Sie sind auf Seite 1von 13

GEOPHYSICS, VOL. 74, NO. 2 共MARCH-APRIL 2009兲; P. S11–S23, 5 FIGS.

10.1190/1.3052116
Downloaded 08/18/19 to 103.47.158.161. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

True-amplitude Gaussian-beam migration

Samuel H. Gray1 and Norman Bleistein2

steep dips, and its decomposition of neighboring seismic traces into


ABSTRACT angular information allows it to propagate different angles to any
subsurface location independently, achieving multipathing. Hill’s
Gaussian-beam depth migration and related beam migra- prestack method 共2001兲 extends his work by presenting an efficient
tion methods can image multiple arrivals, so they provide an 3D migration, suitable for data acquired along a narrow azimuth.
accurate, flexible alternative to conventional single-arrival Nowack et al. 共2003兲 and Gray 共2005兲 remove the narrow azimuth
Kirchhoff migration. Also, they are not subject to the steep- restriction by presenting variations suitable for 共possibly wide-azi-
dip limitations of many 共so-called wave-equation兲 methods muth兲 shot-record migration. All of this work establishes Gaussian-
that use a one-way wave equation in depth to downward-con- beam migration as a powerful imaging technique, with accuracy
tinue wavefields. Previous presentations of Gaussian-beam comparable to wave-equation migration and flexibility comparable
migration have emphasized its kinematic imaging capabili- to Kirchhoff migration.
ties without addressing its amplitude fidelity. We offer two Albertin et al. 共2004兲 present the first treatment of amplitude pres-
true-amplitude versions of Gaussian-beam migration. The ervation in Gaussian-beam migration, with an approach suitable for
first version combines aspects of the classic derivation of common-offset, narrow-azimuth data. Aside from being limited to a
prestack Gaussian-beam migration with recent results on
single acquisition geometry, the approach of Albertin et al. relies on
true-amplitude wave-equation migration, yields an expres-
a common-offset Beylkin determinant that avoids caustic problems
sion involving a crosscorrelation imaging condition. To pro-
by allowing time to be a complex quantity. 共The Beylkin determi-
vide amplitude-versus-angle 共AVA兲 information, true-ampli-
nant is a Jacobian of transformation from image coordinates, when
tude wave-equation migration requires postmigration map-
the image is formed, to surface coordinates, where the imaging sum-
ping from lateral distance 共between image location and
mation takes place兲. This Jacobian is much more difficult to compute
source location兲 to subsurface opening angle. However,
than the Jacobian for common-shot migration.
Gaussian-beam migration does not require postmigration
We present an alternative approach, based on a generalized strate-
mapping to provide AVA data. Instead, the amplitudes and di-
rections of the Gaussian beams provide information that the gy for true-amplitude migration 共Bleistein et al., 2005兲. 共Here true-
migration can use to produce AVA gathers as part of the amplitude migration means to produce amplitudes on reflectors that
migration process. The second version of true-amplitude are proportional, up to a constant factor, to P-wave angle-dependent
Gaussian-beam migration is an expression involving a de- plane-wave reflection coefficients.兲 This strategy has been applied to
convolution imaging condition, yielding amplitude-varia- true-amplitude wave-equation migration 共Zhang et al., 2007兲, and
tion-with-offset 共AVO兲 information on migrated shot-do- we follow the application closely. Our formulation also depends on
main common-image gathers. the Jacobian of a mapping from subsurface coordinates to surface
coordinates; however, our Jacobian is much more straightforward to
understand and compute than that of Albertin et al. 共2004兲. 共On the
other hand, optimizing the Gaussian-beam migration formulas in-
INTRODUCTION volves high-frequency asymptotics, which complicate the final mi-
gration weights.兲
Beginning with Hill’s 共1990兲 paper on poststack migration, Gaus- Our Jacobian is also more general, in that it is not limited to imag-
sian-beam migration has been recognized as an accurate and versa- ing data acquired using a particular geometry even though it was first
tile migration technique. Its reliance on ray tracing allows it to image presented in the context of imaging complete wavefields. It applies,

Manuscript received by the Editor 20 May 2008; revised manuscript received 8 October 2008; published online 11 February 2009.
1
CGGVeritas, Calgary, Alberta, Canada. Email: sam.gray@cggveritas.com.
2
Colorado School of Mines, Department of Geophysics, Center for Wave Phenomena, Golden, Colorado, U.S.A. Email: normblei@gmail.com; norm@dix
.mines.edu.
© 2009 Society of Exploration Geophysicists. All rights reserved.

S11
S12 Gray and Bleistein

for example, to both common-offset and common-shot data; its only hoff migration; but, like our method, they provide greater accuracy
real requirement is that all input data 共e.g., all offsets or all shots兲 and allow for imaging into ADCIGs.
containing information from a particular image location be included GRT migration takes a dual approach to Kirchhoff migration but
to provide accurate amplitudes from that location. Zhang et al. obtains the same result. As described by Miller et al. 共1987兲, GRT
共2007兲 apply this Jacobian as a postmigration mapping from shot- migration reconstructs each subsurface scatterer by accumulating
domain common-image gathers 共SDCIGs兲, whose traces are in-
Downloaded 08/18/19 to 103.47.158.161. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

reflection data from all isochrons that intersect at the scatterer loca-
dexed by the lateral distance between image location and source lo- tion. At actual reflector locations, the live data that dominate the sum
cation, to angle-domain common-image gathers 共ADCIGs兲, whose are from the isochron that passes tangent to the reflector surface. The
traces are indexed by the opening angle at the imaged reflectors. conceptual distinction of GRT migration lies in performing the im-
This approach to true-amplitude imaging can be applied to any aging operation in the subsurface, as opposed to Kirchhoff migra-
migration method — in particular, to Gaussian-beam migration. In tion’s summation of data expressed in terms of surface coordinates.
addition, because the ray-based Gaussian beam contains the angle Although the isochrons are determined by the acquisition geometry
information as the imaging takes place, we can apply the mapping as well as the subsurface velocity structure, the quality of the image
implicitly during imaging, migrating directly into ADCIGs. and its amplitudes is determined by the geometry of the isochrons at
We present a derivation of true-amplitude Gaussian-beam migra- the scatterer locations. As with Kirchhoff and Gaussian-beam mi-
tion that follows those of Zhang et al. 共2007兲 and Hill 共2001兲. The grations, a reflector is well imaged only if a portion of an isochron
general framework applies to any migration method: crosscorrelat- passes tangent to it. In Gaussian-beam migration, the isochrons are
ing downward-continued wavefields from source and receiver loca- composed of local portions, each determined by the real part of the
tions. We write the wavefields in terms of Green’s functions for the traveltime surface for one of the localized Gaussian beams.
wave equation, and we express the Green’s functions as Gaussian- Several workers 共e.g., Deng and McMechan, 2007兲 mention
beam expansions. Our derivation is in terms of wavefields, but a shortcomings of true-amplitude migration, either from complexities
modification of the Green’s functions by Hill 共2001兲 allows us to in the migration velocity leading to uneven subsurface illumination
consider how migration acts on single traces 共even after they have or from elastic or anelastic effects that lie outside acoustic inverse
been slant stacked into neighboring beam center locations兲, as in theory. Such statements apply equally to all migration methods.
Kirchhoff migration. The derivation contains details of the steepest- Also, shortcomings in processing before migration, or interpretation
descent calculation used by Hill 共2001兲 to reduce the number of inte- after migration, can compromise the fidelity of migrated amplitudes.
grals, which correspond to loops in the computer implementation. 共In fact, the simpler of our two examples illustrates the effect of one
共We perform this analysis only for the 2D case; the corresponding of these problems — incomplete illumination resulting from finite
3D analysis is beyond the mathematical scope of this paper.兲 recording or migration aperture — on migrated amplitudes.兲 Al-
Zhang et al. 共2007兲 show that the crosscorrelation imaging condi- though these shortcomings imply limitations in our ability to ana-
tion is appropriate for true-amplitude migration when it is followed lyze migrated amplitudes except in the simplest situations, we take
by mapping from distance to angle. Here, we show that the same is the position that migrated amplitudes should be as accurate as possi-
true for Gaussian-beam migration. In addition, we show that the ble within the acoustic limits assumed by most migration methods. If
crosscorrelation imaging condition is appropriate for true-amplitude the cost of including accurate migration amplitudes is only a small
Gaussian-beam migration when imaging directly into ADCIGs us- part of the total migration cost, the effort of including accurate am-
ing ray information available to the migration, without the need to plitudes is warranted.
image into intermediate SDCIGs. We also show the relationship be-
tween the crosscorrelation imaging condition and the deconvolution TRUE-AMPLITUDE GAUSSIAN-BEAM
imaging condition usually used for true-amplitude Kirchhoff migra- MIGRATION
tion. A very simple example explains this relationship.
Because the action of Gaussian-beam migration on a single input
The 2D formula — Crosscorrelation imaging condition
trace can be analyzed in a manner similar to Kirchhoff and general- Prestack Gaussian-beam migration was developed first for marine
ized Radon transform 共GRT兲 migrations, we mention some connec- streamer-style common-offset, common-azimuth records 共Hill,
tions with those methods. As described by Hanitzsch 共1997兲, early 2001; Albertin et al., 2004兲 and later for common-shot records
versions of true-amplitude Kirchhoff migration relied on Bleistein’s 共Nowack et al., 2003; Gray, 2005兲. Because our presentation relies
共1987兲 modification of Beylkin’s 共1985兲 theory of imaging disconti- on observations about true-amplitude wave-equation migration, we
nuities. Bleistein 共1987兲 treats the imaging problem as an inverse begin with recorded data in the form of an actual wavefield, i.e.,
problem that integrates reflection data over acquisition surface coor- common-shot records. Although this approach differs from Hill’s
dinates. He incorporates Beylkin’s 共1985兲 geometric term into his 共2001兲 derivation of Gaussian-beam migration, most of the steps are
solution for the weight factor that makes the composition of the in- identical. In fact, if we consider the process that the method applies
version operator with the forward modeling operator a formal identi- to any given input trace in a well-sampled survey, we find that the or-
ty, i.e., the inverse operator applied to reflection data yields a reflec- der of the input data is irrelevant; we can apply the method to input
tivity model. traces gathered in any order with essentially identical migrated re-
Many subsequent true-amplitude Kirchhoff migration results rely sults. If the input traces are well sampled spatially 共receiver loca-
on this approach, which is limited to situations where the wavefields tions and shot locations at regular inline x and crossline y intervals
are well behaved 共i.e., away from caustics兲. Later generalizations al- with spacing that is fine enough to allow the unaliased slant stack of
low for imaging in greater structural complexity by removing the the highest temporal frequencies present in the recorded data at the
caustic restriction and allowing for multiple arrivals 共e.g., Xu et al., highest value of horizontal slowness兲 and if the subsurface illumina-
2001; Bleistein et al., 2005兲. These generalizations are much more tion at an image location from the downward-continued wavefields
complicated than conventional single-arrival true-amplitude Kirch- from source and receiver locations is nonzero, then Gaussian-beam
True-amplitude Gaussian-beam migration S13

migration will produce a migrated common-image gather 共CIG兲 at corded wavefield pU, the response to a 2D line source, is dimension-
that location that allows amplitude-versus-angle 共AVA兲 analysis. less. Inserting expressions 2 and 3 into equation 1 results in
In a second departure from Hill’s 共2001兲 derivation, we derive
true-amplitude Gaussian-beam migration in two dimensions only
and present the 3D expression without derivation. For efficiency, the
2D and 3D derivations rely on a saddle-point 共or steepest-descent兲
I共x;xs兲 ⳱ⳮ4 冕 d␻ i␻ 兩␻ 兩
cos ␪ s
Vs
G*共x;xs ; ␻ 兲
Downloaded 08/18/19 to 103.47.158.161. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

integral 共Hill, 2001兲, an asymptotic technique that generalizes the


method of stationary phase to complex exponential functions whose
phase is also complex. 共This happens because the formulation of
⫻ 冕 dxr pU共xr ;xs ; ␻ 兲
cos ␪ r
Vr
G*共x;xr ; ␻ 兲. 共4兲

Gaussian beams, a type of dynamic ray tracing, allows time t, which The choice of Green’s function determines the migration method.
appears in the phase, to be a complex quantity: t ⳱ tRe Ⳮ itIm, where Choosing a Green’s function of the form AART exp关i␻ ␶ 兴, where AART
i ⳱ 冑ⳮ1. By contrast, the dynamic ray tracing used by standard and ␶ are both real, will lead to Kirchhoff migration. Choosing a nu-
Kirchhoff migration requires time along a raypath to remain real: t merical wavefield extrapolator based on a one-way wave equation
⳱ tRe.兲 The 2D derivation requires a saddle-point integral in a single will lead to a wave-equation migration.
complex variable, and the 3D derivation requires the integral in two Instead of these choices, we write the Green’s function as a sum-
complex variables. Because the theory for saddle-point integrals in mation of Gaussian beams 共Hill, 1990, 2001兲:
more than one complex variable is very technical without shedding


light on the problem at hand, we present it elsewhere.
i dpx
We begin with the expression for a single migrated 2D shot record G共x;x⬘ ; ␻ 兲 ⳱ uGB共x;x⬘ ; ␻ 兲
with shot location xs ⳱ 共xs,0兲, receiver locations xr ⳱ 共xr,0兲, and 4␲ pz


image locations x ⳱ 共x,z兲, using the imaging condition that is the
i dpx
crosscorrelation of the downward-continued upgoing and downgo- ⳱ AGB exp关i␻ TGB兴. 共5兲
ing wavefields 共Zhang et al., 2007兲: 4␲ pz

I共x;xs兲 ⳱ ⳮi 冕 d␻ sgn共␻ 兲pU共x;xs ; ␻ 兲pD


* 共x;x ; ␻ 兲,
s
This expression is an integral of individual Gaussian beams uGB over
horizontal slownesses px; pz ⳱ cos ␪ /V is the vertical slowness. In
contrast to the amplitudes and times in equations 2 and 3, Gaussian-
共1兲 beam amplitude AGB and traveltime TGB 共developed in Appendices A
and B兲 are complex quantities. From now on, we drop the subscript
where pU and pD are, respectively, the upgoing 共recorded兲 and the
from AGB and TGB.
downgoing 共source兲 wavefields and * denotes complex conjuga-
tion. With the postmigration mapping described by Zhang et al. Equations 4 and 5 show the versatility of the Gaussian-beam ap-
共2007兲, this expression can be modified to convert SDCIGs into AD- proach. Each term uGB represents a partial wavefield in a limited spa-
CIGs. Still following Zhang et al. 共2007兲, we write asymptotic forms tial region surrounding the beam’s central raypath. These beams are
for pU and pD* , also expressing them in terms of Green’s functions for independent of one another as they range through the subsurface, al-
the 2D wave equation: lowing the overlapping beams to contribute multiple arrivals to the
wavefield at any subsurface location. This differs from Kirchhoff

pU共x;xs ; ␻ 兲 ⳱ 2 冕 dxr pU共xr ;xs ; ␻ 兲


cos ␪ r
Vr
冑ⳮ i␻ AART,r
migration, whose Green’s functions can accommodate multipathing
only with considerable effort. As mentioned, several workers have
made this extra effort.
⫻ exp关ⳮ i␻ ␶ r兴 For Gaussian-beam migration, Hill 共1990, 2001兲 chooses initial


conditions for the dynamic ray-tracing equations that ensure 共1兲 lo-
⳱ⳮ2 dxr pU共xr ;xs ; ␻ 兲 calized planar beams uGB at initial source and receiver locations and
共2兲 regular, nonsingular behavior of the individual beams through-
out the subsurface. These depend on sensible choices for parameters
cos ␪ r w0, the initial half-width of the Gaussian beams, and ␻ r, the refer-
⫻ i␻ G*共x;xr ; ␻ 兲, 共2兲
Vr ence frequency. 共A typical value for ␻ r is 2␲ ⫻ 10 Hz, and a typical
value for w0 is one wavelength at the reference frequency.兲
cos ␪ s The first of these conditions allows us to think of decomposing the
* 共x;x ; ␻ 兲 ⳱ ⳮ2
pD s
冑ⳮ i␻ AART,s exp关ⳮ i␻ ␶ s兴 source and recorded wavefields pD and pU into local plane-wave
Vs components that match the initial directions of the Gaussian beams.
cos ␪ s In particular, it allows us to consider propagating these components
⳱2 i␻ G*共x;xs ; ␻ 兲. 共3兲 into the subsurface using individual beams uGB. If this can be accom-
Vs
plished, then we can migrate the local plane-wave components from
Here, Vs and Vr are surface velocities at the source and receiver a set of surface locations that is much sparser than the original set of
points; ␪ s and ␪ r are takeoff angles of the rays from source and re- source and receiver locations. Hill 共1990, 2001兲 provides recipes for
ceiver points to image location x; AART,s, AART,r, and ␶ s, ␶ r are real- this procedure, which is the heart of Gaussian-beam migration. In
valued amplitudes and traveltimes from asymptotic ray theory; and essence, Gaussian-beam migration replaces the migration of indi-
G共x;x⬘ ; ␻ 兲 is the Green’s function for the wave equation. The di- vidual data traces over all directions into the subsurface by the
mensionalities of pU and pD* are different: in two dimensions, the migration of localized, directional components of the wavefield with
source wavefield pD* has dimensions of inverse length, but the re- unique initial directions into the subsurface. Thus, at selected beam-
S14 Gray and Bleistein

center locations, local slant stacks are performed and the data mi-
grated over those initial directions using Gaussian beams as the
propagators.
Ds共L,pLx, ␻ 兲 ⳱ 冏 冏 冕

␻r
3/2
dxr pU共xr ;xs ; ␻ 兲

To make this rigorous, we first describe the local slant stack, then ⫻ exp关ⳮi␻ pLx共xr ⳮ L兲兴

冋冏冏 册
the partition of unity that permits each input trace to contribute to a
␻ 共xr ⳮ L兲2
Downloaded 08/18/19 to 103.47.158.161. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

number of beam centers with weights that sum to unity. The local ⫻ exp ⳮ 共8兲
slant stack acts as a simple delay applied to an input trace; the delay ␻r 2w20
accounts for the traveltime difference between a plane wave arriving
at the actual receiver location and the same plane wave arriving at and where AL and TL are Gaussian-beam amplitude and time from
the beam center location L ⳱ 共L,0兲. In the frequency domain, this is beam center location L. Notation for receiver slowness changes
accomplished by applying the phase shift exp关ⳮi␻ px共xr ⳮ L兲兴 to from pr to pL to emphasize that all receiver-side rays emanate from
the input trace whose receiver is located at 共xr,0兲. Hill 共2001兲 incor- L.
porates this phase shift as a modification to his Green’s function ex- Next, we transform from source and receiver slownesses ps and pL
pansion 5, and Gray 共2005兲 suggests a modification to this phase to midpoint and offset slownesses pm and ph 共Hill, 2001; Gray, 2005兲,
shift when the elevations of L and xr are different. 共Effectively, this
factor is used to approximate the wavefield resulting from a point pmx ⳱ pLx Ⳮ psx , 共9兲
source at 共x⬘,0兲 when the source of ray propagation is at a nearby lo-
cation 共L,0兲.兲 phx ⳱ pLx ⳮ psx ,
Because the Gaussian-beam wavefront is perfectly planar only at
its initial location on the beam center and gradually accumulates to arrive at the migration formula:
nonzero curvature away from the beam center, approximating the
共nearly linear兲 delay by the linear slant stack incurs some kinematic
error; however, this error usually is negligible. Potentially more seri-
ous is the amplitude error incurred when the origin of the Green’s
I共x;xs兲 ⳱
8␲
⌬L␻ r
2 冑2␲ w0 ⫻ 兺
L
冕 d␻ i␻ 冕冕 dpmxdphx
psz pLz
function 共the receiver location兲 is far from the initial point of the ray cos ␪ s cos ␪ L * *
tracing 共the beam center兲. However, Hill’s choice of partition of uni- As AL exp关ⳮ i␻ 共Ts* Ⳮ TL*兲兴
Vs VL
ty ensures that the traces far from a beam center are downweighted
exponentially relative to traces close to a beam center, whose ampli- ⫻Ds共L,pLx, ␻ 兲. 共10兲
tude error is small. We use this approximation in our synthetic exam-
ples; its error is not enough to degrade the migrated amplitudes sig- This formula expresses a migrated record as a sum over beam-center
nificantly. locations, where each beam center contributes partial images from
The partition of unity is given by Hill 共1990兲 as many directions. We note again that A and T are complex valued be-
cause of our use of Green’s functions composed of Gaussian beams.

1⬇
⌬L
冑2␲ w0 冑冏 冏 兺 冋 冏 冏

␻r L
exp ⳮ
␻ 共xr ⳮ L兲2
␻r 2w20

, 共6兲
To reduce the computational load of equation 10, Hill 共2001兲 sug-
gests evaluating the integral over phx with its steepest-descent ap-
proximation. 共If A and T were both real valued, we would instead use
a stationary-phase approximation. In the 2D case, in fact, the steep-
where ⌬L is the spacing between beam-center locations.Asimple in- est-descent approximation to the integral agrees formally with a sta-
terpretation of this formula is as a discrete approximation, with ap- tionary-phase approximation, with the normally real quantities in
propriate change of variable, to the formula for the infinite integral of the stationary-phase formula replaced by complex quantities.兲
a Gaussian function: 兰ⳮ ⬁ 冑
⬁ exp共ⳮx /2兲dx ⳱ 2 ␲ . Other partitions of
2 In Hill’s 共2001兲 common-offset derivation, local slant-stacked re-
unity are possible, but equation 6 has benign effects on amplitude flection data D depend directly on pmx; but in the present common-
and it generalizes easily to three dimensions. shot formulation, D depends on pmx through pLx. In the steepest-de-
When we insert equations 5 共modified by the phase shift to ac- scent calculation for fixed pmx, we find the value of phx that minimiz-
count for the difference between receiver and beam-center loca- es the imaginary part of the total traveltime T ⳱ Ts Ⳮ TL. This sta-
0
tions兲 and 6 into equation 4 and rearrange terms, we obtain tionary value phx is used with pmx in equation 9 to determine critical
0 0
values psx , pLx ; these are used to evaluate D at the time corresponding
0 0 0
to the total time along beams psx and pLx . The critical values psx ,
⌬L␻ r
I共x;xs兲 ⳱ pLx are also related to vertical slownesses psz ⳱ cos ␪ s /Vs, pLz
0 0 0

4␲ 2冑2␲ w0 ⳱ cos ␪ L /VL, canceling those factors appearing in equation 10. We

冕 冕
show the details of the steepest-descent integral in Appendix A; in-
cos ␪ s dpsx *
⫻兺 d␻ i␻ A exp关ⳮ i␻ Ts*兴 serting its result into equation 10 leads to
Vs psz s

冕 冕
L
⌬L␻ r

cos ␪ L
VL
冕 dpLx *
A exp关ⳮ i␻ TL*兴Ds共L,pLx, ␻ 兲,
pLz L
I共x;xs兲 ⳱ 兺
8 ␲ 2w 0 L
d␻ 冑i␻ dpmx

共7兲 As*AL* exp关ⳮ i␻ T*兴


⫻ Ds共L,pLx
0
, ␻ 兲, 共11兲
冑T*⬙共phx0 兲
where
True-amplitude Gaussian-beam migration S15

where T* ⳱ Ts* Ⳮ TL*. In this equation, As* and Ts* are evaluated Equations 2–6 are substituted into equation 13, and source and re-
along beam psx 0
; and D, AL*, and TL* are evaluated along beam pLx
0
. The ceiver slownesses ps, pr are transformed to midpoint and offset slow-
real parts of Ts* and TL* identify traveltimes from source and receiver nesses pm, ph in the numerator exactly as in equation 9 to arrive at a
beam-center locations to the image location. migration formula:
The second derivative, T*⬙共phx 0
兲, is assumed to be nonzero; it is
⌬L␻ r cos ␪ s
Downloaded 08/18/19 to 103.47.158.161. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

evaluated in Appendix B as R共x;xs兲 ⳱ ⳮ


16␲ 3冑2␲ w0 Vs
T *⬙共p0h兲 ⳱ ⳮ␻ rw20 冋 1 Im关Qs兴
Vs2 psz
2
Qs
1 Im关QL兴
Ⳮ 2 2
VL pLz QL
, 册 ⫻兺 冕 d␻
i␻
*
pD pD
冕冕 dpmxdphx cos ␪ s
psz pLz Vs
L
共12兲
cos ␪ L * *
where Qs and QL are complex dynamic ray-tracing quantities de- ⫻ As AL exp关ⳮi␻ T*兴Ds共L,pLx, ␻ 兲. 共14兲
scribed in Appendix B and Im关Q兴 denotes the imaginary part of Q. VL
For either source or receiver, the vertical slowness pz ⳱ cos ␪ /V,
This formula is similar to equation 10 except for the presence of
where V is velocity at source or receiver beam center location and ␪
pD pD* in the denominator and an extra factor of ⳮ共1/2␲ 兲共cos ␪ s /Vs兲
is takeoff angle relative to the vertical of the raypath from source or
⫻兩␻ 兩. We approximate the integral over ph in the numerator by the
beam center to image location.
method of steepest descent, with the same result as before, multi-
plied by psz ⳱ 冑1 ⳮ Vs2 psx
The total error of our final migration formula is the combination of 2
/Vs, evaluated at the critical value of psx.
errors from 共1兲 our approximation in equation 4 for the Green’s func-
The product pD pD* also can be expressed as a product of integrals
tions, 共2兲 the weighted time shifts applied to associate each input
using equations 3 and 5; because the product has no phase, we do not
trace with a variety of beam centers 共the slant stack兲, 共3兲 the approxi-
need the machinery of the steepest-descent method to evaluate it. It
mation in the steepest-descent integral, and 共4兲 kinematic and dy-
can be evaluated numerically at each image location, arriving at an
namic errors arising from insufficient ray illumination in the subsur-
expression for the total illumination from the source. Instead, using
face. Results from numerous numerical studies suggest that the fail-
equations 3, 5, and B-8, we approximate each integral by its leading-
ure of the Gaussian-beam approximation of equation 4 can be a
order asymptotic expression, arriving at
problem when the velocity structure is extraordinarily complex and
detailed. On the other hand, our ability to estimate highly complicat-
ed migration velocities remains limited, and the advantages of effi-
ciency, flexibility, and steep-dip performance of Gaussian-beam mi-
R共x;xs兲 ⳱ ⳮ
⌬L␻ r cos ␪ s
8 ␲ 2w 0 V s L
兺 冕 d␻ 冑i␻ 冕 dpmx

gration usually result in better-than-adequate imaging — even com- As*AL*兩Ts⬙共psx


0
兲兩exp关ⳮi␻ T*兴
pared with wave-equation and reverse-time methods — in complex ⫻ Ds共L,pLx
0
, ␻ 兲.
velocity structures that are not specified with perfect precision. As 兩As兩 冑T*⬙共phx
2 0

mentioned, the kinematic and dynamic errors from the slant-stack 共15兲
approximation appear to be negligible, even for land data with mod-
erately varying surface elevation and near-surface velocity. As in equation 11, A and T are complex. When multiple arrivals from
The approximation in the steepest-descent integral can be serious the source occur at an image location, the expression in the denomi-
when the magnitude of T ⬙共phx 0
兲 is small. It is difficult to predict when nator will be inaccurate and equation 15 will be a kinematically ac-
this can happen, except to say that it will occur only when the behav- curate imaging formula without amplitude fidelity. As Zhang et al.
ior of the Gaussian beams is pathological; this, in turn, is associated 共2007兲 point out in their equation 9, using equation 15 produces SD-
with velocity complexities. In any case, even if Gaussian-beam mi- CIGs whose traces contain reflection coefficients 共multiplied by a
gration 共or any migration兲 produces good images in the presence of constant factor兲; as with the crosscorrelation imaging condition 11,
extreme velocity variation, uneven subsurface illumination will pre- the traces in each SDCIG are indexed by the lateral distance between
clude the analysis of migrated amplitudes. Uneven subsurface illu- image location and source location.
mination is more likely to occur with Gaussian-beam migration, Both of our derivations can be interpreted in two different ways.
with its reliance on ray tracing, than with wave-equation migration. First, we can view them as procedures for downward-continuing
wavefields followed by application of an imaging condition. This is
the classical approach to presenting wave-equation migration; in
The 2D formula: Deconvolution imaging condition fact, we rely on a recent advance in true-amplitude wave-equation
Minor changes allow us to produce a migration formula using the migration theory 共Zhang et al., 2007兲 to derive equation 11 from the
deconvolution imaging condition. This imaging condition is written crosscorrelation imaging condition. Alternatively, we can view the
differently from equation 1; in fact, we modify the standard decon- derivations as procedures for migrating individual input traces onto
volution imaging condition used for wave-equation migration to the output locations. For each input trace and a given beam center and
2D true-amplitude migration expression derived by Keho and Bey- emergence angle, we first perform a weighted delay that associates
doun 共1988兲. This is equivalent to expression 9 of Zhang et al. the trace with the emergence angle at the beam center and then map
共2007兲: the trace into the subsurface using complex amplitude and travel-
time functions. This interpretation generalizes standard Kirchhoff

R共x;xs兲 ⳱
1 cos ␪ s
2␲ Vs
冕 d␻ i␻
pU共x;xs ; ␻ 兲pD
* 共x;x ; ␻ 兲

pD共x;xs ; ␻ 兲pD
s
* 共x;x ; ␻ 兲
s
.
migration to local slant-stack migration, variations of which have
gained popularity in recent years.
As a consequence of the second interpretation, the order of the in-
共13兲 put traces is largely irrelevant. For example, the same derivations
S16 Gray and Bleistein

can be applied to common-offset input data volumes 共Hill, 2001兲. * 兴兩exp关ⳮi␻ T*兴
As*AL*兩det关Ts,ij
⫻ Ds共L,pLx
0 0
,pLy , ␻ 兲.
兩As兩2冑det关Tij*兴
For each input trace and a given beam center and incident and emer-
gence angles, the initial delay associates the trace with the incident
and emergence angles at the source and receiver locations corre- 共17兲
sponding to the beam center 共Hill, 2001, Figure 2b兲. At the end of the
Downloaded 08/18/19 to 103.47.158.161. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

procedure, which includes mapping the migrated data to opening an- In equations 16 and 17,

冏 冏 冕冕
gles, each input trace will have been migrated in a true-amplitude
sense. True-amplitude SDCIGs will be output when the deconvolu- ␻ 3
Ds共L,pLx,pLy, ␻ 兲 ⳱ dxrdy r pU共xr ;xs ; ␻ 兲
tion imaging condition is applied; true-amplitude ADCIGs will be ␻r
output when the crosscorrelation imaging condition is applied.
⫻ exp关i␻ pL · 共xr ⳮ L兲兴

冋冏冏 册
As an implementation detail, we note the amplitudes A in formu-
las 11 and 14 are complex. These normalized amplitudes are given ␻ 兩xr ⳮ L兩2
in Appendix B as 冑VQ0 /V0Q, where V is local wavespeed, Q is a ⫻ exp ⳮ 共18兲
complex dynamic ray-tracing quantity, and subscript zero refers to ␻r 2w20
the initial location of the ray. Although most of the other quantities is the 2D generalization of the weighted slant stack of equation 8, L
are real or complex in some familiar fashion, complex-valued Q is ⳱ 共Lx,Ly,0兲, xr ⳱ 共xr,y r,0兲, pL ⳱ 共pLx,pLy,pLz兲, and the sum is a
relatively unfamiliar. In Gaussian-beam migration, it is real initially double sum over Lx and Ly. Furthermore, Tij* ⳱ 关⳵ 2T*„ph0 …/⳵ phx⳵ phy兴
and gradually acquires a nonzero imaginary part as a ray propagates is the Hessian matrix of second derivatives of T* evaluated at critical
from the source or the receiver beam center into the subsurface. The point ph0 , replacing T*⬙ in the steepest-descent equation 12, and Ts,ij
*
phases of Q from the source and receiver then combine with the is a similar Hessian for Ts*. Under our assumption that Qs and QL are
phase of exp关ⳮi␻ 共Ts* Ⳮ TL*兲兴/冑T*⬙共p0h兲 to form a total phase shift diagonal, the determinant of Tij* is given by

冋 册
that needs to be applied to the local slant-stack data D while comput-
ing the image. ␻ rw20 ␻ rw20
det关Tij*兴 ⳱ det Im关Qs兴Qsⳮ1 Ⳮ Im关QL兴QLⳮ1 ,
Vs psz VL pLz
The 3D true-amplitude migration formulas 共19兲
Except for the presence of quantities in all three spatial dimen- * is the same as equation 19 without the
and the determinant of Ts,ij
sions, the derivations of 3D true-amplitude migration formulas pro- second term on the right-hand side. Complex matrices Qs and QL are
ceed as for the 2D derivations. Hill 共2001兲 and Zhang et al. 共2007兲 scalar multiples of the 2 ⫻ 2 identity matrix. The scalars have the
provide a template for these derivations. As mentioned, however, the same values as solutions Qs and QL of the 2D dynamic ray equations,
steepest-descent integrals that make up the final step are more com- making it easy to evaluate the determinant numerically.
plicated in the 3D case. Here, we present the final migration expres-
sions for both crosscorrelation and deconvolution imaging condi-
tions. As a technical condition, in three dimensions, P and Q are 2D Mapping surface offset to opening angles
complex matrices that are diagonal initially 共Hill, 2001兲; the matrix With SDCIGs, it is easier to attach a physical meaning to the de-
quotient PQⳮ1 must remain diagonally dominant for the following convolution imaging condition than to the crosscorrelation imaging
formulas to hold. We force this condition by assuming that second condition; understanding this meaning helps in mapping Gaussian-
spatial derivatives of velocity appearing in the coupled differential beam migrated data from distance to subsurface opening angle. The
equations for P and Q can be neglected. This is equivalent to assum- deconvolution imaging condition forms the quotient of the upgoing
ing that the velocity is a piecewise linear function of the spatial coor- 共reflected兲 wavefield with the downgoing 共incident兲 wavefield. At a
dinates except for isolated negligible terms near velocity interfaces. reflector location, the quotient is the reflection coefficient at the an-
If the velocity function is smoothed so that its relative variations on gle of incidence, or half-opening angle, measured at a particular lat-
the order of a wavelength are small, this is not a damaging assump- eral offset from image point to source location. This heuristic obser-
tion for kinematic ray behavior, although it can affect ray amplitudes vation can be justified rigorously and used to help form ADCIGs for
beneath significant velocity boundaries, such as salt interfaces. Gaussian-beam migration.
Neglecting the second spatial derivatives of velocity leads to sca- With Gaussian-beam migration, we know the locations of the
lar matrices P and Q, which facilitates the evaluation of the steepest- source and receiver beam centers. If, at a reflector, we also know the
descent integrals. Then the 3D migration expressions for crosscorre- ray directions from those locations, we can assign the migrated am-
lation and deconvolution imaging conditions are, respectively, plitude to the opening angle. Gaussian-beam migration provides this

冕 冕冕
information as it migrates data: The subsurface ray angles at image
冑3V共x兲⌬Lx⌬Ly␻ r2
I共x;xs兲 ⳱ ⳮ
16␲ 2w20
兺L d␻ dpmxdpmy
locations from source and receiver beam centers can be determined
and applied after migration to map offsets in SDCIGs to angles in
ADCIGs. This is different from 共and more cumbersome than兲 apply-
As*AL* exp关ⳮi␻ T*兴 ing ray directions during imaging to produce ADCIGs without going
⫻ Ds共L,pLx
0 0
,pLy , ␻ 兲, 共16兲 through the intermediate step of producing SDCIGs 共Gray, 2007兲.
冑det关Tij*兴 Where true-amplitude migration is valid, these ADCIGs are suitable
for AVA analysis. In the absence of the mapping from offset to inci-
冑3⌬Lx⌬Ly␻ r2 cos ␪ s
冕 冕冕
dence angle, Gaussian-beam migration using the deconvolution im-
R共x;xs兲 ⳱ ⳮ
16␲ 2w20 Vs
兺L d␻ dpmxdpmy aging condition produces SDCIGs that can be applied in amplitude-
versus-offset 共AVO兲 studies in areas of mild structural complexity.
True-amplitude Gaussian-beam migration S17

In the wave-equation migration formulation of Zhang et al. computer memory and disk needed to hold the larger gathers. As our
共2007兲, the crosscorrelation imaging condition 1 provides SDCIGs; use of wide-azimuth data increases, so will our use of CIGs indexed
the migrated amplitude in a particular offset bin is directly propor- by two quantities — either x- and y-offset or opening and azimuth
tional to the product of angle-dependent reflection coefficient multi- angles.
plied by the angular width of the offset bin. In particular, the angular
Downloaded 08/18/19 to 103.47.158.161. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

width of a 2D offset bin is


EXAMPLES
⌬␪ ⳱ 8␲ AART,s
2
cos
␪ s⌬x
Vs
冉 冊
, 共20兲 We present two examples: a constant-velocity synthetic and a
complex structural synthetic. The simplicity of the constant-velocity
example allows us to demonstrate the nature of the mapping from
where AART,s is the asymptotic ray-theoretic amplitude along the ray lateral distance to subsurface angle. This example models reflections
from the source point to the image point, ␪ s is the takeoff angle rela- from density contrasts in a medium of constant velocity 2000 m/s.
tive to the vertical of that ray at the source point, ⌬x is the spatial Four horizontal reflectors with identical reflection coefficients are
width of the offset bin 共maximum offset minus minimum offset placed at depths of 1000, 2000, 3000, and 4000 m. A single shot
within the bin兲, and Vs is the wavespeed at the source point. From record, with a recording aperture of 7000 m on either side of the shot
equation 20, near-offset bins 共small ␪ s兲 of a given size contain a larg- point, is migrated; the results of true-amplitude GBM are shown in
er range of opening angles than far-offset bins 共large ␪ s兲 of the same Figures 1–3. Half-opening angles were limited to 60° in the migra-
size. In other words, the crosscorrelation imaging condition produc- tion.
es angle-dependent reflection coefficients normalized by d␪ s /dx or In Figure 1, the crosscorrelation imaging condition is used, and in
reflection coefficient densities. Equation 20 provides the map from Figure 2 the deconvolution imaging condition is used. Figure 1a
lateral distance to opening angle. shows a decay in amplitudes with depth, similar to the horizontal-re-
Zhang et al. 共2007兲 point out the difficulty of this procedure for flector example of Zhang et al. 共2007兲. Because of the lateral transla-
wave-equation migration when rays, not directly related to the tional symmetry of the problem for this simple geometry, the migrat-
downward continuation, are used for the mapping. As an alternative,
they propose a modified imaging condition that produces CIGs in-
a)
dexed by subsurface offset; these can be transformed into ADCIGs
without relying on ray information. Obviously, this extra work is not
needed for Gaussian-beam migration, where the ray information is 1000
available and well behaved when the conditions for true-amplitude
imaging are met. Also, as mentioned, the two-step process of form-
ing SDCIGs and transforming them into ADCIGs is relatively cum- 2000
bersome.
Depth (m)

When SDCIGs are formed, deconvolution imaging condition 15


produces reflection coefficients as a function of lateral distance, but
3000
crosscorrelation imaging condition 11 produces reflection coeffi-
cient density. Reflection coefficients are more useful than a density
that needs to be mapped into reflection coefficients, so the deconvo-
lution imaging condition is theoretically preferable when migrating 4000

into SDCIGs. 共As we show, however, the situation reverses when we


migrate directly into ADCIGs: there, the crosscorrelation imaging
condition produces reflection coefficients.兲 On the practical side, the 5000
-5000 0 5000
deconvolution imaging condition contains a quotient that needs to Distance (m)
be regularized, but the crosscorrelation imaging condition contains
only a product.Although the regularization can be straightforward to b)
perform, it might compromise the accuracy of the migrated ampli-
tudes, especially in areas where the source illumination is weak.
An extra issue that arises in three dimensions is absent in two di-
mensions: the presence of a second angle, the subsurface azimuth.
Typically, this angle is ignored, with all azimuths summing into each
surface or subsurface offset or angle bin. If a reflector is illuminated
equally at all azimuths, then summing all azimuths together scales
the migrated amplitudes equally for each opening angle, with no
change to the migrated amplitudes except for a single scale factor. If
this is not the case, summing all azimuths together scales the migrat-
ed amplitudes differently at different opening angles, introducing an
error into the AVA analysis. In Gaussian-beam migration, the same
angle information used to identify the opening angle at an image lo- Figure 1. 共a兲 Migrated image 共crosscorrelation imaging condition兲
from a single shot record in a constant-velocity medium with hori-
cation is also available to identify the subsurface azimuth angle. Full zontal reflectors caused by identical density contrasts. Amplitudes
use of this information will result in ADCIGs indexed by both open- decay visibly from top to bottom. 共b兲 Normalized amplitudes along
ing angle and subsurface azimuth angle, with considerable extra the reflectors in 共a兲.
S18 Gray and Bleistein

ed record is equivalent to an SDCIG. To convert the gather into an with the same fidelity as in Figure 1, with aperture truncation effects
ADCIG, we must perform one of the mappings proposed by Zhang that are more visible because of the relatively larger amplitudes pro-
et al. 共2007兲. In this constant-velocity, horizontal reflector case, duced at greater distances.
equation 20 provides an analytic relative decay function for the am- As mentioned, when the crosscorrelation imaging condition pro-
plitudes, vertically and laterally. In two dimensions, squared ampli- vides SDCIGs, the migrated amplitude in a particular distance bin is
is proportional to 1/冑x2 Ⳮ z2 and cos ␪ s ⳱ z/冑x2 Ⳮ z2, so
Downloaded 08/18/19 to 103.47.158.161. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

2
tude AART,s directly proportional to the product of angle-dependent reflection
the crosscorrelation imaging condition produces a constant angle- coefficient multiplied by angular width of the distance bin. The same
dependent reflection coefficient scaled by a factor proportional to the imaging condition can be used with subsurface angle information
decay function z/共x2 Ⳮ z2兲. available in Gaussian-beam migration to produce ADCIGs directly
Figure 1b shows the migrated amplitudes along the reflectors; in the migration. This is done by 共1兲 summing the ray angles from the
these amplitudes are the approximations, generated by the migra- source and beam center at each image location to find the opening
tion, to the decay function. Figure 1b also illustrates the exact decay angle for migrating a sample onto an image location and 共2兲 placing
function for the shallowest reflector, showing good agreement with the migrated amplitude into the bin corresponding to that opening
the migrated amplitudes. Migration aperture truncation artifacts be- angle. Then the amplitude in an angle bin is proportional to the prod-
gin to interfere with the migrated amplitudes at distances corre- uct of angle-dependent reflection coefficient multiplied by angular
sponding to half-opening angles approaching 60°, and migrated am- width of the angle bin. In other words, the laterally and vertically
plitudes decay rapidly beyond those distances, nullifying any benefit varying normalization of equation 20 becomes proportional to the
of amplitude preservation. width of the angle bins. Because the angle bins are all the same size,
In Figure 2 共deconvolution imaging condition兲, the goal is to pro- the normalization becomes a constant, so the crosscorrelation imag-
duce amplitudes proportional to reflection coefficients, without any ing condition is appropriate for migrating directly in the subsurface
vertically and laterally varying scaling. This goal is accomplished opening angle domain. 共For the same reason, the deconvolution im-

a) a)

Figure 3(a)
1000

2000
Depth (m)

3000

4000

5000
0 80
Angle (°)

b) b)

Figure 3. 共a兲 Migrated CIG, with the crosscorrelation imaging con-


Figure 2. 共a兲 Same as Figure 1a except that the deconvolution imag- dition applied in the opening-angle domain. Angle bins range from
ing condition is applied. Amplitudes are similar for all reflectors. 共b兲 zero to 80°. Amplitudes are similar for all reflectors. 共b兲 Normalized
Normalized amplitudes along the reflectors in 共a兲. Amplitude arti- amplitudes along the reflectors in 共a兲. Amplitudes have not been pre-
facts near the maximum offsets are migration aperture truncation ef- served as accurately as in Figures 1 and 2 because the binning of off-
fects. sets into angles is inaccurate, especially at shallow depths.
True-amplitude Gaussian-beam migration S19

aging condition produces ADGICs whose traces have the inverse records, using the crosscorrelation imaging condition. Figure 3
of the decay function implied by equation 20. Thus, using the cross- shows an ADCIG at a single location. Disappointingly, the ampli-
correlation imaging condition produces a decay at far offsets in the tude behavior is worse than for the SDCIG amplitudes of Figure 1.
SDCIGs as in Figure 1, and using the deconvolution imaging condi- This problem is caused by the granularity of the opening-angle bin-
tion produces an unwanted extra gain in ADCIGs at large angles.兲 ning. Specifically, the small number of migration grid points in each
Downloaded 08/18/19 to 103.47.158.161. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

We illustrate the crosscorrelation imaging condition on angle angle bin at the smallest depths causes the angular width of the offset
gathers in Figure 3. Instead of mapping the migrated record of Figure bins in equation 20 to be calculated with significant jitter. Far from
1 from distance to angle as Zhang et al. 共2007兲 do, we build ADCIGs the source, where the angle bins contain a large number of image lo-
directly by migrating a range of 共translationally invariant兲 shot cations, this jitter decreases.
Although the aperture truncation effects slightly visible in Figure
1b and very visible in Figure 2b are present in Figure 3b, their effect
is dominated by the inaccuracy of the numerical approximation to
equation 20. As this example shows, these truncation effects can be
hard to detect 共especially when the crosscorrelation imaging condi-
tion is used兲, and they are not always considered when amplitude
analysis on CIGs is performed. However, as Figure 2b shows, aper-
ture truncation effects may cause significant errors on amplitude
analysis.
Our second example uses the Sigsbee2a model data set 共Paffen-
holz, 2001兲. We produced the migrated stack of Figure 4 by using the
crosscorrelation imaging condition to migrate the input traces into
ADCIGs, followed by stack. We show some ADCIGs in Figure 5,
corresponding to those shown by Zhang et al. 共2007兲. Qualitatively,
these gathers resemble those obtained by true-amplitude wave-
equation migration, but there are differences in amplitude details.
Generally, the Gaussian-beam migration amplitudes are weaker be-
neath salt than the wave-equation migration amplitudes. Because of
the complexity of the model, it is impossible to determine which
Figure 4. Stacked Gaussian-beam migrated image of the Sigsbee2a method produces more correct amplitudes, but the generally lower
data set using the true-amplitude crosscorrelation imaging condi- noise level in the wave-equation migrated gathers and stack suggest
tion. that the relative amplitudes produced by true-amplitude wave-equa-
tion migration are better.

CONCLUSIONS
We have derived true-amplitude versions of Gaussian-beam mi-
gration. Our derivation combines two migration methods 共classical
Gaussian-beam migration and true-amplitude wave-equation mi-
gration兲, applying a commonly used criterion for true-amplitude mi-
gration. This application produces expressions 11 and 15 in two di-
mensions and expressions 16 and 17 in three dimensions for true-
amplitude migration with crosscorrelation and deconvolution imag-
ing conditions. Our final expressions are similar to published results
on Gaussian-beam migration, with modified migration weights. The
modifications arise from a detailed analysis of a steepest-descent
calculation used to approximate a computationally intensive inte-
gral; the modified weights prove useful in migration. Migration
weights obtained without using the results of the steepest-descent
calculation produce migrated amplitudes that are similar to those
produced here, modified by a slowly varying scale factor. That is, the
2D terms involving T*⬙共p0h兲 in equation 12 and the 3D terms involv-
ing detTij* in equation 19 vary slowly in the far field of the source
and receiver beam-center locations, where the high-frequency ap-
proximation of the steepest-descent calculation is valid. Therefore,
neglecting these terms produces an amplitude error that is slowly
varying and can be difficult to measure. This illustrates that high-fre-
Figure 5. ADCIGs at five locations from true-amplitude Gaussian- quency asymptotic methods such as steepest descent produce first-
beam migration. Several diffraction and reflection events are order effects on the kinematic behavior of wavefields but only sec-
marked as in Zhang et al. 共2007兲. ond-order effects on the dynamic behavior.
S20 Gray and Bleistein

We also have provided physical interpretations of the amplitudes We proceed by assuming there is a point zs on the interval of inte-
on migrated traces using both imaging conditions when migrating gration for which
into CIGs indexed by surface distance or using subsurface angle.
The deconvolution imaging condition produces reflection coeffi- d⌿ 共zs兲 d2⌿ 共zs兲
⳱ 0, ⌿⬙ ⳱ ⫽ 0. 共A-2兲
cients 共multiplied by a constant factor兲 when migrating into SD- dz dz2
Downloaded 08/18/19 to 103.47.158.161. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

CIGs, and the crosscorrelation imaging condition produces reflec-


tion coefficients 共multiplied by a constant factor兲 when migrating These conditions echo the conditions of a simple stationary point in
into ADCIGs. Our examples illustrate these observations. the method of stationary phase. However, the interpretation in the
complex plane is somewhat different. In particular, for 兩z ⳮ zs兩
small,
ACKNOWLEDGMENTS
We thank Ross Hill for several extremely interesting conversa- 共z ⳮ zs兲2
⌿ 共z兲 ⳮ ⌿ 共zs兲 ⬇ ⌿ ⬙. 共A-3兲
tions on this topic. We also thank Uwe Albertin, Yu Zhang, and an 2
anonymous reviewer for very constructive comments.
Let u共x,y兲 ⳱ Re关⌿ 共z兲 ⳮ ⌿ 共zs兲兴 be the real part of the complex dif-
ference in ⌿ . Because of the quadratic approximation in equation
APPENDIX A
A-3, it is fairly easy to show that the surface u共x,y兲 is locally hyper-
bolic — visually, a saddle. Hence, the point zs is interchangeably
METHOD OF STEEPEST DESCENT FOR called a stationary point or a saddle point.
ASYMPTOTIC EVALUATION OF CERTAIN The Cauchy integral theorem tells us that the path of integration
COMPLEX INTEGRALS can be deformed through the saddle point in a direction that is fortu-
itous for asymptotic approximation of the integral, i.e., on a path
In this appendix, we describe the asymptotic expansion of an in-
where ⳮu rises up to the saddle point and then falls again on the oth-
tegral with a complex exponent by the method of steepest descent. In
er side. If we think of those two half-paths as directed away from the
our 2D Gaussian-beam migration application, the integral is charac-
saddle point for the moment, they are the paths of steepest descent of
terized by a single complex-valued function of a real variable. In-
the real part of the exponent, hence the term method of steepest
volving only a single complex variable, the application is a 1D steep-
descent. Along those paths, the exponential has the form
est-descent integral. In 3D Gaussian-beam migration, the integral
exp关ⳮas2 /2兴, with a real and s a new real variable of integration. In
over x- and y-offset ray parameters leads to a double integral over
this case, the integral can be evaluated asymptotically by the Laplace
two complex-valued functions of two real variables, or a 2D steep-
method 共Bleistein, 1984兲, which is the analog for real exponents of
est-descent integral. Although there is no general 2D method of
the method of stationary phase for imaginary exponents.
steepest descent 共in two complex variables兲, it is still possible to
In fact, we can identify those two directions of steepest descent at
evaluate this integral asymptotically. Doing so, however, is beyond
the saddle point by exploiting the approximation of the function dif-
the mathematical scope of this paper; we derive only the 1D formula,
ference in equation A-3. The minus sign in the exponent is already
which allows us to approximate equation 10 by equation 11.
built into the structure of the integrand in equation A-1, so the direc-
The 1D method of steepest descent is a standard tool in classical
tion we seek can be characterized by the complex angle or argument
mathematical physics. We begin by thinking of the original integra-
of the approximation in equation A-3. This direction maximizes the
tion over variable phx along the real axis as a more general contour in-
rate of decay of the exponential. This happens if the difference in
tegral in the complex phx plane. That is, we consider the integral on
equation A-3 is purely real so that its angle is equal to some multiple
an interval along the real axis, generically denoted I, as
of 2␲ . Because the angle of the product of two complex numbers is

I共␻ 兲 ⳱ 冕 dzf共z兲exp关ⳮ␻⌿ 共z兲兴. 共A-1兲


the sum of their angles,

arg关⌿ ⬙兴 Ⳮ 2 arg关z ⳮ zs兴 ⳱ 0,2␲ , . . . 共A-4兲


In our application, we must identify complex variable z ⳱ phx, com- which is equivalent to
plex amplitude
1
cos ␪ s cos ␪ L As*AL* arg关z ⳮ zs兴 ⳱ ⳮ arg关⌿ ⬙兴,
f⳱ Ds , 2
Vs VL psz pLz
and complex time ⌿ ⳱ iT* to use the asymptotic formula derived 1
here. However, the derivation is simplest if we start with the form arg关z ⳮ zs兴 ⳱ ⳮ arg关⌿ ⬙兴 Ⳮ ␲ , . . . , 共A-5兲
2
A-1.
We seek an asymptotic expansion of this integral for large ␻ . where arg means angle of a complex number relative to the positive
Technically, we should use a dimensionless large parameter — for real axis.
example, normalizing ␻ by some reference frequency 共different These are the only two unique choices. We expect that the direc-
from the reference frequency used in Gaussian-beam migration兲. tion of choice will be a rotation of the contour of integration — the
However, such a dimensionless parameter will be proportional to ␻ , real line, positively oriented — through an acute angle. Thus, from
defined as high frequency, so that we can proceed with this form, the two unique choices of direction, we choose
avoiding unnecessary notational clutter, knowing that the scale fac-
tors attached to ␻ will produce the desired dimensionless large pa- 1
arg关z ⳮ zs兴 ⳱ ⳮ Arg关⌿ ⬙兴, 共A-6兲
rameter. 2
True-amplitude Gaussian-beam migration S21

with Arg denoting the angle that picks out this principal argument of 1
the second derivative making an acute angle with the direction of the x共0兲 ⳱ x⬘, p共0兲 ⳱ 共sin ␪ ,cos ␪ 兲, ␶ 共0兲 ⳱ 0.
V共0兲
positive x-axis. Given this information, we can apply the method of
steepest descent 共Bleistein 1984, his equation 7.3.11兲 to the integral 共B-2兲
in equation A-1 to obtain the contribution to the asymptotic expan-
In addition, quantities PGB and QGB used in Gaussian-beam migra-
Downloaded 08/18/19 to 103.47.158.161. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

sion of I from the saddle point zs as follows:


tion satisfy the dynamic equations and initial conditions

冑 冋 册
dQGB dPGB 1 ⳵ 2V
2␲ Arg关⌿ ⬙兴 ⳱ V共s兲PGB共s兲, ⳱ⳮ 2 QGB共s兲,
I共␻ 兲 ⬇ f共zs兲exp ⳮ⌿ 共zs兲 ⳮ ds ds V 共s兲 ⳵ n2
␻ 兩 ⌿ ⬙兩 2


␻ rw20 i
2␲ QGB共0兲 ⳱ , PGB共0兲 ⳱ . 共B-3兲
⳱ f共zs兲exp关ⳮ⌿ 共zs兲兴, ␻ ⬎ 0. 共A-7兲 V共0兲, V共0兲
␻⌿ ⬙
We distinguish between the complex functions QGB and PGB for
Gaussian beams and the corresponding real solutions of asymptotic
That is, with the form of the original integral in equation A-1, the
ray theory, QART and PART. The latter satisfy the same system of equa-
phase shift in the exponent provides just the right factor to yield a
tions B-2, but the initial conditions change as follows:
complex square root of ⌿ ⬙, using as the principal argument the
choice that makes an acute angle with the direction of the original 1
contour of integration. If ␻ is negative, the formula will contain the QART共0兲 ⳱ 0, PART共0兲 ⳱ . 共B-4兲
V共0兲
appropriate phase shift to transform this expansion into its complex
conjugate. By comparing these initial conditions with those for QGB and PGB in
Equation A-7, with phx substituted for z, 共cos ␪ s /Vs兲共cos ␪ L /VL兲 equation B-3, we conclude that
⫻共As*AL* /psz pLz兲Ds substituted for f, and iT* substituted for ⌿ , yields
equation 11 as an approximate evaluation of equation 10. QART共s兲 ⳱ Im关QGB共s兲兴, PART ⳱ Im关PGB共s兲兴. 共B-5兲
We need this relationship in the following.

APPENDIX B Asymptotic expansion of the Gaussian-beam integral for a


single Green’s function
STEEPEST-DESCENT APPROXIMATION
In equation 5, the final integral expresses the Green’s function as
APPLIED TO GAUSSIAN BEAMS an integral of contributions from a set of individual beams indexed
by horizontal slowness px, with each term uGB representing a partial
In this appendix, we derive the expression in equation 12 for the
wavefield in a limited spatial region surrounding the beam’s central
second derivative T ⬙共p0h兲 in terms of the complex amplitudes Qs and
raypath. We begin our analysis by approximating this integral by the
QL of the two Gaussian-beam expansions appearing in the imaging
method of steepest descent. We later apply that analysis to an evalua-
formula for I共x;xs兲, equation 10. We need to start from an explicit
tion of T*⬙共phx
0
兲 in equation 12.
representation of a single Gaussian beam uGB as used in equation 5:
To approximate the Green’s function 共expression 5兲, we follow
the discussion at the beginning of Appendix A, setting

A⳱ 冑 V共s兲Q共0兲
V共0兲Q共s兲
, ⌿ 共px兲 ⳱ ⳮ iT共px兲. 共B-6兲
In addition to its dependence on slowness px, the function ⌿ depends
1 on location x. However, this dependence does not concern us here,
T ⬅ T共s,n兲 ⳱ ␶ 共s兲 Ⳮ PQⳮ1n2 . 共B-1兲 so all indicated derivatives of ⌿ and T* are with respect to px. As in
2
Appendix A, the saddle point is determined by setting
The quantities Q and P are determined along the central ray as solu- d⌿ dT
tions of a system of dynamic ray equations, with ray-centered coor- ⳱ ⳮi ⳱ 0. 共B-7兲
dpx dpx
dinates s 共measured along the central ray from its initial location at s
⳱ 0兲 and n 共measured perpendicular to the central ray at points s兲. For homogeneous media, one of the central rays passes through
Within a particular Gaussian beam, complex amplitude and time A subsurface location x; the saddle point occurs at this central ray. This
and T are functions of position; several Gaussian beams from a given is true because the imaginary part of T is zero for this central ray; that
source or receiver location might strike a particular subsurface loca- is, the imaginary part of T is a minimum and, correspondingly, the
tion, so A and T are also functions of slowness px. This dependence real part of T is a maximum for this critical slowness value, which we
of A and T on px interests us here. The kinematic ray equations and call px0. This is equally true for heterogeneous media, except in two
initial conditions 共at s ⳱ 0兲 for position x, time ␶ , and slowness vec- cases: 共1兲 when no central ray from source or receiver beam center
tor p ⳱ ⵜ ␶ are location x⬘ passes close to x 共a shadow zone兲 and 共2兲 when, in the dis-

冉 冊
crete sampling of central rays used in the sum approximating the in-
dx dp 1 d␶ 1 tegral, one or more central rays from x⬘ passes near the location x
⳱ V共s兲p, ⳱ ⵜ , ⳱ ,
ds ds V共s兲 ds V共s兲 without actually passing through it. Where an actual saddle point oc-
S22 Gray and Bleistein

curs, we can apply the steepest-descent formula of Appendix A to QART共s兲 QGB共0兲 Im关QGB共s兲兴 QGB共0兲
obtain the leading order asymptotic expansion of the Gaussian-beam T ⬙共p0x 兲 ⳱ ⳱
integral:
QGB共s兲 pz V共0兲
2
QGB共s兲 pz2V共0兲

A exp关ⳮ␻⌿ 共p0x 兲兴 A exp关i␻ T共p0x 兲兴 Im关QGB共s兲兴 ␻ rw20


G共x;x⬘ ; ␻ 兲 ⬇ ⳱ ⳱ , 共B-11兲
2pz冑2␲␻⌿ ⬙共p0x 兲 2pz冑ⳮ2␲ i␻ T⬙共p0x 兲
QGB共s兲 pz2V2共0兲
Downloaded 08/18/19 to 103.47.158.161. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

where pz ⳱ 冑1 ⳮ V2共0兲px2 /V共0兲. The first equality in equation B-11


A⳱ 冑 V共s兲Q共0兲
V共0兲Q共s兲
. 共B-8兲 arises from equating the amplitudes and solving for T ⬙共px0兲. The sec-
ond equality exploits the derived relationship between QART and QGB
in equation B-5. The final equality uses the prescribed initial data of
This asymptotic approximation is valid only if T ⬙共px0兲 is nonzero; equation B-3.
T ⬙共px0兲 will be zero when x is at a caustic of the family of central rays. Substituting the final value of T ⬙共px0兲 from equation B-11 into the
Equation B-8 approximates the superposition of multiple contri- denominator of equation B-8 for the asymptotic expansion of the
butions to the wavefield by a single term; this seems to negate the ad- Green’s function provides the right asymptotic expansion away
vantage of the Gaussian-beam expansion of the Green’s function, from caustics of central rays. A convenient cancellation of QGB共s兲
which is multiple contributions to the wavefield from a range of ini- occurs, leading to the same asymptotic leading term as is stated for
tial directions. In our downward continuation of wavefields from the GART in equation B-10. This is not our main objective here, however,
source and receiver locations, we use the approximation in a manner so we do not carry out the details. We need identity B-11 for T ⬙共px0兲
that incorporates most of the partial contributions. Thus, as Hill following.
共2001兲 shows, Gaussian-beam migration retains most of the arrivals,
even though it relies on the steepest-descent approximation.
As mentioned, saddle points do not exist in a shadow zone. In this
Multiple Gaussian-beam integrations for the product of
case, no contribution to the wavefield 共and therefore no steepest-de- Green’s functions
scent calculation兲 is performed because either the source or beam- The imaging condition of equation 10 involves a product of com-
center location does not see the subsurface location. True saddle plex conjugates of Green’s functions. Thus, creating an image of the
points also do not exist when one or more central rays from a source subsurface requires two additional integrations to obtain the migrat-
or receiver location pass near the subsurface location x but none ac- ed amplitude at an image location, compared to the computational
tually pass through it, as in Hill’s 共2001兲 Figure 2. In this case, we requirements of a standard Kirchhoff true-amplitude migration.
may still seek the dominant contribution to the integral and include it Here, following an approach proposed by Hill 共2001兲, we carry out
in the wavefield at x. As in the steepest descent calculation, the dom- the asymptotic expansion of the integral over phx by the method of
inant contribution comes from the central ray whose imaginary time steepest descent. In addition to producing Hill’s complex traveltime,
is a minimum. If this value of imaginary time is close to zero, then we also produce the amplitude correction from the steepest-descent
the central ray is very close to a stationary value and the steepest-de- calculation.
scent calculation can be performed with negligible error. If the mini- Let us define


mum value of imaginary time is not close to zero, the central ray is
not close to a stationary value and the steepest descent calculation is
in error. However, the large value of imaginary time will supply a J共x;xs兲 ⳱ dphxAs*AL* exp关ⳮi␻ T*共phx兲兴Ds共L,prx, ␻ 兲
large exponential decay to the wavefield contribution from this cen-
tral ray, and both the contribution to the wavefield and the error will 共B-12兲
be negligible. In such quasi-saddle point cases, we can perform the with T*共phx兲 ⳱ Ts* Ⳮ TL*. Here, J is the innermost integral 共for fixed
steepest-descent integral as if the saddle point were a true one. pmx兲 in equation 10, and we have neglected the dependence of the
At a true saddle point identified by the critical slowness value px0, functions on all variables except the variable of integration phx. In
the complex traveltime becomes real: particular, we have neglected the dependence on the other variable
of integration pmx. The asymptotic expansion that we derive here
T共p0x 兲 ⳱ ␶ 共s兲, 共B-9兲 must be applied for each pmx.
in agreement with the traveltime obtained by asymptotic ray theory. We propose to apply the general formula of equation A-7 for ap-
In fact, the asymptotic expansion of the Green’s function in equation proximating the integral J in equation B-12, with ⌿ ⳱ iT * 共phx兲. As
B-8 should agree, at least to leading order, with the expansion ob- happened for the calculation of the asymptotic Green’s function
tained by asymptotic ray theory. The leading order term of that B-11, there may be true simple saddle points 共for which the method
asymptotic expansion is of steepest descent is completely valid兲, quasi-saddle points, or

冋 册
shadow zones. If x is in a shadow zone of either xs or L, no integra-

exp i␻ ␶ 共s兲 Ⳮ i 冉冊␲


4
sgn共␻ 兲
tion will take place and beam center L will not contribute to the im-
age at x. If one or more central rays from both xs and L pass near x


G共x;x⬘ ; ␻ 兲 ⬇ . 共B-10兲 共with no central ray pairs actually striking x兲, there will be a quasi-
2␲ 兩␻ 兩QART共s兲 saddle point. Then we can use the same argument as above to show
2 that the error in using the steepest-descent method to evaluate inte-
V共s兲
gral J is negligible compared with the error in the method itself. As in
When we compare this asymptotic expression for the Green’s func- the evaluation of the Green’s function, we are justified in using the
tion with the one in equation B-8, we obtain an indirect evaluation of steepest-descent method for both true saddle points and quasi-saddle
T⬙共px0兲, namely, 0
points. Doing this yields a critical value phx for which
True-amplitude Gaussian-beam migration S23

J共x;xs兲 ⬇ 冑 2␲
0 As AL Ds共L,prx, ␻ 兲
兩␻ 兩T* ⬙共phx 兲
* *
REFERENCES

冋 册
Albertin, U., D. Yingst, P. Kitchenside, and V. Tcheverda, 2004, True-ampli-
tude beam migration: 74th Annual International Meeting, SEG, Expanded
␲ Abstracts, 398–401.
⫻ exp ⳮ i␻ T *共phx
0
兲Ⳮi sgn共␻ 兲 . Beylkin, G., 1985, Imaging of discontinuities in the inverse scattering prob-
4
Downloaded 08/18/19 to 103.47.158.161. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

lem by inversion of a causal generalized Radon transform: Journal of


Mathematical Physics, 26, 99–108.
共B-13兲 Bleistein, N., 1984, Mathematical methods for wave phenomena: Academic
Press Inc.
Recall that pmx is fixed during the evaluation of integral J. As a result, ——–, 1987, On the imaging of reflectors in the earth: Geophysics, 52,
0
the critical value phx depends on pmx, and critical values of psx and prx 931–942
depend on pmx and phx 0
. Bleistein, N., Y. Zhang, S. Xu, S. H. Gray, and G. Zhang, 2005, Migration/in-
version: Think image point coordinates, process in acquisition surface co-
Finally, from equation B-12, we can write T* ⬙共phx 0
兲 ⳱ T* ⬙共psx兲 ordinates: Inverse Problems, 21, 1715–1744.
Ⳮ T* ⬙共prx兲. Using equations 9 and B-11 to evaluate the individual Deng, F., and G. McMechan, 2007, True-amplitude prestack depth migra-
tion: Geophysics, 72, no.3, S155–S166.
derivatives, we obtain Gray, S. H., 2005, Gaussian beam migration of common-shot records: Geo-

冋 册
physics, 70, no.4, S71–S77.
1 Im共Qs兲 1 Im共QL兲 ——–, 2007, Angle gathers for Gaussian beam migration: 69th Annual Con-
T *⬙共phx
0
兲 ⳱ ⳮ␻ rw20 Ⳮ 2 2 . ference and Exhibition, EAGE, Extended Abstracts.
Vs2 psz
2
Qs Vr prz QL Hanitzsch, C., 1997, Comparison of weights in prestack amplitude-preserv-
ing Kirchhoff depth migration: Geophysics, 62, 1812–1816.
共B-14兲 Hill, N. R., 1990, Gaussian beam migration: Geophysics, 55, 1416–1428.
——–, 2001, Prestack Gaussian-beam depth migration: Geophysics, 66,
Equation B-14 evaluates T*⬙ in the asymptotic expression for J in 1240–1250.
equation B-12. We then substitute this expression into equation 11 to Keho, T. H., and W. B. Beydoun, 1988, Paraxial ray Kirchhoff migration:
Geophysics, 53, 1540–1546.
reduce the double integral over pmx and phx into a single integral over Miller, D., M. Oristaglio, and G. Beylkin, 1987, Anew slant on seismic imag-
pmx, which is equation 12. ing: Migration and integral geometry: Geophysics, 52, 943–964.
To complete the discussion, we must mention that the incident or Nowack, R. L., M. K. Sen, and P. L. Stoffa, 2003, Gaussian beam migration
for sparse common-shot and common-receiver data: 73rd Annual Interna-
reflected wavefield might have a caustic of central rays at some im- tional Meeting, SEG, Expanded Abstracts, 1114–1117.
age locations. Even though the regular behavior of the critical pair共s兲 Paffenholz, J., 2001, Sigsbee2 synthetic data set: Image quality as function of
migration algorithm and velocity model error: 71st Annual International
of Gaussian beams will produce nonzero amplitudes at such loca- Meeting, SEG, Workshop W-5.
tions, the interpretation of the peak amplitude at those image loca- Xu, S., H. Chauris, G. Lambaré, and M. Noble, 2001, Common-angle migra-
tions in terms of the geometric optics reflection coefficient will not tion: A strategy for imaging complex media: Geophysics, 66, 1877–1894.
Zhang, Y., S. Xu, N. Bleistein, and G. Zhang, 2007, True-amplitude, angle-
be valid. Briefly, such locations will produce nonsimple saddle domain, common-image gathers from one-way wave-equation migra-
points, violating the validity of the steepest-descent analysis. tions: Geophysics, 72, no.1, S49–S58.

Das könnte Ihnen auch gefallen