave interferometry. Figure 1a
shows a plane wave, radiated by an impulsive unit source at x x
S
and t 0, propagating in the rightward direction along the x-axis.
We assume that the propagation velocity c is constant and the medi-
um is lossless. There are two receivers along the x-axis at x
A
and x
B
.
Figure 1b shows the response observed by the first receiver at x
A
.We
denote this response as Gx
A
,x
S
,t, where G stands for the Green’s
function. Throughout this paper, we use the common convention
that the first two arguments in Gx
A
,x
S
,t denote the receiver and
source coordinates, respectively here, x
A
and x
S
, whereas the last
argument denotes time t or angular frequency
. In our example, this
Green’s function consists of an impulse at t
A
x
A
x
S
/ c; there-
fore, Gx
A
,x
S
,t
t t
A
, where
t is the Dirac delta function.
Similarly, the response at x
B
is given by Gx
B
,x
S
,t
t t
B
, with
t
B
x
B
x
S
/ c Figure 1c.
Seismic interferometry involves the crosscorrelation of responses
at two receivers, in this case at x
A
and x
B
. Looking at Figure 1a,itap-
pears that the raypaths associated with Gx
A
,x
S
,t and Gx
B
,x
S
,t
have the path from x
S
to x
A
in common. The traveltime along this
common path cancels in the crosscorrelation process, leaving the
traveltime along the remaining path from x
A
to x
B
, i.e., t
B
t
A
x
B
x
A
/ c. Hence, the crosscorrelation of the responses in Figure 1b
and c is an impulse at t
B
t
A
see Figure 1d. This impulse can be in-
terpreted as the response of a source at x
A
observed by a receiver at
x
B
, i.e., the Green’s function Gx
B
,x
A
,t. An interesting observation is
that the propagation velocity c and the position of the actual source
x
S
need not be known. The traveltimes along the common path from
x
S
to x
A
compensate each other, independent of the propagation ve-
locity and the length of this path. Similarly, if the source impulse
would occur at t t
S
instead of at t 0, the impulses observed at x
A
and x
B
would be shifted by the same amount of time t
S
, which would
be canceled in the crosscorrelation. Thus, the absolute time t
S
at
which the source emits its pulse need not be known.
Let us discuss this example a bit more precisely. We denote the
crosscorrelation of the impulse responses at x
A
and x
B
as
Gx
B
,x
S
,t Gx
A
,x
S
,t. The asterisk denotes temporal convolu-
tion, but the time reversal of the second Green’s function turns the
convolution into a correlation, defined as Gx
B
,x
S
,t Gx
A
,x
S
, t
Gx
B
,x
S
,t t
Gx
A
,x
S
,t
dt
. Substituting the delta functions
into the right-hand side gives
t t
t
B
t
t
A
dt
t
t
B
t
A
兲兲
t x
B
x
A
/ c. This is indeed the Green’s func-
tion Gx
B
,x
A
,t, propagating from x
A
to x
B
. Because we started this
derivation with the crosscorrelation of the Green’s functions, we
have obtained the following 1D Green’s function representation:
Figure 1. A 1D example of direct-wave interferometry. a A plane
wave traveling rightward along the x-axis, emitted by an impulsive
source at x x
S
and t 0. b The response observed by a receiver
at x
A
. This is the Green’s function Gx
A
,x
S
,t. c As in b but for a re-
ceiver at x
B
. d Crosscorrelation of the responses at x
A
and x
B
. This is
interpreted as the response of a source at x
A
, observed at x
B
, i.e.,
Gx
B
,x
A
,t.
75A196 Wapenaar et al.
Downloaded 28 Sep 2012 to 131.180.130.198. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/

Gx
B
,x
A
,t Gx
B
,x
S
,t Gx
A
,x
S
, t. 1
This representation formulates the principle that the crosscorrelation
of observations at two receivers x
A
and x
B
gives the response at one
of those receivers x
B
as if there were a source at the other receiver
x
A
. It also shows why seismic interferometry is often called
Green’s function retrieval.
Note that the source is not necessarily an impulse. If the source
function is defined by some wavelet st, then the responses at x
A
and
x
B
can be written as ux
A
,x
S
,t Gx
A
,x
S
,t st and ux
B
,x
S
,t
Gx
B
,x
S
,t st, respectively. Let S
s
t be the autocorrelation of
the wavelet, i.e., S
s
t st s t. Then the crosscorrelation of
ux
A
,x
S
,t and ux
B
,x
S
,t gives the right-hand side of equation 1, con-
volved with S
s
t. This is equal to the left-hand side of equation 1,
convolved with S
s
t. Therefore,
Gx
B
,x
A
,t S
s
t ux
B
,x
S
,t ux
A
,x
S
, t. 2
In words: If the source function is a wavelet instead of an impulse,
then the crosscorrelation of the responses at two receivers gives the
Green’s function between these receivers, convolved with the auto-
correlation of the source function.
This principle holds true for any source function, including noise.
Figure 2a and b shows the responses at x
A
and x
B
, respectively, of a
bandlimited noise source Nt at x
S
the central frequency of the
noise is 30 Hz; the figure shows only 4 s of a total of 160 s of noise.
In this numerical example, the distance between the receivers is
1200 m and the propagation velocity is 2000 m / s; hence, the travel-
time between these receivers is 0.6 s. As a consequence, the noise re-
sponse at x
B
in Figure 2b is 0.6 s delayed with respect to the response
at x
A
in Figure 2a similar to the impulse in Figure 1c delayed with re-
spect to the impulse in Figure 1b. Crosscorrelation of these noise re-
sponses gives, analogous to equation 2, the impulse response be-
tween x
A
and x
B
, convolved with S
N
t, i.e., the autocorrelation of the
noise Nt. The correlation is shown in Figure 2c, which indeed re-
veals a bandlimited impulse centered at t 0.6 s the traveltime
from x
A
to x
B
. Note that from registrations at two receivers of a noise
field from an unknown source in a medium with unknown propaga-
tion velocity, we have obtained a bandlimited version of the Green’s
function. By dividing the distance between the receivers 1200 m
by the traveltime estimated from the bandlimited Green’s function
0.6 s, we obtain an estimate of the propagation velocity between
the receivers 2000 m / s. This illustrates that direct-wave interfer-
ometry can be used for tomographic inversion.
Until now, we considered a single plane wave propagating in the
positive x -direction. In Figure 3a, we consider the same configura-
tion as in Figure 1a, but now an impulsive unit source at x x
S
radi-
ates a leftward-propagating plane wave. Figure 3b is the response at
x
A
, given by Gx
A
,x
S
,t
t t
A
, with t
A
x
S
x
A
/ c. Similarly,
the response at x
B
is Gx
B
,x
S
,t
t t
B
, with t
B
x
S
x
B
/ c
Figure 3c. The crosscorrelation of these responses gives
t t
B
t
A
兲兲
(t x
B
x
A
/ c), which is equal to the time-reversed
Green’s function Gx
B
,x
A
, t. So, for the configuration of Figure
3a, we obtain the following Green’s function representation:
Gx
B
,x
A
, t G x
B
,x
S
,t Gx
A
,x
S
,t. 3
We can combine equations 1 and 3 as follows:
Gx
B
,x
A
,t Gx
B
,x
A
,t
i1
2
Gx
B
,x
S
i
,t Gx
A
,x
S
i
,t,
4
where x
S
i
for i 1,2 stands for x
S
and x
S
, respectively.
For the 1D situation, this combination may not seem very useful.
We analyze it here, however, because this representation better re-
sembles the 2D and 3D representations we encounter later. Note that
because Gx
B
,x
A
,t is the causal response of an impulse at t 0
meaning it is nonzero only for t 0, it does not overlap with
Gx
B
,x
A
, t which is nonzero only for t 0. Hence, Gx
B
,x
A
,t
can be resolved from the left-hand side of equation 4 by extracting
the causal part. If the source function is a wavelet st with autocor-
relation S
s
t, we obtain, analogous to equation 2,
Figure 2. As in Figure 1 but this time for a noise source Nt at x
S
. a
The response observed at x
A
, i.e., ux
A
,x
S
,t Gx
A
,x
S
,t Nt. b
As in a but for a receiver at x
B
. c The crosscorrelation, which is
equal to Gx
B
,x
A
,t S
N
t, with S
N
t the autocorrelation of the
noise.
Figure 3. As in Figure 1 but this time for a leftward-traveling impul-
sive plane wave. The crosscorrelation in d is interpreted as the
time-reversed Green’s function Gx
B
,x
A
,t.
Tutorial on interferometry: Part 1 75A197
Downloaded 28 Sep 2012 to 131.180.130.198. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/

Gx
B
,x
A
,t Gx
B
,x
A
, t兲其 S
s
t
i1
2
ux
B
,x
S
i
,t ux
A
,x
S
i
, t. 5
Here, Gx
B
,x
A
,t S
s
t may have some overlap with Gx
B
,x
A
,
t S
s
t for small t, depending on the length of the autocorrela-
tion function S
s
t. Therefore, Gx
B
,x
A
,t S
s
t can be extracted
from the left-hand side of equation 5, except for small distances x
B
x
A
.
The right-hand sides of equations 4 and 5 state that the crosscorre-
lation is applied to the responses of each source separately, after
which the summation over the sources is carried out. For impulsive
sources or transient wavelets st, these steps should not be inter-
changed. Let us see why. Suppose the sources at x
S
and x
S
act simul-
taneously, as illustrated in Figure 4a. Then the response at x
A
would
be given by ux
A
,t
i1
2
Gx
A
,x
S
i
,t st and the response at x
B
by
ux
B
,t
j1
2
Gx
B
,x
S
j
,t st. These responses are shown in Fig-
ure 4b and c for an impulsive source st
t兲兲. The crosscorrela-
tion of these responses, shown in Figure 4d, contains two cross-
terms at t
B
t
A
and t
B
t
A
that have no physical meaning. Hence, for
impulsive or transient sources, the order of crosscorrelation and
summation matters.
The situation is different for noise sources. Consider two simul-
taneously acting noise sources N
1
t and N
2
t at x
S
and x
S
, respec-
tively. The responses at x
A
and x
B
are given by ux
A
,t
i1
2
Gx
A
,x
S
i
,t N
i
t and ux
B
,t
j1
2
Gx
B
,x
S
j
,t N
j
t, re-
spectively see Figure 5a and b. Because each of these responses is
the superposition of a rightward- and a leftward-propagating wave,
the response in Figure 5b is not a shifted version of that in Figure 5a
unlike the responses in Figure 2a and b. We assume that the noise
sources are uncorrelated; thus, N
j
t N
i
t兲典
ij
S
N
t, where
ij
is the Kronecker delta function and · denotes ensemble averag-
ing. In practice, the ensemble averaging is replaced by integrating
over sufficiently long time. In the numerical example the duration of
the noise signals is again 160 s only 4 s of noise is shown in Figure
5a and b. For the crosscorrelation of the responses at x
A
and x
B
,we
can now write
ux
B
,t ux
A
, t兲典
j1
2
i1
2
Gx
B
,x
S
j
,t N
j
t
Gx
A
,x
S
i
, t N
i
t
i1
2
Gx
B
,x
S
i
,t Gx
A
,x
S
i
, t S
N
t. 6
Combining equation 6 with equation 4, we finally obtain
Gx
B
,x
A
,t Gx
B
,x
A
, t兲其 S
N
t ux
B
,t ux
A
,t兲典.
7
Expression 7 shows that the crosscorrelation of two observed fields
at x
A
and x
B
, each of which is the superposition of rightward- and left-
ward-propagating noise fields, gives the Green’s function between
x
A
and x
B
plus its time-reversed version, convolved with the autocor-
relation of the noise see Figure 5c. The cross-terms, unlike Figure
4d, do not contribute because the noise sources N
1
t and N
2
t are
uncorrelated.
Miyazawa et al. 2008 apply equation 7 with x
A
and x
B
at different
depths along a borehole in the presence of industrial noise at Cold
Lake, Alberta, Canada. By choosing for u different components of
multicomponent sensors in the borehole, they retrieve separate
Green’s functions for P- and S-waves, the latter with different polar-
izations. From the arrival times in the Green’s functions, they derive
the different propagation velocities and accurately quantify shear-
wave splitting.
Despite the relative simplicity of our 1D analysis of direct-wave
interferometry, we can make several observations about seismic in-
terferometry that also hold true for more general situations. First, we
can distinguish between interferometry for impulsive or transient
Figure 4. As in Figures 1 and 3 but with simultaneously rightward-
and leftward-traveling impulsive plane waves. The crosscorrelation
in d contains cross-terms that have no physical meaning.
Figure 5. As in Figure 4 but this time with simultaneously rightward-
and leftward-traveling uncorrelated noise fields. The crosscorrela-
tion in c contains no cross-terms.
75A198 Wapenaar et al.
Downloaded 28 Sep 2012 to 131.180.130.198. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/

sources on the one hand equations 4 and 5 and interferometry for
noise sources on the other hand equation 7. In the case of impulsive
or transient sources, the responses of each source must be crosscor-
related separately, after which a summation over the sources takes
place. In the case of uncorrelated noise sources, a single crosscorre-
lation suffices.
Second, it appears that an isotropic illumination of the receivers is
required to obtain a time-symmetric response between the receivers
of which the causal part is the actual response. In one dimension,
isotropic illumination means equal illumination by rightward- and
leftward-propagating waves. In two and three di-
mensions, it means equal illumination from all di-
rections discussed in the next section.
Finally, instead of the time-symmetric re-
sponse Gx
B
,x
A
,t Gx
B
,x
A
,t, in the litera-
ture we often encounter an antisymmetric re-
sponse Gx
B
,x
A
,t Gx
B
,x
A
,t. This is merely
a result of differently defined Green’s functions.
Note that a simple time differentiation of the
Green’s functions would turn the symmetric re-
sponse into an antisymmetric one, and vice versa
see Wapenaar and Fokkema 2006 for a more
detailed discussion on this aspect.
2D and 3D analysis of direct-wave
interferometry
We extend our discussion of direct-wave inter-
ferometry to configurations with more dimen-
sions. In the following discussion, we mainly use
heuristic arguments, illustrated with a numerical
example. For a more precise derivation based on
stationary-phase analysis, we refer to Snieder
2004.
Consider the 2D configuration shown in Figure
6a. The horizontal dashed line corresponds to the
1D configuration of Figure 1a, with two receivers
at x
A
and x
B
, 1200 m apart x denotes a Cartesian
coordinate vector. The propagation velocity c is
2000 m / s, and the medium is again assumed to
be lossless. Instead of plane-wave sources, we
have many point sources denoted by the small
black dots, distributed over a “pineapple slice,”
emitting transient signals with a central frequen-
cy of 30 Hz. In polar coordinates, the positions of
the sources are denoted by r
S
,
S
. The angle
S
is equidistantly sampled
S
0.25° , whereas
the distance r
S
to the center of the slice is chosen
randomly between 2000 and 3000 m. The re-
sponses at the two receivers at x
A
and x
B
are
shown in Figure 6b and c, respectively, as a func-
tion of the polar source coordinate
S
for dis-
play purposes, only every sixteenth trace is
shown. These responses are crosscorrelated for
each source separately, and the crosscorrelations
are shown in Figure 6d, again as a function of
S
.
Such a gather is often called a correlation gather.
Note that the traveltimes in this correlation gather
vary smoothly with
S
, despite the randomness of
the traveltimes in Figure 6b and c. This is because in the crosscorre-
lation process only the time difference along the paths to x
A
and x
B
matters.
The source in Figure 6a with
S
0° plays the same role as the
plane-wave source at x
S
in Figure 1a. For this source, the crosscorre-
lation gives a signal at x
B
x
A
/ c 0.6 s, seen in the trace at
S
0° in Figure 6d. Similarly, the source at
S
180° plays the same
role as the plane-wave source at x
S
in Figure 3a and leads to the trace
at
S
180° in Figure 6d with a signal at 0.6 s. Analogous to
equation 5, we sum the crosscorrelations of all sources, i.e., we sum
Figure 6. A 2D example of direct-wave interferometry. a Distribution of point sources,
isotropically illuminating the receivers at x
A
and x
B
. The thick dashed lines indicate the
Fresnel zones. b Responses at x
A
as a function of the polar source coordinate
S
. c
Responses at x
B
. d Crosscorrelation of the responses at x
A
and x
B
. The dashed lines indi-
cate the Fresnel zones. e The sum of the correlations in d. This is interpreted as
Gx
B
,x
A
,t Gx
B
,x
A
, t兲其 S
s
t. The main contributions come from sources in the
Fresnel zones indicated in a and d. f Single crosscorrelation of the responses at
x
A
and x
B
of simultaneously acting uncorrelated noise sources. The duration of the noise
signals was 9600 s.
Tutorial on interferometry: Part 1 75A199
Downloaded 28 Sep 2012 to 131.180.130.198. Redistribution subject to SEG license or copyright; see Terms of Use at http://segdl.org/

Citations
More filters
Journal ArticleDOI

Shear wave imaging from traffic noise using seismic interferometry by cross-coherence

TL;DR: In this article, the authors apply the cross-coherence method to the seismic interferometry of traffic noise, which originates from roads and railways, to retrieve both body waves and surface-waves.
Journal ArticleDOI

Surface and borehole ground-penetrating-radar developments

TL;DR: The specific aspects of borehole radar are discussed and recent developments to become more sensitive to orientation and to exploit the supplementary information in different components in polarimetric uses of radar data are described.
Journal ArticleDOI

Geophysical Monitoring of Moisture-Induced Landslides: A Review

TL;DR: Whiteley et al. as discussed by the authors presented a review of the state of the art of geophysical monitoring applied to moisture-induced landslides, focusing on technical and practical uses of time-lapse methods in geophysics applied to monitoring moistureinduced landslide.
Journal ArticleDOI

Thickness and structure of the martian crust from InSight seismic data

TL;DR: In this paper, the authors determine the structure of the crust beneath the InSight landing site on Mars using both marsquake recordings and the ambient wavefield, and find that the observations are consistent with models with at least two and possibly three interfaces.
Journal ArticleDOI

Estimating near‐surface shear wave velocities in Japan by applying seismic interferometry to KiK‐net data

TL;DR: In this article, the authors estimate shear wave velocities in shallow subsurface throughout Japan by applying seismic interferometry to the data recorded with KiK-net, a strong motion network in Japan.
References
More filters
Journal ArticleDOI

Irreversibility and Generalized Noise

TL;DR: In this article, a relation between the generalized resistance and the generalized forces in linear dissipative systems is obtained, which forms the extension of the Nyquist relation for the voltage fluctuations in electrical impedances.
Book ChapterDOI

The Dispersion of Surface Waves on Multilayered Media

TL;DR: In this paper, a matrix formalism developed by W. T. Thomson is used to obtain the phase velocity dispersion equations for elastic surface waves of Rayleigh and Love type on multilayered solid media.
Book

Interferometry and Synthesis in Radio Astronomy

TL;DR: In this paper, a theory of interferometry and synthesis imaging analysis of the Interferometer Response Geometric Relationships and Polarimetry Antennas and Arrays Response of the receiving system Design of the Analog Receiving System Digital Signal Processing Very-Long-Baseline Interferometry Calibration and Fourier Transformation of Visibility Data Deconvolution, Adaptive Calibrration, and Applications Interferometers Techniques for Astrometry and Geodesy Propagation Effects Van Cittert-Zernike Theorem, Spatial Coherence, and
Journal ArticleDOI

High-resolution surface-wave tomography from ambient seismic noise.

TL;DR: Cross-correlation of 1 month of ambient seismic noise recorded at USArray stations in California yields hundreds of short-period surface-wave group-speed measurements on interstation paths that are used to construct tomographic images of the principal geological units of California.
Journal ArticleDOI

Transmission of Elastic Waves through a Stratified Solid Medium

TL;DR: In this article, the transmission of a plane elastic wave at oblique incidence through a stratified solid medium consisting of any number of parallel plates of different material and thickness is studied theoretically.
Related Papers (5)
Frequently Asked Questions (1)
Q1. What contributions have the authors mentioned in the paper "Utorial on seismic interferometry: art 1 — basic principles and applications" ?

In this paper, a simple situation of an impulsive plane wave propagating along the x-axis gives the Green 's function of the direct wave between these receivers.