Acoustic anisotropic wavefields through perturbation theory

Similar documents
Traveltime approximations for inhomogeneous transversely isotropic media with a horizontal symmetry axis

3D VTI traveltime tomography for near-surface imaging Lina Zhang*, Jie Zhang, Wei Zhang, University of Science and Technology of China (USTC)

Compensating visco-acoustic effects in anisotropic resverse-time migration Sang Suh, Kwangjin Yoon, James Cai, and Bin Wang, TGS

Wave-equation tomography for anisotropic parameters

An acoustic wave equation for orthorhombic anisotropy

An efficient wave extrapolation method for tilted orthorhombic media using effective ellipsoidal models

Chapter 1. Introduction EARTH MODEL BUILDING

H005 Pre-salt Depth Imaging of the Deepwater Santos Basin, Brazil

Prestack exploding reflector modelling and migration for anisotropic media

Correspondence between the low- and high-frequency limits for anisotropic parameters in a layered medium

Stanford Exploration Project, Report 97, July 8, 1998, pages

Anisotropic tomography for TTI and VTI media Yang He* and Jun Cai, TGS

Residual migration in VTI media using anisotropy continuation

Geophysical Journal International

Elastic wave-mode separation for VTI media

Nonhyperbolic Reflection Moveout for Orthorhombic Media

SUMMARY ANGLE DECOMPOSITION INTRODUCTION. A conventional cross-correlation imaging condition for wave-equation migration is (Claerbout, 1985)

Moveout approximation for P waves in a homogeneous VTI medium

Elastic wavefield separation for VTI media

Fourier finite-difference wave propagation a

AVAZ inversion for fracture orientation and intensity: a physical modeling study

2010 SEG SEG Denver 2010 Annual Meeting

Reflection moveout and parameter estimation for horizontal transverse isotropy

Angle-domain common-image gathers from anisotropic Gaussian beam migration and its application to anisotropy-induced imaging errors analysis

TTI Anisotropic Depth Migration: Which Tilt Estimate Should We Use?

Improvement of stacking image by anisotropic velocity analysis using P-wave seismic data

P185 TTI Anisotropic Depth Migration - Which Tilt Estimate Should We Use?

SEG Houston 2009 International Exposition and Annual Meeting

Lowrank RTM for converted wave imaging

Anisotropic Seismic Imaging and Inversion for Subsurface Characterization at the Blue Mountain Geothermal Field in Nevada

Attenuation compensation in viscoacoustic reserve-time migration Jianyong Bai*, Guoquan Chen, David Yingst, and Jacques Leveille, ION Geophysical

Václav Bucha. Department of Geophysics Faculty of Mathematics and Physics Charles University in Prague. SW3D meeting June 6-7, 2016 C OM S TR 3 D

Effects of Fracture Parameters in an Anisotropy Model on P-Wave Azimuthal Amplitude Responses

Depth-domain velocity analysis in VTI media using surface P-wave data: Is it feasible?

A synthetic example of anisotropic P -wave. processing for a model from the Gulf of Mexico

The Deconvolution of Multicomponent Trace Vectors

Edinburgh Research Explorer

Body-wave radiation patterns and AVO in transversely isotropic media

Seismic Waves in Complex 3 D Structures, 26 (2016), (ISSN , online at

Analysis of image gathers in factorized VTI media

Seismic inversion for the parameters of two orthogonal fracture sets in a VTI background medium

SEG/New Orleans 2006 Annual Meeting. Non-orthogonal Riemannian wavefield extrapolation Jeff Shragge, Stanford University

ANISOTROPIC PRESTACK DEPTH MIGRATION: AN OFFSHORE AFRICA CASE STUDY

P137 Our Experiences of 3D Synthetic Seismic Modeling with Tip-wave Superposition Method and Effective Coefficients

Elastic least-squares reverse time migration

3-D description of normal moveout in anisotropic inhomogeneous media

Implicit 3-D depth migration by wavefield extrapolation with helical boundary conditions

Kirchhoff prestack depth migration in simple models of various anisotropy

Attenuation compensation in least-squares reverse time migration using the visco-acoustic wave equation

An Improvement of Velocity Variation with Offset (VVO) Method in Estimating Anisotropic Parameters

Finite difference elastic modeling of the topography and the weathering layer

Dip-constrained tomography with weighting flow for paleo-canyons: a case study in Para- Maranhao Basin, Brazil Guang Chen and Lingli Hu, CGG

Title: Exact wavefield extrapolation for elastic reverse-time migration

The signature of attenuation and anisotropy on AVO and inversion sensitivities

G022 Multi-azimuth Seismic Data Imaging in the Presence of Orthorhombic Anisotropy

Lowrank finite-differences and lowrank Fourier finite-differences for seismic wave extrapolation in the acoustic approximation a

NMO ellipse for a stratified medium with laterally varying velocity

Interval anisotropic parameters estimation in a least squares sense Case histories from West Africa

One-step extrapolation method for reverse time migration

Complex-beam Migration and Land Depth Tianfei Zhu CGGVeritas, Calgary, Alberta, Canada

SHEAR-WAVE REFLECTION MOVEOUT FOR AZIMUTHALLY ANISOTROPIC MEDIA

AVAZ and VVAZ practical analysis to estimate anisotropic properties

Comparison between least-squares reverse time migration and full-waveform inversion

On anelliptic approximations for qp velocities in VTI media

C031 Quantifying Structural Uncertainty in Anisotropic Depth Imaging - Gulf of Mexico Case Study

Fowler DMO and time migration for transversely isotropic media

Kirchhoff prestack depth migration in simple models with differently rotated elasticity tensor: orthorhombic and triclinic anisotropy

Pseudo-acoustic wavefield propagation for anisotropic media

Finite difference modelling of the full acoustic wave equation in Matlab

Chapter 7. Seismic imaging. 7.1 Assumptions and vocabulary

Snell s law in transversely isotropic media using linearized group velocities and related quantities

Acoustic Anisotropy Measurements and Interpretation in Deviated Wells

Direct nonlinear traveltime inversion in layered VTI media Paul J. Fowler*, Alexander Jackson, Joseph Gaffney, and David Boreham, WesternGeco

Waveform inversion for attenuation estimation in anisotropic media Tong Bai & Ilya Tsvankin Center for Wave Phenomena, Colorado School of Mines

Velocity Update Using High Resolution Tomography in Santos Basin, Brazil Lingli Hu and Jianhang Zhou, CGGVeritas

Morse, P. and H. Feshbach, 1953, Methods of Theoretical Physics: Cambridge University

v v Downloaded 01/11/16 to Redistribution subject to SEG license or copyright; see Terms of Use at

Traveltime sensitivity kernels: Banana-doughnuts or just plain bananas? a

Vollständige Inversion seismischer Wellenfelder - Erderkundung im oberflächennahen Bereich

Multicomponent imaging with distributed acoustic sensing Ivan Lim Chen Ning & Paul Sava, Center for Wave Phenomena, Colorado School of Mines

Wenyong Pan and Lianjie Huang. Los Alamos National Laboratory, Geophysics Group, MS D452, Los Alamos, NM 87545, USA

Seismic tomography with co-located soft data

Anisotropic inversion and imaging of PP and PS reflection data in the North Sea

2012 SEG SEG Las Vegas 2012 Annual Meeting Page 1

arxiv: v1 [math.na] 1 Apr 2015

A frequency-space 2-D scalar wave extrapolator using extended 25-point finite-difference operator

Seismic wave extrapolation using lowrank symbol approximation

Plane-wave migration in tilted coordinates

Observation of shear-wave splitting from microseismicity induced by hydraulic fracturing: A non-vti story

Youzuo Lin and Lianjie Huang

P191 Bayesian Linearized AVAZ Inversion in HTI Fractured Media

Downloaded 10/15/15 to Redistribution subject to SEG license or copyright; see Terms of Use at

Migration velocity analysis in factorized VTI media

Exact elastic impedance in orthorhombic media

Stanford Exploration Project, Report 123, October 31, 2005, pages 25 49

Elastic least-squares reverse time migration using the energy norm Daniel Rocha & Paul Sava Center for Wave Phenomena, Colorado School of Mines

Modeling seismic wave propagation during fluid injection in a fractured network: Effects of pore fluid pressure on time-lapse seismic signatures

CHARACTERIZATION OF SATURATED POROUS ROCKS WITH OBLIQUELY DIPPING FRACTURES. Jiao Xue and Robert H. Tatham

Inversion of normal moveout for monoclinic media

AVAZ inversion for fracture orientation and intensity: a physical modeling study

Transcription:

GEOPHYSICS, VOL. 78, NO. 5 (SEPTEMBER-OCTOBER 2013); P. WC41 WC50, 22 FIGS. 10.1190/GEO2012-0391.1 Acoustic anisotropic wavefields through perturbation theory Tariq Alkhalifah 1 ABSTRACT Solving the anisotropic acoustic wave equation numerically using finite-difference methods introduces many problems and media restriction requirements, and it rarely contributes to the ability to resolve the anisotropy parameters. Among these restrictions are the inability to handle media with η < 0 and the presence of shear-wave artifacts in the solution. Both limitations do not exist in the solution of the elliptical anisotropic acoustic wave equation. Using perturbation theory in developing the solution of the anisotropic acoustic wave equation allows direct access to the desired limitation-free solutions, that is, solutions perturbed from the elliptical anisotropic background medium. It also provides a platform for parameter estimation because of the ability to isolate the wavefield dependency on the perturbed INTRODUCTION The acoustic anisotropic wave equation is now heavily used in industry as anisotropy inclusion in imaging the subsurface is becoming a necessity. The necessity stems from the fact that anisotropic migrations have recently yielded superior images to those attained using the conventional isotropic assumption of the earth subsurface, even in complex media (Zhou et al., 2004; Huang et al., 2008). This preference for anisotropic imaging is practiced in spite of the many limitations associated with solving the anisotropic acoustic wave equation (Alkhalifah, 2000). These limitations include constraints on the anisotropic models, specifically that η 0. Also, in its conventional implementation, wavefields extracted from the fourth-order anisotropic acoustic wave equation include shear-wave artifacts. It is also a fact that solving the anisotropic acoustic wave equation costs more than solving its isotropic counterpart, and the cost for the tilted anisotropic case could be four times higher. anisotropy parameters. As a result, I derive partial differential equations that relate changes in the wavefield to perturbations in the anisotropy parameters. The solutions of the perturbation equations represented the coefficients of a Taylor-series-type expansion of the wavefield as a function of the perturbed parameter, which is in this case η or the tilt of the symmetry axis. The expansion with respect to the symmetry axis allows use of an acoustic transversely isotropic media with a vertical symmetry axis (VTI) kernel to estimate the background wavefield and the corresponding perturbation coefficients. The VTI extrapolation kernel is about one-fourth the cost of the transversely isotropic model with a tilt in the symmetry axis kernel. Thus, for a small symmetry axis tilt, the cost of migration using a first-order expansion can be reduced. The effectiveness of the approach was demonstrated on the Marmousi model. These limitations pale in comparison with the challenges involved in estimating the anisotropy parameters. It is hard enough to resolve a velocity model that may change vertically and laterally on a fine scale imagine doing so for more than one velocity (or parameter) and specifically resolving up to five parameters in the widely used transversely isotropic (TI) model with a tilt in the symmetry axis (TTI) in 3D. Despite the many approximations used to simplify the anisotropic description, including weak anisotropy approximations of the polarization and phase angles (Pšenčík and Gajewski, 1998), parameter estimation remains difficult in complex media. A dip-oriented TI assumption reduces the required parameters to three (Alkhalifah and Sava, 2010). Nevertheless, resolving even three parameters is at least three times more difficult. The first-order dependence of wavefields, and specifically the acoustic wave equation, on media parameters is described by the Born approximation. It is a single scattering approximation, used in seismic applications to approximate the perturbed wavefield Manuscript received by the Editor 22 September 2012; revised manuscript received 28 March 2013; published online 25 July 2013; corrected version published online 17 September 2013. 1 King Abdulaziz City for Science and Technology (KACST), Astronomy and Geophysical Research Institute, Riyadh, Saudi Arabia. E-mail: tariq.alkhalifah@kaust.edu.sa. 2013 Society of Exploration Geophysicists. All rights reserved. WC41

WC42 Alkhalifah due to small perturbations in the reference medium. It is also, in its inverse form, used to help infer medium parameters from observed wavefields (Cohen and Bleistein, 1977; Panning et al., 2009). In the spirit of the Born approximation, the partial differential equations used here describe the wavefield first- and second-order dependence on perturbations in anisotropy parameters from a background isotropic or elliptically anisotropic medium. This type of formulation allows isolation of the anisotropy influence even in complex media, and thus, it provides the opportunity to measure this influence. In acoustic TI media with a vertical symmetry axis (VTI), Alkhalifah (1998) demonstrated that a new representation in terms of just three parameters is sufficient to solve for the wavefield. These three parameters are the P-wave vertical velocity (or symmetry-axis direction) v v, the normal-moveout velocity for a horizontal reflector pffiffiffiffiffiffiffiffiffiffiffiffiffi v ¼ v v 1 þ 2δ; (1) and the anisotropy coefficient v 2 η 0.5 h v 2 1 ¼ ϵ δ 1 þ 2δ ; (2) where v h is the horizontal velocity and ϵ and δ are Thomsen (1986) parameters. Alkhalifah (2000) introduces the acoustic wave equation as an alternative to the elastic wave equation. It produces a very similar description of the P-wave kinematics including the geometrical amplitude behavior to that of the elastic solution, but at a reduced cost and effort. It is especially useful for imaging applications because we tend to condition the data to comply with the acoustic assumption. Following in the footsteps of Alkhalifah (2000), I apply perturbation theory to the acoustic anisotropic wave equation in hope of alleviating some of the aforementioned weaknesses of the original equation. I focus on perturbing from a background elliptical anisotropic model as solving the wave equation for such a model is free of all the limitations associated with the VTI equation. This implies that I develop solutions as a function of a polynomial series in terms of powers in η, and also the tilt. Despite the constant η and tilt assumption, the solution should be valid for expansion parameter changes causing wavelength perturbations not exceeding the seismic wavelength. I demonstrate the usefulness of the approach on synthetic data, including the Marmousi model. PERTURBATION THEORY The general procedure of the perturbation method is to identify a small parameter (i.e., tilt angle or η), such that when this parameter is set to zero the problem (for example, the differential equation) becomes easily soluble. Thematically, the approach decomposes a tough problem into an infinite number of relatively easy ones. Hence, the perturbation method is most useful when the first few steps reveal the important features of the solution and the remaining ones give small corrections. Another feature of the perturbation method is that it does not add additional solutions to the solutions already obtained from the unperturbed medium (i.e., background medium). The method perturbs those solutions obtained for the background medium (in this case, elliptical anisotropic or VTI) to accommodate the perturbation in the medium. To approach the TTI problem, two routes can be taken. Compute a VTI background wavefield solution and perturb from that solution as a function of tilt angle to approximate the TTI solution. One advantage to this approach is the relatively low cost of the VTI solution, and it allows me to readily invert for a smooth tilt angle field. However, accuracy will depend directly on the size of the tilt angle. Alternatively, I compute a tilted elliptical anisotropic solution and perturb that solution in terms of η to approximate the TTI behavior. A feature of this approach is the simplicity of the elliptical wave equation (second order) with solutions free of shear-wave artifacts. In the following, I develop both approaches. Perturbation in tilt angle For a TTI medium, the derivatives in the acoustic VTI wave equation (Alkhalifah, 2000) are taken with respect to the symmetry axis direction, and thus, I have to rotate the derivatives using the following Jacobian in 3D: 0 1 cos ϕ cos θ sin ϕ cos θ sin θ @ sin ϕ cos ϕ 0 A; (3) cos ϕ sin θ sin ϕ sin θ cos θ to obtain a wave equation corresponding to the conventional computational coordinates governed by the acquisition surface. In equation 3, θ is the angle of the symmetry axis measured from the vertical and ϕ corresponds to the azimuth of the vertical plane that contains the symmetry axis measured from the x-axis (typically the inline direction). The TTI (or VTI) acoustic wave equation, despite its fourth-order nature, can be described by a pair of second-order linear partial differential equations, and specifically I use the form of Zhang et al. (2011) given by 2 p t 2 ¼ v2 h H 2½pŠþv 2 vh 1 ½qŠ; (4) 2 q t 2 ¼ v2 H 2 ½pŠþv 2 vh 1 ½qŠ: (5) Here, p and q are two wavefields that depend on space x and time t and H 1 and H 2 are differential operators applied to the wavefields in square brackets. They include rotation to honor the tilt in the axis of symmetry. In VTI media, H 1 and H 2 reduce to 2 z 2 and ð 2 x 2 Þþð 2 y 2 Þ, respectively. To honor tilt, I rotate these derivatives using the Jacobian (equation 3) (Zhang et al., 2011). The velocities v v, v h, and v are those along the symmetry axis, orthogonal to the symmetry axis and NMO velocities, respectively, used to parametrize a generic TTI medium. Also, the operators H 1 and H 2 depend on medium parameters θ and ϕ, describing the symmetry axis direction. Assuming the symmetry axis tilt, θ, to be small and independent of position (constant), and focusing on the 2D case, ϕ ¼ 0 (x-axis). I consider the following trial solutions to equations 4 and 5: pðx; z; tþ p 0 ðx; z; tþþp 1 ðx; z; tþ sin θ þ p 2 ðx; z; tþsin 2 θ; (6) and

Acoustic anisotropic wavefields WC43 qðx; z; tþ q 0 ðx; z; tþþq 1 ðx; z; tþ sin θ þ q 2 ðx; z; tþsin 2 θ: (7) Setting θ ¼ 0 in equations 4 and 5 allows us to obtain the acoustic wave equation for VTI media, as follows: 2 p 0 t 2 v2 h H 0 2 ½p 0Š v 2 vh 0 1 ½q 0Š¼0; (8) 2 q 0 t 2 v2 H 0 2 ½p 0Š v 2 vh 0 1 ½q 0Š¼0; (9) where H1 0 and H 2 0 are the VTI differential operators, and as described previously H1 0 ¼ H 1ðθ ¼ 0Þ ¼ 2 z 2 and H2 0 ¼ H 2 ðθ ¼ 0Þ ¼ 2 x 2. This will allow me to solve for p 0 and q 0. If I insert the trial solutions 6 and p7, respectively, into equations 4 and 5, and I expand the cos θ ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 sin 2 θ using Taylor s series for small sin θ, I can rearrange the terms for both equations to have the general form P ia i sin i ðθþ ¼0. For this polynomial to equal zero for a nonzero θ, the coefficients a i must equal zero, for i ¼ 0; 1; 2; :::. Thus, a 0 ¼ 0 results in equations 8 and 9. While a 1 ¼ 0 allows the first-order coefficients p 1 and q 1 to satisfy a similar VTI wave equation, but with a source function that depends on the background VTI wavefields (p 0 and q 0 ), as follows: 2 p 1 t 2 v2 h H 0 2 ½p 1Š v 2 vh 0 1 ½q 1Š¼2v 2 h Hxz ½p 0 Š 2v 2 vh xz ½q 0 Š; (10) 2 p 1 t 2 v2 H 0 2 ½p 1Š v 2 vh 0 1 ½q 1Š¼2v 2 H xz ½p 0 Š 2v 2 vh xz ½q 0 Š; (11) where H xz ¼ 2 x z. The second-order coefficients p 2 and q 2 also satisfy the same VTI wave equation, extracted from a 2 ¼ 0, as follows: 2 p 2 t 2 2 q 2 t 2 v2 h H 0 2 ½p 2Š v 2 vh 0 1 ½q 2Š¼v 2 h ½2Hxz ½p 1 ŠþH 0 1 ½p 0Š H2 0½p 0ŠŠ v 2 v½2h xz ½q 1 Š þ H1 0½q 0Š H2 0½q 0ŠŠ; (12) v2 H 0 2 ½p 2Š v 2 vh 0 1 ½q 2Š¼v 2 ½2H xz ½p 1 ŠþH 0 1 ½p 0Š H2 0½p 0ŠŠ v 2 v½2h xz ½q 1 Š þ H1 0½q 0Š H2 0½q 0ŠŠ: (13) As I go to higher order, it is clear that the coefficients will satisfy the same VTI wave equation with more complicated source functions. Thus, accuracy will cost more. For 3D media, I will need to solve equations 4 and 5 for VTI media, with source functions that include additional derivatives in y that depend on the order of the desired coefficients for the expansion. Perturbation in η In this case, I consider the following trial solutions to the TTI wave equations 4 and 5: and pðx; z; tþ p 0 ðx; z; tþþp 1 ðx; z; tþη þ p 2 ðx; z; tþη 2 ; (14) qðx; z; tþ q 0 ðx; z; tþþq 1 ðx; z; tþη þ q 2 ðx; z; tþη 2 : (15) Setting η ¼ 0 in equations 4 and 5 allows me to obtain the acoustic wave equation for tilted elliptical anisotropic media, as follows: 2 p 0 t 2 v2 H 2 ½p 0 Š v 2 vh 1 ½q 0 Š¼0; (16) 2 q 0 t 2 v2 H 2 ½p 0 Š v 2 vh 1 ½q 0 Š¼0; (17) where H 1 and H 2 are the TTI operators described above. Because the two equations are similar, q 0 ¼ p 0 and satisfy 2 p 0 t 2 v2 H 2 ½p 0 Š v 2 vh 1 ½p 0 Š¼0: (18) The first-order coefficients p 1 and q 1 satisfy a similar tilted elliptically anisotropic wave equation, but with a source function that depends on the background elliptically anisotropic wavefield (p 0 and q 0 ), as follows: 2 p 1 t 2 v2 H 2 ½p 1 Š v 2 vh 1 ½q 1 Š¼2v 2 H 2 ½p 0 Š; (19) 2 q 1 t 2 v2 H 2 ½p 1 Š v 2 vh 1 ½q 1 Š¼0: (20) The second-order coefficients, p 2 and q 2, also satisfy the same TI wave equation as follows: 2 p 2 t 2 v2 H 2 ½p 2 Š v 2 vh 1 ½q 2 Š¼2v 2 H 2 ½p 1 Š; (21) 2 q 2 t 2 v2 H 2 ½p 2 Š v 2 vh 1 ½q 2 Š¼0: (22) This formulation also allows me to constrain the symmetry axis to be normal to the reflector dip and create a mechanism to invert for η using equations 14 and 15. Again, higher-order coefficients satisfy the same second-order tilted elliptical anisotropic wave equation with a source function dependent on the lower order coefficient solutions. For the η perturbation, the complexity of this source function does not increase as I solve for higher order coefficients.

WC44 Alkhalifah NUMERICAL EXAMPLES The accuracy of the perturbation-based solutions depends mainly on the amount of perturbation in the parameters. In the numerical tests, I will solve the TTI wave equation by either perturbing from a background VTI medium, and thus the perturbation parameter is the tilt angle, or perturbing from a background tilted elliptical anisotropic medium, with the perturbation parameter η. I compare both perturbation solutions with the exact acoustic solution. The finitedifference approximation for all the wave equations is second order in time and fourth order in space. I start with a homogeneous example. Homogeneous medium I consider wave propagation in an acoustic TTI media with velocity equal to 2km s, δ ¼ 0.05, and η ¼ 0.05. The source, given by a Ricker wavelet with a peak frequency of 15 Hz, is set in the middle of the propagation medium. The space sampling interval in x and z is 0.008 km. Snapshots of the wavefields at 0.7 s for a tilt angle of 10 are shown in Figure 1. The difference between the exact acoustic solution shown in Figure 1a and the one based on the one-term perturbation expansion (Figure 1b) is generally small (Figure 1c). This observation, of course, depends on the order of the expansion, but more so on the amount of tilt and the strength of anisotropy. Figure 2 shows snapshots of the wavefield obtained from the first-order perturbation for three different tilt angles: 10, 20, and 30. The kinematic errors clearly increase with tilt angle as I compare the snapshots with the exact analytically derived traveltime solution (dashed curve). The difference between these wavefields and the exact acoustic TTI wave equation is shown in Figure 3 for the three tilt angles. The errors increase with the larger perturbation given by the larger tilt angle from vertical. In fact, the magnitude of the error for the 30 tilt is comparable in size to the original wavefield. However, most of these errors, as the kinematics demonstrated, are amplitude related. Also, note the presence of the numerical dispersion in the snapshots of the solution. This is a result of the relatively coarse-spaced grid of 0.008 km used in the finite-difference implementation. On the other hand, focusing on perturbations in η from an elliptical TTI medium background as an alternative to the tilt perturbation allows me to focus on the η influence. Figure 4 compares the exact TTI solution to that perturbed from a background tilted elliptical anisotropic medium using only the first-order approximation in equations 14 and 15. The difference, shown in Figure 4c, exposes a large angular dependence compared with the errors associated with the tilt perturbation case. This is largely because η has a very strong angular influence on velocity. Figure 5 shows the first-order perturbation solution in η for three different tilts (10, 20, and 30 ) along with the exact traveltime solution given by the dashed curves. The differences with the exact solution (errors), shown in Figure 6 for the three symmetry direction angles, are very similar in size and do not depend on tilt angle. This is expected because the perturbation is in η. In both cases, I can use the second-order expansion for higher accuracy, but at an increase in computational cost. Figure 7 shows the same difference plots shown in Figure 3, but with the secondorder accuracy perturbation. As expected, the differences (errors) are smaller. A similar conclusion is drawn in perturbing from an elliptical anisotropic background, as Figure 8 shows an overall reduction in error compared to that shown in Figure 6. Figure 1. A snapshot of the wavefields at time 0.7 s for a source in the center (a) computed using the actual acoustic TTI wave equation, (b) computed using the first-order perturbation in tilt angle (middle), and (c) the difference for a medium with v ¼ 2km s, η ¼ 0.05, δ ¼ 0.05, and tilt angle of 10. Figure 2. A snapshot of the wavefields at time 0.7 s for a source in the center computed using the first-order perturbation in tilt angle for a tilt angle of (a) 10, (b) 20, and (c) 30 for a medium with v ¼ 2km s, η ¼ 0.05, and δ ¼ 0.05. The dashed curves correspond to the exact analytically derived traveltime solution.

Acoustic anisotropic wavefields WC45 Inhomogeneous medium The interface geometry embedded in the Marmousi model (Figure 9) is based, somewhat, on a profile through the Cuanza basin (Versteeg, 1993). The model contains many reflectors, steep dips, and strong velocity variations in the lateral and the vertical directions (with a minimum velocity of 1500 m s and a maximum of 5500 m s). The point source considered here is a Ricker wavelet with a 25-Hz peak frequency. I initially consider a constant η of 0.05. Figure 3. A snapshot of the wavefields difference between that computed using the first-order perturbation in tilt angle and the exact acoustic solution for a tilt angle of (a) 10, (b) 20, and (c) 30 for a medium with v ¼ 2km s, η ¼ 0.05, and δ ¼ 0.05. Figure 4. A snapshot of the wavefields at time 0.7 s for a source in the center (a) computed using the actual acoustic TTI wave equation, (b) computed using the first-order perturbation in η, and (c) the difference for a medium with v ¼ 2km s, η ¼ 0.05, δ ¼ 0.05, and tilt angle of 10. Figure 5. A snapshot of the wavefields at time 0.7 s for a source in the center computed using the first-order perturbation in η for a tilt angle of (a) 10, (b) 20, and (c) 30 for a medium with v ¼ 2km s, η ¼ 0.05, and δ ¼ 0.05. The dashed curves correspond to the exact analytically derived traveltime solution. Figure 6. A snapshot of the wavefield difference between that computed using the first-order perturbation in η and the exact acoustic solution for a tilt angle of (a) 10, (b) 20, and (c) 30 for a medium with v ¼ 2km s, η ¼ 0.05, and δ ¼ 0.05.

WC46 Alkhalifah Again, using the fourth-order finite-difference approximation in space and second order in time, I solve the wave equation for a source located at the surface at lateral position 5 km. The space sampling interval in x and z is 0.012 km. A snapshot of the wavefields at 1.2 s for a tilt angle of 10 is shown in Figure 10. The difference between the exact acoustic solution shown in Figure 10a and the one based on one term of the perturbation expansion in tilt angle (Figure 10b) is generally small (Figure 10c). This observation, of course, depends on the order of the expansion, but more so on the amount of tilt and the strength of anisotropy. Figure 11 shows snapshots of the wavefield obtained from the first-order perturbation in tilt angle for three different tilt angles: 10, 20, and 30. The difference between these wavefields and the exact acoustic TTI wave equation is shown in Figure 12 for the three tilt angles. The errors increase with the larger perturbation given by the larger tilt. On the other hand, focusing on perturbations in η, Figure 13 compares the exact TTI solution to that perturbed from a background tilted elliptical anisotropic medium using only the first-order approximation in equations 14 and 15. The difference, shown in Figure 13c, exposes a large angular dependency compared to the tilt perturbation. This is largely because η has a very strong angular influence on velocity. Figure 14 shows the first-order perturbation solution in η for three different tilts (10, 20, and 30 ). The differences with the exact solution (errors) shown in Figure 22 for the three symmetry direction angles are very similar and do not depend on tilt angle. Figure 15 shows the wavefield difference between that computed using the first-order perturbation in η and the exact acoustic solution for a tilt angle of (a) 10, (b) 20, and (c) 30 for a medium with η ¼ 0.05 and δ ¼ 0.05. In both cases, the second-order expansion can be used for higher accuracy, but at an increase in the computational cost. Figure 16 shows the same difference plots shown in Figure 12, but with the second-order accuracy perturbation. As expected, the difference (errors) are smaller. A similar conclusion is drawn in perturbing from an elliptical anisotropic background, as Figure 17 shows an overall reduction in error compared to that shown in Figure 22. Again, all images are displayed at the same scale. The elliptically based perturbation has a more angular dependent error than the tilt angle perturbation. Complex η spatial distribution The approximate solution based on a perturbation in η requires theoretically that η be constant. Likewise, the solution corresponding to perturbations in the symmetry axis angle requires that the symmetry angle be constant. In both cases, smooth variations are tolerable, especially if such smooth variations inject an almost stationary behavior within the seismic predominant wavelength. However, a perturbation in η does not set any restrictions on the complexity of the symmetry axis field, and similarly a perturbation in symmetry axis does not require any restrictions on the η field. Alkhalifah (2000) developed a complex η field for the Marmousi model. Figure 18 shows this η with values ranging between 0 and 0.27. I use this model without smoothing to test the perturbation in the angle of symmetry. The model parameters as well as the frequency of the source are similar to that used in the previous subsection. The only difference is the complex η model (Figure 18). Thus, Figure 19 shows that for a 10 symmetry axis tilt, the errors for the first-order perturbation are comparable to those experienced earlier. As the tilt increases, the errors increase (see Figure 20). Nevertheless, these errors are reasonable for inversion applications because we may want to search for the best tilt that fits the data. Figure 7. A snapshot of the wavefield difference between that computed using the second-order perturbation in tilt angle and the exact acoustic solution for a tilt angle of (a) 10, (b) 20, and (c) 30 for a medium with v ¼ 2km s, η ¼ 0.05, and δ ¼ 0.05. Figure 8. A snapshot of the wavefield difference between that computed using the second-order perturbation in η and the exact acoustic solution for a tilt angle of (a) 10, (b) 20, and (c) 30 for a medium with v ¼ 2km s, η ¼ 0.05, and δ ¼ 0.05.

Acoustic anisotropic wavefields WC47 Figure 9. The Marmousi velocity model. To test how much smoothness is needed in the complex η model in Figure 18 for it to provide acceptable approximate solutions (artifact free) for the perturbations in η, I convolve this complex η model with two different-length pyramid-shaped smoothing operators (triangle along each axis), as shown in Figure 21. This is equivalent to multiplying the η model in the Fourier domain by a 2D squared sinc function, which acts as a low-pass filter. Figure 22 shows the errors in the first-order perturbation in η solutions for the Figure 10. A snapshot of the wavefields for a source in the center at time 1.2 s (a) computed using the actual acoustic TTI wave equation, (b) computed using the first-order perturbation in tilt angle, and (c) the difference for a medium with η ¼ 0.05, δ ¼ 0.05, and tilt angle of 10. Figure 11. A snapshot of the wavefields for a source in the center at time 1.2 s computed using the first-order perturbation in tilt angle for a tilt angle of (a) 10, (b) 20, and (c) 30 for a medium with η ¼ 0.05 and δ ¼ 0.05. Figure 12. A snapshot of the wavefield difference between that computed using the first-order perturbation in tilt angle and the exact acoustic solution for a tilt angle of (a) 10, (b) 20, and (c) 30 for a medium with η ¼ 0.05 and δ ¼ 0.05. Figure 13. A snapshot of the wavefields at time 1.2 s for a source in the center (a) computed using the actual acoustic TTI wave equation, (b) computed using the first-order perturbation in η, and (c) the difference for a medium with η ¼ 0.05, δ ¼ 0.05, and tilt angle of 10.

Alkhalifah WC48 two smooth η models using a smoothing operator with a length of 3 samples (Figure 22b), 10 samples (Figure 22a), compared to no smoothing (Figure 22c). Because the difference is with respect to the exact acoustic TTI operator with the same smoothed models, all three difference plots show practically the same magnitude of error. However, for the model with no smoothing applied to a) b) η (Figure 22a), there are slightly more errors and possibly some artifacts, specifically around lateral location 4 km and depth around 2 km, as well as in other places. Such artifacts are far less observable as I smooth the η model (recall that I am perturbing in η), and ultimately no potential artifacts are expected for a constant η model. c) Figure 14. A snapshot of the wavefields at time 1.2 s for a source in the center computed using the first-order perturbation in η for a tilt angle of (a) 10, (b) 20, and (c) 30 for a medium with η ¼ 0.05 and δ ¼ 0.05. a) b) c) Figure 15. A snapshot of the wavefield difference between that computed using the first-order perturbation in η and the exact acoustic solution for a tilt angle of (a) 10, (b) 20, and (c) 30 for a medium with η ¼ 0.05 and δ ¼ 0.05. a) b) c) Figure 16. A snapshot of the wavefield difference between that computed using the second-order perturbation in tilt angle and the exact acoustic solution for a tilt angle of (a) 10, (b) 20, and (c) 30 for a medium with η ¼ 0.05 and δ ¼ 0.05. a) b) c) Figure 17. A snapshot of the wavefield difference between that computed using the second-order perturbation in η and the exact acoustic solution for a tilt angle of (a) 10, (b) 20, and (c) 30 for a medium with η ¼ 0.05 and δ ¼ 0.05.

Acoustic anisotropic wavefields WC49 DISCUSSION Wavefield perturbation allows us to approximately isolate the parts of the wavefield attributable to the perturbed parameters. In this case, the perturbed parameters are the anisotropic parameters and specifically η and the symmetry axis direction. Both parameters tend to be relatively small, but they are usually difficult to estimate Figure 18. The η field for the Marmousi model. using conventional methods. Using perturbation theory, I use the wavefield in its full inhomogeneous capable glory to relate changes in it to changes in these medium parameters. It has a Born approximation framework, which is useful for waveform inversion. This framework provides an efficient way to update the wavefield to reflect the anisotropic influence; however, because of the nature of the perturbation, it requires that the perturbation parameters be constant at least over the maximum wavelength. It is also incapable of dealing with large perturbations in medium parameters specifically, those inducing wavefield perturbations beyond the capability of the polynomial expansion in representing the sinusoidal nature of the wavefield. The length associated with such a limitation is governed by the highest frequency. The perturbation of the acoustic wavefield, to take into account the anisotropy influence of the medium, allows for avoidance of many of the limitations associated with obtaining such wavefields Figure 19. A snapshot of the wavefields at time 1.2 s for a source in the center (a) computed using the actual acoustic TTI wave equation, (b) computed using the first-order perturbation in tilt angle, and (c) the difference for a medium with δ ¼ 0.0, tilt angle of 10, and η given by Figure 18. Figure 20. A snapshot of the wavefield difference between that computed using the first-order perturbation in tilt angle and the exact acoustic solution for a tilt angle of (a) 10, (b) 20, and (c) 30 for a medium δ ¼ 0.0, and η given by Figure 18. Figure 21. (a) The η model in Figure 18, (b) smoothed using a triangulated filter with length of 0.075 km in each direction, and (c) 0.25 km in each direction.

WC50 Alkhalifah Figure 22. A snapshot of the wavefield difference between that computed using the first-order perturbation in η and the exact acoustic solution for the smoothed η models in Figure 21, respectively. in the conventional way. The η perturbation of the wavefield from a background elliptical anisotropic (vertical or tilted) medium are free of shear-wave artifacts. Unlike the original equation, it also works for negative η. The perturbation with respect to the tilt angle allows us to rely on a relatively efficient (far more efficient than the TTI) VTI engine for calculating the background solutions and the perturbation coefficients. The accuracy of the wavefield perturbation solution also depends on the frequency content because lower frequencies yield lower perturbation errors, especially because the perturbation, unlike the classical Born type, is based on a complex background and smooth variation in the perturbed parameter. Thus, it provides a platform for parameter estimation in line with the needs associated with full waveform inversion governed by the phase component of the wavefield. The perturbation in tilt angle yields an overall uniform distribution of error with angle mostly in amplitude. On the other hand, a perturbation in η has larger errors in the direction normal to tilt angle, where the influence of η is the largest. CONCLUSIONS Using perturbation theory, I derive partial differential equations to solve for the expansion coefficients of the acoustic anisotropic wavefield solution in terms of η and the symmetry axis. Such coefficients provide insights into the sensitivity of the acoustic wavefield to such anisotropy parameters. They also provide solutions free of some of the limitations associated with the original TTI acoustic wave equation. Though the synthetic tests on homogenous and inhomogeneous media showed higher accuracy for the η perturbation, the perturbation in η and symmetry axis tilt provided generally accurate wavefields. ACKNOWLEDGMENTS I thank KACST for their support. I thank the associate editor, S. Hestholm, and two anonymous reviewers for their valuable review of the paper. REFERENCES Alkhalifah, T., 1998, Acoustic approximations for processing in transversely isotropic media: Geophysics, 63, 623 631, doi: 10.1190/1.1444361. Alkhalifah, T., 2000, An acoustic wave equation for anisotropic media: Geophysics, 65, 1239 1250, doi: 10.1190/1.1444815. Alkhalifah, T., and P. Sava, 2010, A transversely isotropic medium with a tilted symmetry axis normal to the reflector: Geophysics, 75, no. 3, A19 A24, doi: 10.1190/1.3409114. Cohen, J., and N. Bleistein, 1977, Seismic waveform modelling in a 3-D earth using the Born approximation: Potential shortcomings and a remedy: Journal of Applied Mathematics, 32, 784 799. Huang, T., S. Xu, J. Wang, G. Ionescu, and M. Richardson, 2008, The benefit of TTI tomography for dual azimuth data in Gulf of Mexico: 78th Annual International Meeting, SEG, Expanded Abstracts, 222 226. Panning, M., Y. Capdeville, and A. Romanowicz, 2009, An inverse method for determining small variations in propagation speed: Geophysical Journal International, 177, 161 178, doi: 10.1111/j.1365-246X.2008.04050.x. Pšenčík, I., and D. Gajewski, 1998, Polarization, phase velocity, and NMO velocity of qp-waves in arbitrary weakly anisotropic media: Geophysics, 63, 1754 1766, doi: 10.1190/1.1444470. Thomsen, L., 1986, Weak elastic anisotropy: Geophysics, 51, 1954 1966, doi: 10.1190/1.1442051. Versteeg, R. J., 1993, Sensitivity of prestack depth migration to the velocity model: Geophysics, 58, 873 882, doi: 10.1190/1.1443471. Zhang, Y., H. Zhang, and G. Zhang, 2011, A stable TTI reverse time migration and its implementation: Geophysics, 76, no. 3, WA3 WA11, doi: 10.1190/1.3554411. Zhou, H., D. Pham, S. Gray, and B. Wang, 2004, Tomographic velocity analysis in strongly anisotropic TTI media: 74th Annual International Meeting, SEG, Expanded Abstracts, 2347 2350.