+ All Categories
Home > Documents > Chapter 1 Downward-continuation methods

Chapter 1 Downward-continuation methods

Date post: 28-Dec-2021
Category:
Upload: others
View: 4 times
Download: 0 times
Share this document with a friend
65
Chapter 1 Downward-continuation methods The numerical solution of the one-way wave equation is the cornerstone of all downward-continuation migration methods. Over the years a wide variety of so- lutions have been proposed based on approximations of the SSR operator intro- 1
Transcript
Page 1: Chapter 1 Downward-continuation methods

Chapter 1

Downward-continuation methods

The numerical solution of the one-way wave equation is the cornerstone of alldownward-continuation migration methods. Over the years a wide variety of so-lutions have been proposed based on approximations of the SSR operator intro-

1

Page 2: Chapter 1 Downward-continuation methods

duced in Chapter??. There is no absolute optimum among these methods, sincethe particular problem at hand determines the ideal balance of accuracy, computa-tional cost, flexibility, and robustness in the selection of a downward-continuationmethod. This chapter provides a fairly complete overview of the numerical methodsthat have been developed for downward-continuing wavefields.

Downward-continuation methods can be classified according to the computa-tional domain in which the SSR operator is applied. The most efficient, but alsothe least flexible, are the migrations performing the computations in thefrequency-wavenumber domain (ω-k). Those purely in thefrequency-space domain (ω-x)are at the other extreme of the spectrum; that is, they are the most flexible but tend tobe also the most computationally intensive. Between these extremes, there are themixed-domain (ω-k/ω-x) methods. We start our discussion with theω-k methods,and then quickly move on to the more usefulω-k/ω-x methods.

Page 3: Chapter 1 Downward-continuation methods

1.1 Frequency-wavenumber (ω-k) domain meth-odsWhen the propagation velocity is assumed to be a function only of depth, wavefieldscan be downward-continued efficiently in the frequency-wavenumber (ω-k) domainby simple multiplication with the SSR downward-continuation operator introducedin Section??:

Pz+1z (ω,k) = Pz (ω,k)ei 1z

√ω2

v(z)2−|k|

2

, (1.1)

whereω is the temporal frequency,k is the vector of the horizontal wavenumbers,andv (z) is the depth-dependent velocity.

The downward-continuation operator expressed in equation (1.1) describes adip-dependent(k/ω) phase shift of the wavefield, and thus it is often calledphase-shift operator (Gazdag, 1978). When the medium velocity is strictly a functionof depth, the application of equation (1.1) downward-continues the data withoutany approximation of the kinematics of waves propagating along one directionwith respect to the vertical axis. (Depending on the sign of the exponential thewaves can propagate either upward or downward.) However, the effects of veloc-

Page 4: Chapter 1 Downward-continuation methods

ity variations along the depth direction on geometrical spreading are not correctlytaken into account by the phase-shift operator, and consequently the amplitudes ofthe downward-continued wavefield are not correct. In Chapter?? we will discussin detail the issues that influence the migrated amplitudes when using wavefield-continuation methods. In that chapter we will also discuss correction terms thatcan be introduced to correct for the mishandling of geometrical spreading by thephase-shift operator.

The remainder of this chapter presents a wide-range of methods to solve theSSR equation when lateral velocity variations are not negligible, and thus the accu-racy of the phase-shift operator is unacceptable. To adapt the phase-shift operatorto lateral velocity variations it is necessary to perform at least part of the computa-tion in the space domain. Several methods perform part of the computation in thewavenumber domain and part in the space domain; these methods can be catego-rized as mixedω-k/ω-x-domain methods.

Page 5: Chapter 1 Downward-continuation methods

1.3 Frequency-space (ω-x) methodsPure space-domain (ω-x) methods are preferable to mixed-domain methods whenthe velocity variations are so strong that too many reference velocities are requiredfor accurate mixed-domain migrations. Theω-x domain methods are based on theapproximation of the SSR operator with a finite-length convolutional filter in thespace domain. Becauseω-x methods are applied as convolutions in the space do-main, the coefficients of the depth-extrapolation filter can vary as a function of themidpoint axes. Therefore,ω-x methods can easily accommodate strong velocityvariations without any additional computational cost and complexity. However, be-cause the midpoint-wavenumber vectork is two-dimensional, a direct implemen-tation of equation (??) would lead to a two-dimensional convolutional filter. Two-dimensional filters are often cumbersome to design, and they are expensive to apply,because their computational cost grows with the square of the filter length. There-fore, alternative methods for implementing equation (??) in the space domain arenecessary to makeω-x migrations competitive withω-k/ω-x migrations.

Section1.3.1presents a simple splitting method to approximate the two-dimensionalconvolution with a cascade of two one-dimensional convolutions. Splitting is sim-ple and efficient, but it introduces errors that are unacceptable in many practical

Page 6: Chapter 1 Downward-continuation methods

cases. Section1.3.2introduces methods that have been recently introduced to avoidsplitting and reduce the large computational cost incurred by full 2-D convolutionalfilters.

1.3.1. Splitting methods for 3-D downward continuationThe challenge withω-x methods is that a straightforward representation of equa-tion (??) as a 2-D convolutional filter along the midpoint axes results in a fairlyexpensive algorithm. The cost function for such an algorithm is the following:

ZoffDwn(ω−X) ∝ κZoffDwn(ω−X) ×(Nzξ ×Nxξ ×Nyξ

(Nt×Nop2

). (1.17)

The dependency of the cost function on the square of the filter length Nop makes themethod unattractive, hence the desire to split the 2-D convolution into the cascadeof two 1-D convolutions. With splitting, the cost function is reduced to a moreattractive functional:

ZoffDwnSpl(ω−X) ∝ κZoffDwnSpl(ω−X)×(Nzξ ×Nxξ ×Nyξ

)× (Nt×2×Nop).

(1.18)

Page 7: Chapter 1 Downward-continuation methods

Unfortunately, the 2-D convolution expressed in equation (??) does not fullyseparate into the cascade of two 1-D convolutions along the midpoint axes, althoughequation (??) is very similar to the expression for the summation surfaces of zero-offset Kirchhoff migration in equation (??). When operators are represented assummation surfaces, cascading an in-line operator with a cross-line operator is ex-pressed as a substitution of variables; while, cascading convolutional filters in thewavenumber domain is represented as a multiplication of the corresponding trans-fer functions. To split the downward-continuation operator, we need to rewrite theSSR dispersion relation as a sum of terms that are functions of eitherkxm or kym. Anexact decomposition of the square root in this fashion is impossible, and thus theSSR equation can be only approximated as a cascade of 1-D convolutional filters.The SSR equation is usually approximated as follows:

SSR(ω,k) ≈

√4ω2

v (z,x, y)2−k2

xm︸ ︷︷ ︸Convolution inx

+

√4ω2

v (z,x, y)2−k2

ym︸ ︷︷ ︸Convolution iny

−2ω

v (z,x, y). (1.19)

Splitting is exact for dips that are aligned along either of the midpoint axes, for

Page 8: Chapter 1 Downward-continuation methods

which eitherkxm or kym are equal to zero. The error introduced by the approximationin equation (1.19) is maximum for reflections oriented at 45◦ with respect to themidpoint axes and steeply dipping with respect to the surface.

Figure1.17and Figure1.18compare depth slices of the migration impulse re-sponses generated with the exact SSR equation (left) and its splitting approximation(right). Figure1.17corresponds to a a reflector dip of 30◦. It shows that splittingachieves a good approximation of the desired circular response when the reflec-tor dip is moderate. In contrast, Figure1.18corresponds to a reflector dip of 60◦;the results obtained by splitting show an unacceptable anisotropy of the impulseresponse.

Figure1.19 and Figure1.20 show slices extracted from migrated cubes of areal data set collected in the Gulf of Mexico over a salt dome. Figure1.19comparestwo diagonal slices obtained by splitting (left) and by a full downward-continuationoperator (right). The two results show several significant differences both in thequality of the reflectors’ focusing and in their positioning. The dipping reflectorsclose to the salt-sediment boundaries are the ones that are most negatively affectedby splitting. Figure1.20 compares two depth slices through the migrated cubesobtained by splitting (left) and by a full downward-continuation operator (right).

Page 9: Chapter 1 Downward-continuation methods

The most evident differences between the two sections is the mispositioning causedby splitting of the broad reflector at CMP X=12 km.

The anisotropy caused by splitting can be compensated for with correction fil-ters. Li (1991) first proposed a method based on a correction filter. However, othermethods have been recently developed that approximate the SSR equation more ac-curately than do splitting methods; an example is the McClellan method describedin Section1.3 (Hale, 1991). The computational complexity of these new methodsgrows only linearly with the filter length, as does the cost of splitting methods, butthe new methods have more isotropic impulse responses.

1.3.2. McClellan downward continuationThis section presents a method to approximate the two-dimensional convolution thatis more expensive than simple splitting, but for which the computational cost stillgrows linearly with the filter length, as it does for splitting. To reduce the com-putational cost and to simplify the design of the continuation operator, the methodexploits the circular symmetry of the downward-continuation operator. The symme-try of the continuation operator in the space domain directly derives from circular

Page 10: Chapter 1 Downward-continuation methods

Figure 1.17: Depth slices extracted from zero-offset migration impulse responsescorresponding to a reflector dip of 30◦, obtained with the exact operator (left) andsplitting (right). down-dslice-30-overn[CR,M]

Page 11: Chapter 1 Downward-continuation methods

Figure 1.18: Depth slices extracted from zero-offset migration impulse responsescorresponding to a reflector dip of 60◦, obtained with the exact operator (left) andsplitting (right). down-dslice-60-overn[CR,M]

Page 12: Chapter 1 Downward-continuation methods

symmetry of the SSR as a function of the midpoint wavenumbers. SSR is circularlysymmetric in the wavenumber domain because it is a one-dimensional function ofthe midpoint-wavenumber vector magnitude|k|. Therefore, it can be approximatedby the following finite-length summation:

exp[SSR(ω,k)] ≈

Nop−1∑n=−Nop+1

hn (ω/v)exp[i (n |k|)] . (1.20)

The circular symmetry of the operator guarantees that the coefficientshn(ω/v) areeven; that ishn(ω/v) = h−n(ω/v). Therefore, the summation in (1.20) can berewritten as the following summation of only cosine terms:

exp[SSR(ω,k)] ≈ h0 (ω/v)+Nop−1∑

n=1

hn (ω/v)cos(n |k|) . (1.21)

Notice that the previous expression would be exactly the same if we were consider-ing a 2-D downward continuation filter, wherekxm is substituted for|k|. Thereforethedepth-extrapolation filter coefficientshn are function only of the ratioω/v and

Page 13: Chapter 1 Downward-continuation methods

can be designed with any of the methods proposed for designing depth-extrapolationfilters in 2-D.Nautiyal et al.(1993) present an intuitive method for 2-D filter design.

From a practical viewpoint, it is convenient that the coefficients of the depth-extrapolation filter are a function of the ratioω/v and not of both variables indepen-dently. Tables containing the filter coefficients are thus one-dimensional; they areeasy to precompute and can be stored in fast memory without difficulties.

Although equation (1.21) leads to a simpler filter design, its direct applicationwould still require the convolution with the circularly symmetric 2-D filters repre-sented in the wavenumber domain by cos(n|k|). These filters are expensive to applybecause their length grows linearly withn. Fortunately, the McClellan transforms(McClellan and Chan, 1977) can be used to reduce drastically the cost of applyingequation (1.21), as first presented by Hale (1991). The McClellan transforms arebased on the following Chebyshev recursion formula for the cosine function:

cos(n |k|) = 2cos[(n−1)|k|] cos|k|−cos[(n−2)|k|]. (1.22)

Using this recursion formula, cos(n|k|) can be expressed, for any integern, in termsof annth-order polynomial of cos|k|. Using the recursion of equation (1.22), equa-

Page 14: Chapter 1 Downward-continuation methods

tion (1.21) can be written in terms of powers of cos|k| as follows:

exp[SSR(ω,k)] ≈

Nop−1∑n=0

hn (ω/v)cos(n |k|)

=

Nop−1∑n=0

bn (ω/v)cosn |k| , (1.23)

where the coefficientsbn can be obtained from the coefficientshn. Therefore, byusing the McClellan transforms, we can substitute recursive convolutions with thecompact cos|k| filter for the convolution with the long cos(n|k|) filters.

An efficient implementation of the McClellan transforms fully exploits the re-cursive nature of the Chebyshev formula, and requires only Nop− 1 applicationsof the cos|k| filter. Figure1.21 shows a graphical representation of an efficientimplementation of the McClellan transforms, where the filterG is a space-domainapproximation of cos|k|.

The simplest, though least accurate, approximation for cos|k| is expressed in

Page 15: Chapter 1 Downward-continuation methods

the wavenumber domain as follows:

cos|k| ≈ G9pts(kxm,kym

)= −1+

1

2

(1+coskxm

)(1+coskym

). (1.24)

In the space domain, the filter represented by equation (1.24) is the following 3×3convolutional filter:

G9pts=

1/8 1/4 1/8

1/4 -1/2 1/41/8 1/4 1/8

. (1.25)

Hale (1991) also suggested a more accurate approximation of the cos|k| that isonly marginally more expensive. This improved cosine filter is represented in thewavenumber domain as

cos|k| ≈ G17pts(kxm,kym

)= G9pts−

c

2

[1−cos

(2kxm

)][1−cos

(2kym

)], (1.26)

where c is chosen by exactly matchingG17pts to cos|k| at a particular value of|k|

along the diagonalkxm = kym. WhenG17pts is matched at|k| = π/3, c is equal to0.0255. The space representation of theG17pts has 17 coefficients different from

Page 16: Chapter 1 Downward-continuation methods

zero:

G17pts=

-c/8 0 c/4 0 -c/8

0 1/8 1/4 1/8 0c/4 1/4 -(1+c)/2 1/4 c/40 1/8 1/4 1/8 0

-c/8 0 c/4 0 -c/8

. (1.27)

Both the filters in1.25 and1.27 have a quadrantal symmetry that can be furtherexploited to reduce the number of complex multiplications necessary to convolveGwith the wavefield, and thus reducing the computational cost of the method.

Both approximations of theG filter cause errors in the wavefield propagation.These errors mostly cause a slight anisotropy of the effective propagation opera-tor, with obviously theG9pts filter being more anisotropic than theG17pts filter.Biondi and Palacharla(1994) proposed an effective method to reduce this opera-tor anisotropy by alternating the usage ofG9pts with the usage of a filter similar toG9pts, but rotated by 45 degrees with respect to the horizontal axes. The additionalaccuracy is gained at the expenses of a negligible increase in computational cost.

The cost function of McClellan migrations resembles the cost function of split-

Page 17: Chapter 1 Downward-continuation methods

ting [equation (1.18)] because it grows only linearly with the number Nop of coef-ficients in the depth-extrapolation filter; that is,

ZoffDwn(ω−X) ∝ κZoffDwn(ω−X) ×(Nzξ ×Nxξ ×Nyξ ×Nop

)× (Nt) . (1.28)

However, the leading constantκZoffDwn(ω−X) is considerably higher for McClellanmigration than for splitting, because of the convolutions with the filterG.

The main disadvantage of the application of the McClellan transform to down-ward continuation is the difficulty of generalizing the method to the practically im-portant case where the samplings along the two horizontal directions are different(1x 6=

1y). Soubaras(1994) presented a method that, like Hale’s method, takes advantageof the circular symmetry of the 3-D downward-propagation operator, but is moreflexible with respect to spatial sampling.

Figure1.22 shows two cross-line sections extracted from an image cube ob-tained by McClellan migration. The section on the left was obtained with lessaccurate, cheaper filters than those used for the section on the right. For the lessexpensive solution, Nop was equal to 20 andG = G9pts, whereas Nop was equal to40 andG = G17pts for the more expensive solution. The image on the left was ob-tained in a third of the computer time required to compute the image on the right. No

Page 18: Chapter 1 Downward-continuation methods

Figure 1.19: Diagonal slices extracted from zero-offset migrations. The sliceon the left was obtained with splitting, while the slice on the right was ob-tained by use of a full downward-continuation operator. Notice the differencesin focusing and positioning of the sediments’ terminations around the salt dome.down-diag-salt-overn[CR,M]

Page 19: Chapter 1 Downward-continuation methods

Figure 1.20: Depth slices extracted from zero-offset migrations. The slice onthe left was obtained with splitting, while the slice on the right was obtained byuse of a full downward-continuation operator. Notice the different positioningof the reflector at CMP X=12 km, and the different phases of many reflectors.down-dslice-salt-overn[CR,M]

Page 20: Chapter 1 Downward-continuation methods

2G 2Goo oo

V

o

G 2Go o

Vh0

o

Vh1

o

o

2

o 2G

h2 2V

o

o o

h23V

o

o

VV

o

−−

h2h2h2N−1N−2N−1

oO/P

I/P −

o

Figure 1.21: Efficient implementation of the recursive algorithm that implementsthe McClellan transforms.down-cheby[NR]

Page 21: Chapter 1 Downward-continuation methods

Figure 1.22: Cross-line sections obtained by McClellan migration, withG = G9pts and Nop=20 (left), and withG = G17pts and Nop=40 (right).down-xslice-mc-overn[CR,M]

Page 22: Chapter 1 Downward-continuation methods

Figure 1.23: Depth slices obtained by McClellan migration, withG = G9pts andNop=20 (left), and withG = G17pts and Nop=40 (right). down-dslice-mc-overn[CR,M]

Page 23: Chapter 1 Downward-continuation methods

1.2 Mixed frequency-wavenumber/space (ω-k/ω-x) methodsGazdag and Sguazzero(1984) proposed the first method to generalize the phase-shift method to lateral velocity variations. Because their method is based on phaseshift and interpolation, they called itPhase Shift Plus Interpolation (PSPI) mi-gration. Their method can be summarized as follows: 1) propagate the wave-field at each depth step by phase shift using several (two or more) constant veloci-ties (reference velocities), 2) transform the wavefields obtained by these constant-velocity continuations into the space domain, and 3) interpolate between the wave-fields according to the differences between the medium velocity and the referencevelocities. PSPI is still a popular method because of its conceptual simplicity andits flexibility. Up to a limit, the accuracy of PSPI downward-continuation can beincreased at will by simply increasing the number of reference velocities.

I will not discuss PSPI migration in any further detail, but in the next section IpresentSplit-step migration that is closely related to PSPI.

Page 24: Chapter 1 Downward-continuation methods

1.2.1. Split-step migrationSplit-step migration (Stoffa et al., 1990) retains PSPI’s characteristics of simplicityand flexibility but has the conceptual advantage that it is easily related to the moreaccurate methods that are introduced in Section1.2.2. In light of the theory intro-duced in Section1.2.2, the application of a split-step correction can be interpreted asthe application of of a first-order (with respect to velocity perturbations) correctionoperator to the phase-shift operator, whereas the more accurate methods presentedin Section1.2.2apply a higher-order correction operator.

Split-step downward continuation is based on the following approximation ofthe SSR equation:

kz = SSR(ω,k) ≈

(√ω2

v2ref

−|k|2

)︸ ︷︷ ︸

ω−k domain

+

v (z,x, y)−

ω

vref

),︸ ︷︷ ︸

ω−x domain

(1.2)

wherevref is called the reference velocity. The reference velocity is usually setequal to the average velocity in the current extrapolation layer. As indicated by theunderlining braces, the operator implementing the first term in expression (1.2) is

Page 25: Chapter 1 Downward-continuation methods

applied in theω-k domain, whereas the second term represents a spatially varyingtime-shift that can be easily applied in theω-x domain.

Figure1.1 shows the flow-chart of the inner kernel of the split-step algorithm.At every depth step, the wavefield is first transformed into the wavenumber domainby a forward FFT. The transformed wavefield is then downward-propagated withthe SSR operator, assuming the velocity to be constant and equal to the referencevelocity. The subsequent, spatially varying time-shift, corrects, at least to first order,for the discrepancies between the reference velocity and the actual medium velocity.

Performing the downward propagation partially in the wavenumber domain andpartially in the space domain requires the computation of at least a forward and aninverse Fourier transform at each depth step. Because the cost of an FFT of lengthN grows asN log2 N, the cost function of split-step migration is equal to

ZoffDwn(ω−K/ω−X) ∝ κZoffDwn(ω−K/ω−X) ×(Nzξ ×Nxξ ×Nyξ × log2Nxξ × log2Nyξ

)×(Nt) .

(1.3)This cost function is quite attractive; typically Nxξ and Nyξ are of the order of thethousands, and thus log2Nxξ and log2Nyξ are approximately equal to 10.

The split-step approximation in equation (1.2) is exact when either the velocity

Page 26: Chapter 1 Downward-continuation methods

Figure 1.1: Flow-chart of theinner kernel of the split-stepalgorithm. At every depthstep, the wavefield is first trans-formed into the wavenumber do-main by a forward FFT. Thetransformed wavefield is thendownward propagated with theSSR operator, assuming the con-stant velocityvref. The follow-ing spatially varying time-shiftcorrects, at least to first order,for the discrepancies between thereference velocityvref and theactual medium velocityv(m,z).down-split-cons[NR]

ωv

2

2ref

k2

ωvrefv(z,x,y)

ω

Inverse FFT

Forward FFT

Page 27: Chapter 1 Downward-continuation methods

is constant and equal to the reference velocity, or the midpoint wavenumber is equalto zero. The steeper the reflections and the stronger the lateral variations in velocity,the greater is the error introduced by the split-step operator. Section1.2.2showsexamples of impulse responses that illustrate this dependency of the errors from thereflectors dips and the velocity contrast.

The accuracy of split-step migration can be improved by using more than onereference velocity. In this modified scheme, often calledExtended Split-Step mi-gration (Kessinger, 1992), as many reference wavefields are generated as referencevelocities are used. A single wavefield is then estimated at each depth step by an in-terpolation in the space domain of the reference wavefields corresponding to all ref-erence velocities. At each spatial location, the interpolation weights are computedaccording to the difference between the actual medium velocity and the respectivereference velocity.

Figure1.2 shows the flow-chart of the extended split-step method, which usesmultiple reference velocities. Only one forward FFT is required at each depth step,but as many inverse FFTs are required as references velocities are used. Thereforethe cost function of the simple split-step method expressed in equation (1.3) must bemultiplied by the number of reference velocities. If the velocity variations are large,

Page 28: Chapter 1 Downward-continuation methods

Figure 1.2: Flow-chart of the ex-tended split-step method, whichuses multiple reference veloci-ties. The wavefield is downwardcontinued separately for eachreference velocityvref1,vref2, ...,using the split-step approxima-tion. At the end of each depthstep, the resulting wavefieldsare combined by interpolation.Only one forward FFT is re-quired at each depth step, but asmany inverse FFTs are requiredas reference velocities are used.down-split-var [NR]

ωv

2

2ref

k2

2

ωv

2

2ref

k2

1

1

ωvrefv(z,x,y)

ω2

ωvrefv(z,x,y)

ω

Inverse FFT

Interpolation

Inverse FFT

Forward FFT

Page 29: Chapter 1 Downward-continuation methods

several reference velocities are needed to achieve accurate results. Consequently,when the velocity variations are large, the cost of the mixed-domain method may betoo high compared to the pure space-domain methods presented in Section1.3.

Figures1.3and1.4show the results of using split-step to migrate a zero-offsetdata set recorded over a salt dome in the Gulf of Mexico. This example illustratesthe need for multiple reference velocities when split-step migration is used with arapidly changing medium velocity. The sections on the left are extracted from animage cube that was migrated with one reference velocity, whereas the sections onthe right were obtained with two reference velocities. There are visible differencesbetween the cross-line sections shown in Figure1.3. The use of just one referencevelocity causes the artifacts on the salt flank, where the migration velocity rapidlychanges because of the salt-sediment transition. The mispositioned events are alsovisible in the depth slice shown on the left in Figure1.4. These spurious events over-lap with the sediments’ reflections and distort the image of the salt flanks, makingthe interpretation of the salt boundaries more difficult.

The real data shown in the previous figure require two reference velocities tobe migrated satisfactorily, because the sediment velocity has small lateral variationsand the salt dome has a simple shape. In contrast, the proper zero-offset migra-

Page 30: Chapter 1 Downward-continuation methods

Figure 1.3: Cross-line sections obtained by split-step migration with only one ref-erence velocity (a) and with two reference velocities (b).down-xslice-split-overn[CR,M]

Page 31: Chapter 1 Downward-continuation methods

Figure 1.4: Depth slices obtained by split-step migration with only one referencevelocity (a) and with two reference velocities (b).down-dslice-split-overn[CR,M]

Page 32: Chapter 1 Downward-continuation methods

tion of the SEG-EAGE salt data set requires many more reference velocities. Fig-ures1.5and1.6compare the in-line section migrated with three reference velocitiesand with six reference velocities. The top of the salt and the salt flanks are betterfocused when six reference velocities are used. Figures1.7 and1.8 compare thecross-line section migrated with three reference velocities and with six referencevelocities. The bottom of the salt and the flanks of the canyons in the middle partof the salt body are better focused when six reference velocities are used. However,to obtain the better images, the computational cost was doubled. In the next sec-tion we discuss a family of methods that can be used to improve the accuracy ofmixed-domain downward continuation more cost effectively than by increasing thenumber of reference velocities in an extended split-step downward continuation.

1.2.2. Higher-order mixedω-k/ω-x methodsThe previous examples show that when the lateral velocity variations are strong andcontinuous – that is the velocity function cannot be simply represented by a fewdiscrete values – the extended split-step method would require several referencevelocities to produce good images. In those cases, it is advantageous to use a more

Page 33: Chapter 1 Downward-continuation methods

Figure 1.5: In-line section (CMP Y = 8,010 m) of the zero-offset SEG-EAGE saltdata set migrated with three reference velocities.down-Salt-3v-x8010[CR,M]

Page 34: Chapter 1 Downward-continuation methods

Figure 1.6: In-line section (CMP Y = 8,010 m) of the zero-offset SEG-EAGE saltdata set migrated with six reference velocities.down-Salt-6v-x8010[CR,M]

Page 35: Chapter 1 Downward-continuation methods

Figure 1.7: Cross-line section (CMP X = 7,440) of the zero-offset SEG-EAGE saltdata set migrated with three reference velocities.down-Salt-3v-y7440[CR,M]

Page 36: Chapter 1 Downward-continuation methods

Figure 1.8: Cross-line section (CMP X = 7,440) of the zero-offset SEG-EAGE saltdata set migrated with six reference velocities.down-Salt-6v-y7440[CR,M]

Page 37: Chapter 1 Downward-continuation methods

accurate approximation to the SSR operator than the split-step approximation shownin equation (1.2). These considerations have led to the developments of severalmethods that approximate the SSR by adding additional terms, of higher order inthe spatial derivatives (k), to the split-step approximation. All these methods reduceto the split-step method for vertically propagating waves – that is whenk vanishes –but propagate dipping events with higher accuracy. The most accurate and efficientmethod of this group is theFourier Finite-Difference (FFD) migration introducedby Ristow and Rühl(1994). The extended local Born-Fourier migration (Huanget al., 1999) and pseudo-screen propagator methods (Xie and Wu, 1999) are lessaccurate, but of more straightforward derivation, and thus I will present them first.

• Pseudo-screen methodThe basic idea of all these methods is fairly simple, and is based on a Taylor ex-pansion of the SSR equation in terms of perturbations of the velocity (or slowness)function, as suggested by the following equation:

Pz+1z (ω,k) = Pz (ω,k)eikz1z≈ Pz (ω,k)e

ikrefz 1z+i dkz

ds

∣∣∣sref

1s1z. (1.4)

Page 38: Chapter 1 Downward-continuation methods

The first derivative of the vertical wavenumberkz with respect to slowness is givenby

dkz

ds

∣∣∣∣sref

=ω√

1−|k|

2

ω2s2ref

. (1.5)

The methods differ in how this correction term is applied. A robust and com-putationally efficient way of applying the correction term is by an implicit finite-difference scheme. To apply implicit finite difference we first need to approximatethe square root by a rational function. This goal can be accomplished by a con-tinued fraction expansion of the square root. This method was first introduced ingeophysics to approximate directly the SSR operator (Claerbout, 1985). A goodcompromise between accuracy and computational cost is provided by the so called45-degree approximation. According to this expansion we can write the square rootin the denominator of the derivative in equation (1.5) as√

1− X2 ≈ 1−X2

2−X2

2

=4−3X2

4− X2, (1.6)

Page 39: Chapter 1 Downward-continuation methods

whereX = |k|/ωsref. Inverting this expression, we can recast the derivative as thesum of a constant (1) and a rational function ofX2; that is,

1√

1− X2≈ 1+

2X2

4−3X2, (1.7)

and equation (1.5) can be approximated as follows:

dkz

ds

∣∣∣∣sref

≈ ω

[1+

2v2refX

2

4−3v2refX

2

], (1.8)

where obviouslyvref = 1/sref.The vertical wavenumberkz is now approximated by the sum of three terms:

kz ≈ krefz +ω1s+ω1s

2(k2

x +k2y

)4ω2s2

ref −3(k2

x +k2y

)= kref

z +ksplit−stepz +kf−d

z , (1.9)

where the first term (krefz ) represents the downward continuation with a reference

Page 40: Chapter 1 Downward-continuation methods

velocity and is applied in theω-k domain, the second term (ksplit−stepz ) is the split-

step correction and is applied as a simple phase-shift in theω-x domain, and the thirdterm (kf−d

z ) is applied in theω-x domain as an implicit finite-difference operator.The direct application ofkf−d

z by an implicit finite-difference method requiresthe numerical solution of several banded systems of simultaneous equations. Thecost of solving a banded system is proportional to the number of unknowns mul-tiplied by the width of the diagonal band in the matrix. In this case the cost ishigh because the width of the diagonal band is proportional to the number of gridpoints along one of the two horizontal directions. Splitting, as presented in Sec-tion 1.3.1, can be used to reduce the computational cost. After splitting, the systemof simultaneous equations becomes tridiagonal. The computational cost of solvinga tridiagonal system is substantially smaller than solving the full banded systembecause the width of the diagonal band is only three.

To avoid splitting but maintain computational efficiency, new methods based onthe helix transform could be used to approximatekf−d

z (Rickett et al., 1998). How-ever, because splitting is applied to a residual operator, the anisotropy introducedby splitting kf−d

z is less of a concern than when splitting is applied directly to theSSR (Section1.3.1).

Page 41: Chapter 1 Downward-continuation methods

Splitting is equivalent to approximatingkf−dz as the sum of two terms:

kf−dz ≈ kf−d

zx+kf−d

zy= ω1s

(2k2

x

4ω2s2ref −3k2

x

+2k2

y

4ω2s2ref −3k2

y

). (1.10)

After splitting, each term (kf−dzx

andkf−dzy

) can be applied to the wavefield by using aCrank-Nicolson scheme that is equivalent to approximating the exponential operatoras follows:

Pz′ = Pzeikf−d

zx 1z≈ Pz

1+ i ω1s 2k2x

4ω2s2ref−3k2

x

1z2

1− i ω1s 2k2x

4ω2s2ref−3k2

x

1z2

= Pz4ω2s2

ref −3k2x + i ω1sk2

x1z

4ω2s2ref −3k2

x − i ω1sk2x1z

. (1.11)

The spatial second-derivative operator, represented byk2x, can be approximated by

Page 42: Chapter 1 Downward-continuation methods

a centered first-order finite-difference approximation of the Laplacian:

k2x = −∇

2P ≈−Px−1x +2Px

− Px+1x

12x

. (1.12)

Substituting equation (1.12) into equation (1.11) leads to the tridiagonal system ofequations determined by the following relations:

Px−1xz′ (3+ i ω1s1z)+ Px

z′

(4ω2s2

ref12x −6− i 2ω1s1z

)+ Px+1x

z′ (3+ i ω1s1z) =

Px−1xz (3− i ω1s1z)+ Px

z

(4ω2s2

ref12x −6+ i 2ω1s1z

)+ Px+1x

z (3− i ω1s1z) .

(1.13)

Implicit finite-difference methods have the important theoretical and practicaladvantage of being unconditionally stable when the velocity functions (the refer-ence velocity and the medium velocity) are smooth. However, numerical instabilitymay arise when either of the velocity functions has sharp discontinuities (Ristowand Rühl, 1994). Stability is a necessary condition for a migration method to bepractically useful.Biondi (2002) presents a reformulation of the finite-differencealgorithm outlined in equation (1.13) that guarantees stability even in presence ofsharp discontinuities in the velocity functions. His method is an adaptation of the

Page 43: Chapter 1 Downward-continuation methods

bulletproofing theory developed byGodfrey et al.(1979) andBrown (1979) for the45-degree equation.

The improvements in accuracy achieved by applying the higher-order correc-tion are evident when comparing the migration results of the SEG-EAGE salt dataset obtained using a simple split-step correction and a higher-order correction. Fig-ure 1.9 shows the cross-line section taken from the migrated cubes obtained by apseudo-screen method using three reference velocities. It should be compared withthe same section obtained using a split-step method with three reference velocities(Figure1.7) and six reference velocities (Figure1.8). The flanks of the canyons inthe middle part of the salt body and the bottom of the salt are better focused whenthe higher-order correction is applied. The quality of the image in Figure1.9is closeto the image obtained using six reference velocities and displayed in Figure1.8.

• Fourier Finite Difference (FFD)The Fourier Finite Difference (FFD) correction achieves better accuracy than thepseudo-screen correction because it is based on a direct expansion of the differencebetween the square root evaluated at the medium velocityv and the square rootevaluated at the reference velocityvref, instead of being based on the expansion of

Page 44: Chapter 1 Downward-continuation methods

Figure 1.7: Cross-line section (CMP X = 7,440) of the zero-offset SEG-EAGE saltdata set migrated with three reference velocities.down-Salt-3v-y7440[CR,M]

Page 45: Chapter 1 Downward-continuation methods

Figure 1.8: Cross-line section (CMP X = 7,440) of the zero-offset SEG-EAGE saltdata set migrated with six reference velocities.down-Salt-6v-y7440[CR,M]

Page 46: Chapter 1 Downward-continuation methods

Figure 1.9: Cross-line section (CMP X = 7,440) of the zero-offset SEG-EAGEsalt data set migrated with three reference velocities and higher-order correction.down-Salt-3v-fd-y7440[CR,M]

Page 47: Chapter 1 Downward-continuation methods

Figure 1.7: Cross-line section (CMP X = 7,440) of the zero-offset SEG-EAGE saltdata set migrated with three reference velocities.down-Salt-3v-y7440[CR,M]

Page 48: Chapter 1 Downward-continuation methods

the square root around the reference velocity. The downward-continued wavefieldis approximated as

Pz+1z (ω,k) = Pz (ω,k)eikz1z= Pz

(ω,kx,ky

)eikref

z 1z+i 1kz1s 1s1z, (1.14)

where the Taylor series of the correction term is now

1kz

1s≈ ω

[1+

vrefvX2

2+

vrefv(v2

ref +v2+vrefv

)X4

8+ ...

], (1.15)

and the continued fraction approximation of the correction term is

1kz

1s≈ ω

[1+

2vrefvX2

4−(v2

ref +v2 +vrefv)

X2

]. (1.16)

Notice that equation (1.16) reduces to equation (1.8) if v = vref. Therefore, at thelimit when the difference between the reference velocity and the medium velocity issmall, the two correction terms are equivalent, but they differ for larger corrections.

The gain in accuracy achieved by the FFD correction over the simple split-stepcorrection is illustrated in Figure1.10. It compares the phase curves obtained after

Page 49: Chapter 1 Downward-continuation methods

the split-step correction [equation (1.2)] and the FFD correction [equation (1.16)]were applied. The medium velocityv is equal to 2 km/s, and two reference veloc-ities are assumed: one 10% lower than the medium velocity (1.8 km/s), the otherone 10% higher than the medium velocity (2.2 km/s).

Figure 1.11 shows the impulse responses associated with the phase curvesshown in Figure1.10. The maximum frequency in the data is 42 Hz, and the spatialsampling is 10 m in both directions. Figure1.11c shows the exact impulse responsefor the medium velocity equal to 2 km/s. Figure1.11a shows the impulse responsewith reference velocity equal to 2.2 km/s and split-step correction. Figure1.11bshows the impulse response with reference velocity equal to 2.2 km/s and FFD cor-rection. Figure1.11d shows the impulse response with reference velocity equal to1.8 km/s and FFD correction. Figure1.11e shows the impulse response with refer-ence velocity equal to 1.8 km/s and split-step correction.

In Figure1.11, starting from the panel on the top and moving downward, theimpulse responses get narrower. The shape of the impulse responses obtained us-ing the FFD correction [panels b) and d)] is closer to the correct shape [panel c)]than the shape of the impulse responses obtained using the split-step correction[panels a) and e)]. However, for steep propagation angles the impulse responses

Page 50: Chapter 1 Downward-continuation methods

obtained using the FFD correction show clear signs of frequency-dispersion. Thesenumerical artifacts are caused by the discretization errors of the horizontal Lapla-cian operator represented byX2. To generate these figures I used the first-orderthree-point approximation of the Laplacian described in equation (1.12). The phasecurves shown in Figure1.10neglect this approximation, and thus they represent theeffective phase shift for zero-frequency data. Frequency-dispersion artifacts couldbe reduced if the accuracy of the discrete Laplacian operator were improved, forexample by use of the well-known “1/6 trick” (Claerbout, 1985). Another way toreduce frequency dispersion would be to employ more accurate, but also more com-putationally expensive, approximations of the Laplacian, such as a second-orderfive-point approximation.

To give an idea of the extent of the frequency dispersion, Figure1.12shows theimpulse response with reference velocity equal to 1.8 km/s and FFD correction, butwithout frequency dispersion. This impulse response was computed by applyingthe FFD correction in the frequency-wavenumber domain, taking advantage of thefact that both the reference and the medium velocity are constant. This option is notavailable in realistic situations when the velocities are laterally varying. The com-parison of Figure1.12with Figure1.11d shows that frequency dispersion causes the

Page 51: Chapter 1 Downward-continuation methods

high frequencies to be undercorrected with respect to the low frequencies, and withrespect to the ideal (no dispersion) situation.

Some frequency-dispersion artifacts are unavoidable whenever a finite-differencecorrection (either pseudo-screen or FFD) is applied. However, they can be substan-tially mitigated by taking advantage of the errors being in the opposite directionsfor opposite signs of the velocity correction (e.g Figure1.11b and Figure1.11d).Biondi (2002) presents a method (Fourier Finite-Difference Plus Interpolation -FFDPI) to exploit these opposite directions of the frequency-dispersion errors, sothat the related artifacts are reduced without any additional computational complex-ity.

Figure1.13presents a quantitative illustration of the frequency-dispersion er-rors and of the advantages of interpolating between two wavefields obtained withFFD corrections of opposite signs. It compares the relative phase errors measuredas a function of the propagation angle, for split-step, FFD, FFDPI, and extendedsplit-step. As in Figure1.10, the medium velocityv is equal to 2 km/s, and two ref-erence velocities are assumed: one 10% lower than the medium velocity (1.8 km/s),and one 10% higher than the medium velocity (2.2 km/s). Two temporal frequen-cies of the wavefield were assumed: 0 Hz and 100 Hz. The frequency of 100 Hz

Page 52: Chapter 1 Downward-continuation methods

corresponds to the Nyquist wavenumber for the waves propagating at 90 degreeswith velocity of 2 km/s and spatial sampling of 10 m. Therefore, the 100-Hz er-ror curves shown in Figure1.13correspond to the worst possible case for both theFFD and the FFDPI methods. As expected, the FFD corrections have smaller errorsthan the split-step correction for all angles and both frequencies (0 and 100 Hz). Atzero frequency the FFD corrections have lower error than the extended split-stepmethod, though they are computationally less expensive. However, because of fre-quency dispersion, at 100 Hz the simple FFD correction has a worse behavior thanthe extended split-step method. Interpolation between the two FFD corrections sub-stantially reduces the zero-frequency errors (to less than 1% up to 65 degrees), andbrings the 100 Hz errors below the extended split-step errors.

Interpolation between wavefields reduces also the errors caused by splitting.Section1.3.1analyzes in detail the operator anisotropy introduced by splitting. Fig-ure 1.14shows an example of the phase errors related to this operator anisotropy.It compares relative phase errors as a function of the azimuth measured for a prop-agation angle of 61 degrees. As in the previous figures, the medium velocityv isequal to 2 km/s, and two reference velocities are assumed: one 10% lower than themedium velocity (1.8 km/s), and one 10% higher than the medium velocity (2.2

Page 53: Chapter 1 Downward-continuation methods

km/s). The plots show the phase errors at two frequencies (0 Hz and 100 Hz) forthe FFDPI algorithm, the FFD correction starting from the lower reference veloc-ity, and the FFD correction starting from the higher reference velocity. Notice thatfor both the simple FFD correction cases the azimuthal anisotropy decreases as thefrequency increases, though the average phase error increases as well. But the cru-cial feature of the phase-error function for the FFD correction, is that the azimuthalvariations are in opposite directions when the differences between the reference ve-locities and the medium velocity have opposite signs. Consequently, the phase errorof the interpolation method is contained within the± 1% band and is much lowerthan the error of either of the simple FFD corrections. At higher frequencies (100Hz), the impulse response of FFDPI is almost perfectly isotropic.

Figures1.15and1.16summarize the analysis of the phase errors related to bothfrequency dispersion and operator anisotropy. Figure1.15 shows the depth sliceof three impulse responses superimposed onto each other. The outermost circularevent corresponds to the FFD correction starting from a reference velocity of 2.2km/s. The middle event corresponds to the exact impulse response with the mediumvelocity of 2 km/s. The innermost event corresponds to the FFD corrections startingfrom a reference velocity of 1.8 km/s. The depth of the slices corresponds to a

Page 54: Chapter 1 Downward-continuation methods

propagation angle of 64.2 degrees, which is close to the maximum propagationangle (65.4 degrees) for the high reference velocity (2.2 km/s). As predicted bythe curves shown in Figure1.14, the azimuthal anisotropy is frequency-dependent,and the frequency dispersion is smaller for azimuths oriented at 45 degrees withrespect to the coordinate axes.

The comparison of Figure1.16with Figure1.15demonstrates the reduction inmigration anisotropy achieved by employing FFDPI in conjunction with splitting.Figure1.16 is the merge of two impulse responses along the in-line direction, cutat the same depth as the slices shown in Figure1.15. For negative values of thein-line coordinate, the plot shows the depth slice for the exact impulse response.For positive values of the in-line coordinate, the plot shows the depth slice for theimpulse response obtained by FFDPI. It is evident that the result of the interpolationscheme is much less affected by azimuthal anisotropy and frequency dispersion thanthe results of the two simple FFD corrections showed in Figure1.15.

Page 55: Chapter 1 Downward-continuation methods

Figure 1.10: Phase curvesthat compare the accuracy of theFourier Finite-Difference (FFD)correction with the simple split-step correction. down-ffd_3dsi[CR]

0 10 20 30 40 50 60 70 80 900

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

Propagation angle (degrees)

Phase shift (rad)

Split−step (Vr=1.8 km/s) FFD (Vr=1.8 km/s)Exact (V=2.0 km/s)FFD (Vr=2.2 km/s)Split−step (Vr=2.2 km/s)

Page 56: Chapter 1 Downward-continuation methods

Figure 1.11: Impulse responsefor the medium velocity equalto 2 km/s and obtained with: a)reference velocity equal to 2.2km/s and split-step correction, b)reference velocity equal to 2.2km/s and FFD correction, c) ref-erence velocity equal to 2 km/s(i.e. exact impulse response),d) reference velocity equal to1.8 km/s and FFD correction, e)reference velocity equal to 1.8km/s and split-step correction.Starting from the panel on thetop and moving downward, theimpulse responses get narrower.down-Line-impulse-all-overn[CR,M]

Page 57: Chapter 1 Downward-continuation methods

Figure 1.12: Impulse responsefor the medium velocity equal to2 km/s and obtained with refer-ence velocity equal to 1.8 km/sand FFD correction applied inthe wavenumber domain and thuswithout frequency dispersion.Compare with Figure 1.11d.down-Line-impulse-pos-hd-overn[CR,M]

Page 58: Chapter 1 Downward-continuation methods

Figure 1.13: Relative phase-error curves assumingv=2 km/sand starting from two refer-ence velocities (vref=1.8 km/sandvref=2.2 km/s), for split step,FFD, FFDPI and extended split-step. Two temporal frequenciesof the wavefield were assumed:0 Hz and 100 Hz. The verti-cal solid line indicates the maxi-mum propagation angle (65.4 de-grees) whenvref=2.2 km/s andv=2 km/s. The horizontal solidlines indicate the 1% phase errorlevel. down-errfreq_3dsi[CR]

0 10 20 30 40 50 60 70 80 90−10

−8

−6

−4

−2

0

2

4

6

8

10

Propagation angle (degrees)

Relative phase error (%

) Split step (Vr=1.8 km/s)FFD (Vr=1.8 km/s, f=100 Hz)FFD (Vr=1.8 km/s, f=0 Hz)FFD (Vr=2.2 km/s, f=0 Hz)FFD (Vr=2.2 km/s, f=100 Hz)Split step (Vr=2.2 km/s)FFDPI (f=0 Hz)FFDPI (f=100 Hz)Extended Split Step

Page 59: Chapter 1 Downward-continuation methods

Figure 1.14: Relative phase-error curves for FFD and FFDPI,as a function of the azimuth.The medium velocity was as-sumed to bev=2 km/s and thetwo reference velocities werevref=1.8 km/s andvref=2.2 km/s.Two temporal frequencies of thewavefield were assumed: 0 Hzand 100 Hz. The horizontal solidlines indicate the± 1% phase-error level. down-azimerr_3dsi[CR]

0 50 100 150 200 250 300 350

−20

−10

0

10

20

30

Azimuth (degrees)

Relative phase error (%

)

FFD (Vr=1.8 km/s, f=100 Hz) FFD (Vr=1.8 km/s, f=0 Hz)FFD (Vr=2.2 km/s, f=0 Hz)FFD (Vr=2.2 km/s, f=100 Hz)FFDPI (f=0 Hz)FFDPI (f=100 Hz)

Page 60: Chapter 1 Downward-continuation methods

Figure 1.15: Depth slicesthrough impulse responses: 1) in-nermost event corresponds to theFFD corrections starting from areference velocity of 1.8 km/s, 2)middle event corresponds to theexact impulse response with themedium velocity of 2 km/s, 3)outermost event corresponds tothe FFD corrections starting froma reference velocity of 2.2 km/s.down-Aniso-rist [CR]

Page 61: Chapter 1 Downward-continuation methods

Figure 1.16: Depth slices through impulse responses: 1) left half correspondsto the exact impulse response with the medium velocity of 2 km/s, 2) right halfcorresponds to the FFDPI results.down-Iso-rist [CR]

Page 62: Chapter 1 Downward-continuation methods

significant differences are noticeable between the two sections, with the exceptionthat the more expensive migration preserved slightly better the steep reflections.

Figure 1.23 shows two depth slices extracted from the same migrated cubesshown in Figure1.22. The image obtained with the more expensive filters (right)is slightly sharper than the other one. Again, the main differences are visible in thevicinity of the salt-sediment boundary.

REFERENCESBiondi, B. L., and Palacharla, G., 1994, 3-D depth migration by rotated McClellan

filter: Geophysical Prospecting,43, 1005–1020.52

Biondi, B., 2002, Stable wide-angle Fourier finite-difference downward extrapola-tion of 3-D wavefields: Geophysics,67, no. 3, 872–882.24, 30

Brown, D., 1979, Muir’s rules for matrices: Another look at stability: SEP–20,125–142. 25

Claerbout, J. F., 1985, Imaging the Earth’s Interior: Blackwell Scientific Publica-tions. 20, 29

Page 63: Chapter 1 Downward-continuation methods

Gazdag, J., and Sguazzero, P., 1984, Migration of seismic data by phase-shift plusinterpolation: Geophysics,49, no. 2, 124–131.5

Gazdag, J., 1978, Wave equation migration with the phase-shift method: Geo-physics,43, no. 10, 1342–1351.3

Godfrey, R. J., Muir, F., and Claerbout, J. F., 1979, Stable extrapolation: SEP–16,83–87. 25

Hale, D., 1991, 3-D depth migration via McClellan transformations: Geophysics,56, 1778–1785.45, 49, 51

Huang, L.-J., Fehler, M. C., and Wu, R.-S., 1999, Extended local Born Fouriermigration method: submitted for publication to Geophysics.19

Kessinger, W., 1992, Extended split-step Fourier migration: 62nd Annual Internat.Mtg., Soc. Expl. Geophys., Expanded Abstracts, 917–920.9

Li, Z., 1991, Compensating finite-difference errors in 3-D migration and modeling:Geophysics,56, no. 10, 1650–1660.45

Page 64: Chapter 1 Downward-continuation methods

McClellan, J., and Chan, D., 1977, A 2-D FIR filter structure derived from theChebyshev recursion: IEEE Trans. Circuits Syst.,CAS-24, 372–384. 49

Nautiyal, A., Gray, S. H., Whitmore, N. D., and Garing, J. D., 1993, Stability versusaccuracy for an explicit wavefield extrapolation operator: Geophysics,58, 277–283. 49

Rickett, J., Claerbout, J., and Fomel, S., 1998, Implicit 3-D depth migration bywavefield extrapolation with helical boundary conditions: 68th Ann. Internat.Meeting, Soc. Expl. Geophys., 1124–1127.22

Ristow, D., and Rühl, T., 1994, Fourier finite-difference migration: Geophysics,59,no. 12, 1882–1893.19, 24

Soubaras, R., 1994, Signal-preserving random noise attenuation by the f-x projec-tion:, in 64th Ann. Internat. Mtg Soc. of Expl. Geophys., 1576–1579.53

Stoffa, P. L., Fokkema, J., de Luna Freire, R. M., and Kessinger, W. P., 1990, Split-step Fourier migration: Geophysics,55, no. 4, 410–421.6

Page 65: Chapter 1 Downward-continuation methods

Xie, X. B., and Wu, R. S., 1999, Improving the wide angle accuracy of the screenpropagator for elastic wave propagation:,in 69th Ann. Internat. Mtg Soc. of Expl.Geophys., 1863–1866.19


Recommended