next up previous [pdf]

Next: About this document ... Up: Fomel: Velocity continuation Previous: Acknowledgments

Bibliography

Adler, F., 2002, Kirchhoff image propagation: Geophysics, 67, 126-134.

Al-Yahya, K., and P. Fowler, 1986, Prestack residual migration, in SEP-50: Stanford Exploration Project, 219-230.

Beasley, C., W. Lynn, K. Larner, and H. Nguyen, 1988, Cascaded frequency-wavenumber migration - Removing the restrictions on depth-varying velocity: Geophysics, 53, 881-893.

Belonosova, A. V., and A. S. Alekseev, 1967, in About one formulation of the inverse kinematic problem of seismics for a two-dimensional continuously heterogeneous medium: Nauka, 137-154.

Berkhout, A. J., 1985, Seismic migration: Imaging of acoustic energy by wave field extrapolation: Elsevier, Amsterdam.

Chun, J. H., and C. A. Jacewitz, 1981, Fundamentals of frequency-domain migration: Geophysics, 46, 717-733.

Claerbout, J. F., 1985, Imaging the Earth's Interior: Blackwell Scientific Publications.

----, 1986, Velocity extrapolation by cascaded 15 degree migration, in SEP-48: Stanford Exploration Project, 79-84.

Clayton, R. W., 1978, Common midpoint migration, in SEP-14: Stanford Exploration Project, 21-36.

Courant, R., and D. Hilbert, 1989, Methods of mathematical physics: John Wiley & Sons.

Deregowski, S. M., 1990, Common-offset migrations and velocity analysis: First Break, 08, 224-234.

Dunkin, J. W., and F. K. Levin, 1973, Effect of normal moveout on a seismic pulse: Geophysics, 38, 635-642.

Etgen, J., 1990, Residual prestack migration and interval velocity estimation: PhD thesis, Stanford University.

Fomel, S., 1997, Velocity continuation and the anatomy of prestack residual migration: 67th Ann. Internat. Mtg, Soc. of Expl. Geophys., 1762-1765.

----, 2001, Migration velocity analysis by velocity continuation: 71st Ann. Internat. Mtg, Soc. of Expl. Geophys., 1107-1110.

----, 2003a, Theory of differential offset continuation: Geophysics, in press.

----, 2003b, Time migration velocity analysis by velocity continuation: Geophysics, submitted for publication.

Fomel, S. B., 1994, Method of velocity continuation in the problem of temporal seismic migration: Russian Geology and Geophysics, 35, 100-111.

Fowler, P., 1984, Velocity independent imaging of seismic reflectors: 54th Ann. Internat. Mtg, Soc. of Expl. Geophys., Session:S1.8.

----, 1988, Seismic velocity estimation using prestack time migration: PhD thesis, Stanford University.

Gradshtein, I. S., and I. M. Ryzhik, 1994, Table of integrals, series, and products: Boston: Academic Press.

Hale, I. D., 1983, Dip moveout by Fourier transform: PhD thesis, Stanford University.

Hubral, P., M. Tygel, and J. Schleicher, 1996, Seismic image waves: Geophysical Journal International, 125, 431-442.

Jakubowicz, H., and S. Levin, 1983, A simple exact method of three-dimensional migration - Theory: Geophys. Prosp., 31, 34-56.
(Discussion in GPR-34-06-0927-0939 with reply by author; Comment in GPR-32-02-0347-0350 with reply by author).

Larner, K., and C. Beasley, 1987, Cascaded migrations - Improving the accuracy of finite-difference migration: Geophysics, 52, 618-643.
(Errata in GEO-52-8-1165).

Levin, S., 1986a, Cascaded fifteen degree equations simplified, in SEP-48: Stanford Exploration Project, 101-108.

----, 1986b, Test your migration IQ, in SEP-48: Stanford Exploration Project, 147-160.

Liptow, F., and P. Hubral, 1995, Migrating around in circles: The Leading Edge, 14, 1125-1127.

Liu, H., and G. McMechan, 1996, Dynamic residual prestack depth migration for common offset gathers: UTD Report.

Liu, Z., and N. Bleistein, 1995, Migration velocity analysis: Theory and an iterative algorithm: Geophysics, 60, 142-153.

MacKay, S., and R. Abma, 1992, Imaging and velocity estimation with depth-focusing analysis: Geophysics, 57, 1608-1622.

Mikulich, W., and D. Hale, 1992, Steep-dip v(z) imaging from an ensemble of Stolt-like migrations: Geophysics, 57, 51-59.

Popovici, A. M., 1995, Prestack migration by split-step DSR, in SEP-84: Stanford Exploration Project, 53-60.

Rothman, D. H., S. A. Levin, and F. Rocca, 1985, Residual migration - Applications and limitations: Geophysics, 50, 110-126.

Sava, P., and S. Fomel, 2003, Wave-equation angle-domain common-image gathers: Geophysics, in press.

Schleicher, J., P. Hubral, G. Hocht, and F. Liptow, 1997, Seismic constant-velocity remigration: Geophysics, 62, 589-597.

Schneider, W. A., 1978, Integral formulation for migration in two-dimensions and three-dimensions: Geophysics, 43, 49-76.

Shurtleff, R. N., 1984, An F-K procedure for prestack migration and migration velocity analysis, in Presented at the 46th Annual EAGE Mtg., London: EAGE.

Stolt, R. H., 1978, Migration by Fourier transform: Geophysics, 43, 23-48.
(Discussion and reply in GEO-60-5-1583).

----, 1996, Short note - A prestack residual time migration operator: Geophysics, 61, 605-607.

Tygel, M., J. Schleicher, and P. Hubral, 1994, Pulse distortion in depth migration: Geophysics, 59, 1561-1569.

Yilmaz, O., 1979, Prestack partial migration: PhD thesis, Stanford University.

Yilmaz, O., and R. E. Chambers, 1984, Migration velocity analysis by wave-field extrapolation: Geophysics, 49, 1664-1674.

Yilmaz, O., and J. F. Claerbout, 1980, Prestack partial migration: Geophysics, 45, 1753-1779.

Appendix A

DERIVING THE KINEMATIC EQUATIONS

The main goal of this appendix is to derive the partial differential equation describing the image surface in a depth-midpoint-offset-velocity space.

vlcray
Figure A-1.
Reflection rays in a constant velocity medium (a scheme).
vlcray
[pdf] [png] [xfig]

The derivation starts with observing a simple geometry of reflection in a constant-velocity medium, shown in Figure A-1. The well-known equations for the apparent slowness

\begin{displaymath}
{{\partial t} \over {\partial s}} \,=\,
{ {\sin{\alpha_1}} \over {v}}\;,
\end{displaymath} (52)


\begin{displaymath}
{{\partial t} \over {\partial r}} \,=\,
{{\sin{\alpha_2}} \over {v}}
\end{displaymath} (53)

relate the first-order traveltime derivatives for the reflected waves to the emergence angles of the incident and reflected rays. Here $s$ stands for the source location at the surface, $r$ is the receiver location, $t$ is the reflection traveltime, $v$ is the constant velocity, and $\alpha_1$ and $\alpha_2$ are the angles shown in Figure A-1. Considering the traveltime derivative with respect to the depth of the observation surface $z$ shows that the contributions of the two branches of the reflected ray, added together, form the equation
\begin{displaymath}
- {{\partial t} \over {\partial z}} \,=\,
{{\cos{\alpha_1}} \over {v}} +
{{\cos{\alpha_2}} \over {v}}\;.
\end{displaymath} (54)

It is worth mentioning that the elimination of angles from equations (A-1), (A-2), and (A-3) leads to the famous double-square-root equation,
\begin{displaymath}
- v\,{{\partial t} \over {\partial z}} \,=\,
\sqrt{1 - v^2\,...
...{1 - v^2\,\left({{\partial t} \over {\partial r}}\right)^2}\;,
\end{displaymath} (55)

published in the Russian literature by Belonosova and Alekseev (1967) and commonly used in the form of a pseudo-differential dispersion relation (Claerbout, 1985; Clayton, 1978) for prestack migration (Popovici, 1995; Yilmaz, 1979). Considered locally, equation (A-4) is independent of the constant velocity assumption and enables recursive prestack downward continuation of reflected waves in heterogeneous isotropic media.

Introducing the midpoint coordinate $x = {{s+ r} \over 2}$ and half-offset $h = {{r - s} \over 2}$, one can apply the chain rule and elementary trigonometric equalities to formulas (A-1) and (A-2) and transform these formulas to

\begin{displaymath}
{{\partial t} \over {\partial x}} \,=\,
{{\partial t} \over...
...tial r}} \,=\,
{ {2 \sin{\alpha}\,\cos{\theta}} \over {v}}\;,
\end{displaymath} (56)


\begin{displaymath}
{{\partial t} \over {\partial h}} \,=\,
{{\partial t} \over ...
...ial s}} \,=\,
{ {2 \cos{\alpha}\,\sin{\theta}} \over {v}} \;,
\end{displaymath} (57)

where $\alpha = {{\alpha_1 + \alpha_2} \over 2}$ is the dip angle, and $\theta = {{\alpha_2 - \alpha_1} \over 2}$ is the reflection angle (Claerbout, 1985; Clayton, 1978). Equation (A-3) transforms analogously to
\begin{displaymath}
- {{\partial t} \over {\partial z}} \,=\,
{{2 \cos{\alpha} \cos{\theta}} \over {v}}\;.
\end{displaymath} (58)

This form of equation (A-3) is used to describe the stretching factor of the waveform distortion in depth migration (Tygel et al., 1994).

Dividing (A-5) and (A-6) by (A-7) leads to

\begin{displaymath}
{{\partial z} \over {\partial x}} \,=\,
- \tan{\alpha}\;,
\end{displaymath} (59)


\begin{displaymath}
{{\partial z} \over {\partial h}} \,=\,
- \tan{\theta}\;.
\end{displaymath} (60)

Equation (A-9) is the basis of the angle-gather construction of Sava and Fomel (2003). Substituting formulas (A-8) and (A-9) into equation (A-7) yields yet another form of the double-square-root equation:
\begin{displaymath}
- {{\partial t} \over {\partial z}} \,=\, {2 \over {v}}\,
\l...
...left({\partial z} \over {\partial h}\right)^2}\right]^{-1}\;,
\end{displaymath} (61)

which is analogous to the dispersion relationship of Stolt prestack migration (Stolt, 1978).

The law of sines in the triangle formed by the incident and reflected ray leads to the explicit relationship between the traveltime and the offset:

\begin{displaymath}
v\,t = 2\,h\, {{\cos{\alpha_1}+ \cos{\alpha_2}} \over
\sin{\...
...alpha_1\right)}} = 2\,h\,{\cos{\alpha} \over
\sin{\theta}} \;.
\end{displaymath} (62)

An algebraic combination of formulas (A-11), (A-5), and (A-6) forms the basic kinematic equation of the offset continuation theory (Fomel, 2003a):
\begin{displaymath}
{{\partial t} \over {\partial h}} \,
\left(t^2 + {{4\,h^2} \...
...\,-
\left({{\partial t} \over {\partial x}}\right)^2\right)\;.
\end{displaymath} (63)

Differentiating (A-11) with respect to the velocity $v$ yields

\begin{displaymath}
- v^2\,{{\partial t} \over {\partial v}} =
2\,h\,{\cos{\alpha} \over \sin{\theta}}\;.
\end{displaymath} (64)

Finally, dividing (A-13) by (A-7) produces
\begin{displaymath}
v\,{{\partial z} \over {\partial v}} =
{h \over {\cos{\theta}\,\sin{\theta}}}\;.
\end{displaymath} (65)

Equation (A-14) can be written in a variety of ways with the help of an explicit geometric relationship between the half-offset $h$ and the depth $z$,
\begin{displaymath}
h = z\,
{{\sin{\theta}\,\cos{\theta}} \over
{\cos^2{\alpha}-\sin^2{\theta}}}\;,
\end{displaymath} (66)

which follows directly from the trigonometry of the triangle in Figure A-1 (Fomel, 2003a). For example, equation (A-14) can be transformed to the form obtained by Liu and Bleistein (1995):
\begin{displaymath}
v\,{{\partial z} \over {\partial v}} =
{z \over{\cos^2{\alp...
...sin^2{\theta}}} =
{z \over{\cos{\alpha_1}\,\cos{\alpha_2}}}\;.
\end{displaymath} (67)

In order to separate different factors contributing to the velocity continuation process, one can transform this equation to the form
$\displaystyle v\,{{\partial z} \over {\partial v}}$ $\textstyle =$ $\displaystyle {z \over {\cos^2{\alpha}}} +
{{h^2} \over z}\,\left(1-\tan^2{\alpha}\,\tan^2{\theta}\right)$  
  $\textstyle =$ $\displaystyle z\,\left(1 + \left({{\partial z} \over {\partial x}}\right)^2\rig...
...rtial x}}\right)^2\,
\left({{\partial z} \over {\partial h}}\right)^2\right)\;.$ (68)

Rewritten in terms of the vertical traveltime $\tau = z/v$, it further transforms to equation
\begin{displaymath}
{{\partial \tau} \over {\partial v}} =
v\,\tau\,\left({{\pa...
...
\left({{\partial \tau} \over {\partial h}}\right)^2\right)\;,
\end{displaymath} (69)

equivalent to equation (1) in the main text. Yet another form of the kinematic velocity continuation equation follows from eliminating the reflection angle $\theta$ from equations (A-14) and (A-15). The resultant expression takes the following form:
\begin{displaymath}
v\,{{\partial z} \over {\partial v}} =
{{2\,(z^2 + h^2)} \o...
...+ {{2\,h^2} \over
{\sqrt{z^2 + h^2 \sin^2{2\,\alpha}} + z}}\;.
\end{displaymath} (70)

Appendix B

Derivation of the residual DMO kinematics

This appendix derives the kinematical laws for the residual NMO+DMO transformation in the prestack offset continuation process.

The direct solution of equation (31) is nontrivial. A simpler way to obtain this solution is to decompose residual NMO+DMO into three steps and to evaluate their contributions separately. Let the initial data be the zero-offset reflection event $\tau_0(x_0)$. The first step of the residual NMO+DMO is the inverse DMO operator. One can evaluate the effect of this operator by means of the offset continuation concept (Fomel, 2003a). According to this concept, each point of the input traveltime curve $\tau_0(x_0)$ travels with the change of the offset from zero to $h$ along a special trajectory, which I call a time ray. Time rays are parabolic curves of the form

\begin{displaymath}
x\left(\tau\right) = x_0+{{\tau^2-\tau_0^2\left(x_0\right)} \over
{\tau_0\left(x_0\right)\,\tau_0'\left(x_0\right)}}\;,
\end{displaymath} (71)

with the final points constrained by the equation
\begin{displaymath}
h^2 = \tau^2\,{{\tau^2-\tau_0^2\left(x_0\right)} \over
{\left(\tau_0\left(x_0\right)\,\tau_0'\left(x_0\right)\right)^2}}\;,
\end{displaymath} (72)

where $\tau_0'\left(x_0\right)$ is the derivative of $\tau_0\left(x_0\right)$. The second step of the cumulative residual NMO+DMO process is the residual normal moveout. According to equation (23), residual NMO is a one-trace operation transforming the traveltime $\tau$ to $\tau_1$ as follows:
\begin{displaymath}
\tau_1^2 = \tau^2 + h^2\,d\;,
\end{displaymath} (73)

where
\begin{displaymath}
d = \left({1 \over v_0^2} - {1 \over v_1^2}\right)\;.
\end{displaymath} (74)

The third step is dip moveout corresponding to the new velocity $v_1$. DMO is the offset continuation from $h$ to zero offset along the redefined time rays (Fomel, 2003a)
\begin{displaymath}
x_2\left(\tau_2\right) =
x + {{h\,X} \over {\tau_1^2\,H}}\,\left(\tau_1^2-\tau_2^2\right)\;,
\end{displaymath} (75)

where $H = {{\partial \tau_1} \over {\partial h}}$, and $X =
{{\partial \tau_1} \over {\partial x}}$. The end points of the time rays (B-5) are defined by the equation
\begin{displaymath}
\tau_2^2 = - \tau_1^2\,{{\tau_1\,H} \over {h\,X^2}}\;.
\end{displaymath} (76)

The partial derivatives of the common-offset traveltimes are constrained by the offset continuation kinematic equation
\begin{displaymath}
h\,(H^2 - X^2) = \tau_1\,H\;,
\end{displaymath} (77)

which is equivalent to equation (A-12) in Appendix A. Additionally, as follows from equations (B-3) and the ray invariant equations from (Fomel, 2003a),
\begin{displaymath}
\tau_1\,X = \tau\,{{\partial \tau} \over {\partial x}} =
{{...
...2\,\tau_0'\left(x_0\right)} \over {\tau_0\left(x_0\right)}}\;.
\end{displaymath} (78)

Substituting (B-1-B-4) and (B-7-B-8) into equations (B-5) and (B-6) and performing the algebraic simplifications yields the parametric expressions for velocity rays of the residual NMO+DMO process:
\begin{displaymath}
\left\{
\begin{array}{rcl}
x_2(d) & = & \displaystyle{x_0 + ...
...laystyle{{{\tau_1^2(d)} \over {T_2(d)}}}\;,
\end{array}\right.
\end{displaymath} (79)

where the function $T\left(h,\tau_0(x_0),\tau_0'\left(x_0\right)\right)$ is defined by
\begin{displaymath}
T\left(h,\tau,\tau_x\right) =
{{\tau + \sqrt{\tau^2 + 4\,h^2\,\tau_x^2}} \over 2}\;,
\end{displaymath} (80)


\begin{displaymath}
T_2(d) = \sqrt{T\left(h,\tau_1^2(d),\tau_0'\left(x_0\right)\,
T\left(h,\tau_0(x_0),\tau_0'\left(x_0\right)\right)\right)}\;,
\end{displaymath} (81)

and
\begin{displaymath}
\tau_1^2(d) = \tau_0\,T + d\,h^2\;.
\end{displaymath} (82)

The last step of the cascade of inverse DMO, residual NMO, and DMO is illustrated in Figure B-1. The three plots in the figure show the offset continuation to zero offset of the inverse DMO impulse response shifted by the residual NMO operator. The middle plot corresponds to zero NMO shift, for which the DMO step collapses the wavefront back to a point. Both positive (top plot) and negative (bottom plot) NMO shifts result in the formation of the specific triangular impulse response of the residual NMO+DMO operator. As noticed by Etgen (1990), the size of the triangular operators dramatically decreases with the time increase. For large times (pseudo-depths) of the initial impulses, the operator collapses to a point corresponding to the pure NMO shift.

vlcvoc
Figure B-1.
Kinematic residual NMO+DMO operators constructed by the cascade of inverse DMO, residual NMO, and DMO. The impulse response of inverse DMO is shifted by the residual NMO procedure. Offset continuation back to zero offset forms the impulse response of the residual NMO+DMO operator. Solid lines denote traveltime curves; dashed lines denote the offset continuation trajectories (time rays). Top plot: $v_1/v_0 = 1.2$. Middle plot: $v_1/v_0 = 1$; the inverse DMO impulse response collapses back to the initial impulse. Bottom plot: $v_1/v_0 = 0.8$. The half-offset $h$ in all three plots is 1 km.
vlcvoc
[pdf] [png] [sage]

Appendix C

INTEGRAL VELOCITY CONTINUATION AND KIRCHHOFF MIGRATION

The main goal of this appendix is to prove the equivalence between the result of zero-offset velocity continuation from zero velocity and conventional post-stack migration. After solving the velocity continuation problem in the frequency domain, I transform the solution back to the time-and-space domain and compare it with the conventional Kirchhoff migration operator (Schneider, 1978). The frequency-domain solution has its own value, because it forms the basis for an efficient spectral algorithm for velocity continuation (Fomel, 2003b).

Zero-offset migration based on velocity continuation is the solution of the boundary problem for equation (41) with the boundary condition

\begin{displaymath}
\left.P\right\vert _{v=0} = P_0\;,
\end{displaymath} (83)

where $P_0(t_0,x_0)$ is the zero-offset seismic section, and $P(t,x,v)$ is the continued wavefield. In order to find the solution of the boundary problem composed of (41) and (C-1), it is convenient to apply the function transformation $R(t,x,v) = t\,P(t,x,v)$, the time coordinate transformation $\sigma = t^2/2$, and, finally, the double Fourier transform over the squared time coordinate $\sigma$ and the spatial coordinate $x$:
\begin{displaymath}
\widehat{R}(v) = \int \int\,P(t,x,v)\,
\exp(i \Omega \sigma - i k x )\,t^2\,dt\,dx\;.
\end{displaymath} (84)

With the change of domain, equation (41) transforms to the ordinary differential equation
\begin{displaymath}
{{d\,\widehat{R}} \over {d\,v}} =
i\,{k^2 \over \Omega}\,v\,\widehat{R}\;,
\end{displaymath} (85)

and the boundary condition (C-1) transforms to the initial value condition
\begin{displaymath}
\widehat{R}(0) = \widehat{R}_0\;,
\end{displaymath} (86)

where
\begin{displaymath}
\widehat{R}_0 = \int \int\,P_0(t_0,x_0)\,
\exp(i \Omega \sigma_0 - i k x_0 )\,t_0^2\,dt_0\,dx_0\;,
\end{displaymath} (87)

and $\sigma_0 = t_0^2/2$. The unique solution of the initial value (Cauchy) problem (C-3) - (C-4) is easily found to be
\begin{displaymath}
\widehat{R}(v) = \widehat{R}_0\,
\exp\left( i\,{{k^2} \over {2\,\Omega}}\,v^2\right)\;.
\end{displaymath} (88)

In the transformed domain, velocity continuation appears to be a unitary phase-shift operator. An immediate consequence of this remarkable fact is the cascaded migration decomposition of post-stack migration (Larner and Beasley, 1987):

\begin{displaymath}
\exp\left( i\,{{k^2} \over {2\,\Omega}}\,
(v_1^2 + \cdots + ...
...ots\,
\exp\left( i\,{{k^2} \over {2\,\Omega}}\,v_n^2\right)\;.
\end{displaymath} (89)

Analogously, three-dimensional post-stack migration is decomposed into the two-pass procedure (Jakubowicz and Levin, 1983):
\begin{displaymath}
\exp\left( i\,{{k_1^2+k_2^2} \over {2\,\Omega}}\,v^2\right) ...
...ht)\,
\exp\left( i\,{{k_2^2} \over {2\,\Omega}}\,v^2\right)\;.
\end{displaymath} (90)

The inverse double Fourier transform of both sides of equality (C-6) yields the integral (convolution) operator

\begin{displaymath}
P(t,x,v) = \int\int\,P_0(t_0,x_0)\,K(t_0,x_0;t,x,v)\,dt_0\,dx_0\;,
\end{displaymath} (91)

with the kernel $K$ defined by
\begin{displaymath}
K = {{t_0^2/t} \over {(2\,\pi)^{m+1}}}\,
\int\int\,\exp\left...
... -
{{i\Omega} \over 2}\,(t^2 - t_0^2)
\right)\,dk\,d\Omega\;,
\end{displaymath} (92)

where $m$ is the number of dimensions in $x$ and $k$ ($m$ equals $1$ or $2$). The inner integral on the wavenumber axis $k$ in formula (C-10) is a known table integral (Gradshtein and Ryzhik, 1994). Evaluating this integral simplifies equation (C-10) to the form
\begin{displaymath}
K = {{t_0^2/t} \over {(2\,\pi)^{m/2+1}\,v^m}}\,
\int\,(i\Ome...
...2 - t^2 - {{(x - x_0)^2} \over v^2}\right)\right]\,
d\Omega\;.
\end{displaymath} (93)

The term $(i\Omega)^{m/2}$ is the spectrum of the anti-causal derivative operator ${d \over {d\sigma}}$ of the order $m/2$. Noting the equivalence
\begin{displaymath}
\left({\partial \over {\partial \sigma}}\right)^{m/2} =
\lef...
...t)^{m/2}\,
\left({\partial \over {\partial t}}\right)^{m/2}\;,
\end{displaymath} (94)

which is exact in the 3-D case ($m=2$) and asymptotically correct in the 2-D case ($m=1$), and applying the convolution theorem transforms operator (C-9) to the form
\begin{displaymath}
P(t,x,v) = {1 \over {(2\,\pi)^{m/2}}}\,\int\,
{{\cos{\alpha}...
...t_0}}\right)^{m/2}
P_0\left({\rho \over v},x_0\right)\,dx_0\;,
\end{displaymath} (95)

where $\rho = \sqrt{v^2\,t^2 + (x - x_0)^2}$, and $\cos{\alpha} =
t_0/t$. Operator (C-13) coincides with the Kirchhoff operator of conventional post-stack time migration (Schneider, 1978).


next up previous [pdf]

Next: About this document ... Up: Fomel: Velocity continuation Previous: Acknowledgments

2014-04-01