MULTISCALE GAUSSIAN-BEAM METHOD FOR HIGH-FREQUENCY WAVE PROPAGATION AND INVERSE PROBLEMS By Chao Song A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Applied Mathematics — Doctor of Philosophy 2018 ABSTRACT MULTISCALE GAUSSIAN-BEAM METHOD FOR HIGH-FREQUENCY WAVE PROPAGATION AND INVERSE PROBLEMS By Chao Song The existence of Gaussian beam solution to hyperbolic PDEs has been known to the pure mathematics community since sometime in the 1960s [3]. It enjoys popularity afterwards due to its ability to resolve the caustics problem and its efficiency [49, 28, 31]. In this thesis, we will focus on the extension of the multi-scale Gaussian beam method and its application to seismic wave modeling and inversion. In the first part of thesis, we discuss the application of the multi-scale Gaussian beam method to the inverse problem. A new multi-scale Gaussian beam method is introduced for carrying out true-amplitude prestack migration of acoustic waves. After applying the Born approximation, the migration process is considered as shooting two beams simultaneously from the subsurface point which we want to image. The Multi-scale Gaussian Wavepacket transform provides an efficient and accurate way for both decomposing the perturbation field and initializing Gaussian beam solution. Moreover, we can prescribe both the region of imaging and the range of dipping angles by shooting beams from a subsurface point in the region of imaging. We prove the imaging condition equation rigorously and conduct error analysis. Some numerical approximations are derived to improve the efficiency further. Numerical results in the two-dimensional space demonstrate the performance of the proposed migration algorithm. In the second part of thesis, we propose a new multiscale Gaussian beam method with reinitialization to solve the elastic wave equation in the high frequency regime with different boundary conditions. A novel multiscale transform is proposed to decompose any arbitrary vector-valued function to multiple Gaussian wavepackets with various resolution. After the step of initializing, we derive various rules corresponding to different types of reflection cases. To improve the efficiency and accuracy, we develop a new reinitialization strategy based on the stationary phase approximation method to sharpen each single beam ansatz. This is especially useful and necessary in some reflection cases. Numerical examples with various parameters demonstrate the correctness and robustness of the whole method. There are two boundary conditions considered here, the periodic and the Dirichlet boundary condition. In the end, we show that the convergence rate of the proposed multiscale Gaussian beam method follows the convergence rate of the classical Gaussian beam solution, i.e. O(cid:0) 1√ (cid:1). ω To my family iv ACKNOWLEDGMENTS Firstly, I would like to express my sincere gratitude to my advisor Prof. Qian, Jianliang, for his continuous support of my Ph.D. study and related research, for his patience, moti- vation, and immense knowledge. His guidance helped me in all the time of research and writing of this thesis. I could not have imagined having a better advisor for my Ph.D. study. I would like to express my gratitude to Prof. Zhou, Zhengfang, Prof. Cheng, Yingda and Prof. Tang, Moxun for serving on my thesis committee and critical comments and valuable feedbacks. I also would like to thank all the professors whom I took classes from at MSU, for their uniform excellence in teaching. I would like to thank the support staff of Department of Mathematics for all their help with administrative matters. I would like to acknowledge my dear friends Ruochuan Zhang, Jun Du, Wangtao Lu and Wenbin Li for their friendship over the past six years. They made the experience of graduate school more rewarding. Many thanks go to my parents for supporting me spiritually throughout writing this thesis and my life in general. Last but not the least, I would like to express my profound gratitude to my wife Yu, who has shown me the deepest love, kindness, and patience throughout our time together. This dissertation stands as a testament to your unconditional love and encouragement. v TABLE OF CONTENTS LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix LIST OF ALGORITHMS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xii Chapter 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1 Prestack Inversion Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2 Elastic Wave Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3 Related Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.4 Contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 3 5 7 10 Chapter 2 Single Gaussian beam ansatz and Multiscale Gaussian wavepacket Transform . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1 Single Gaussian beam ansatz . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 Multiscale Gaussian Wavepacket Transform . . . . . . . . . . . . . . . . . . 2.2.1 Multiscale Gaussian Beam Method . . . . . . . . . . . . . . . . . . . 12 12 16 19 Migration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1 Modified Multiscale Gaussian Wavepacket Transform . . . . . . . . . . . . . 3.2 Multiscale Gaussian Wavepacket Inversion . . . . . . . . . . . . . . . . . . . 3.2.1 Setup of the True Amplitude Migration Problem . . . . . . . . . . . . 3.2.2 Born Approximation for the Trace Data . . . . . . . . . . . . . . . . 3.2.3 Multiscale Gaussian Beam Approximation of the Green’s Function . . 3.2.4 True-Amplitude Migration Process . . . . . . . . . . . . . . . . . . . 3.2.5 Motivation for Inverting the multiscale Gaussian-beam transform . . 3.3 Theoretical Validation: The Proof about The Imaging Operator . . . . . . . 3.3.1 Road Map of The Theoretical Analysis . . . . . . . . . . . . . . . . . 3.3.2 Approximation of Gaussian Beams along the Surface . . . . . . . . . 3.3.3 Asymptotic Analysis of Two Beams’ Interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chapter 3 Multiscale Gaussian-Beam Transforms for True Amplitude Prestack 21 21 29 29 31 33 34 37 39 39 42 47 3.3.3.1 Parabolic Scaling Principle 48 3.3.3.2 Difference between Two Interacted Beams’ Traveltime 51 3.3.3.3 Difference between Two Interacted Beams’ Phase and Hessians 55 58 3.3.3.4 Approximation of Two Beams’ Interaction . . . . . . . . . . 3.3.3.5 Integral about Boundary Points r and s . . . . . . . . . . . 61 63 3.3.3.6 Conclusion of Two Beams’ Interaction . . . . . . . . . . . . 66 3.3.4 Asymptotic Analysis of Four Beams’ Interaction . . . . . . . . . . . . Integral about Wavenumber ω . . . . . . . . . . . . . . . . . 66 Integral about Momentum ξ and η: Evaluation of Real Part of Phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3.4.1 3.3.4.2 68 vi 3.3.4.3 3.3.4.4 Integral about Momentum ξ and η: Evaluation of Imaginary Part of Phase . . . . . . . . . . . . . . . . . . . . . . . . . . Integral about Momentum ξ and η . . . . . . . . . . . . . . Implementation of the Prestack Imaging Operator . . . . . . . . . . . . . . . 3.4 3.5 Numerical Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.5.1 Approximation of Beams along the surface . . . . . . . . . . . . . . . 3.5.2 The Correctness of Prestack Imaging Operator . . . . . . . . . . . . . 3.5.3 Single Source Migration Test . . . . . . . . . . . . . . . . . . . . . . . 3.5.3.1 Example 1: Constant Background Slowness . . . . . . . . . 3.5.3.2 Example 2: Multiple Flat Reflectors . . . . . . . . . . . . . 3.5.3.3 Example 3: Linear Background Slowness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.5.4.1 Example 4: Constant Slowness with Dipped Layer . . . . . 3.5.4.2 Example 5: Flat Layer in Lateral Background Velocity . . . 3.5.4.3 Example 6: Slowness with Caustics . . . . . . . . . . . . . 3.5.4.4 Example 7: Polluted Trace Data . . . . . . . . . . . . . . . 3.5.4 Multiple Source Migration Test 71 76 81 85 85 86 87 87 88 89 89 89 91 92 92 Chapter 4 Fast Multiscale Gaussian Beam Method for Elastic Wave Equa- 95 tions in Bounded Domains . . . . . . . . . . . . . . . . . . . . . . 95 4.1 Asymptotic Method for the Elastic Wave equation . . . . . . . . . . . . . . . 96 4.2 The Asymptotic Ansatz Solution to the Elastic Wave . . . . . . . . . . . . . 98 4.2.1 P-wave and S-wave’s Eikonal Equations . . . . . . . . . . . . . . . . . 4.2.2 Transport Equation Governing P-wave’s Amplitude Vectors . . . . . 101 4.2.3 Transport Equation Governing S-wave’s Amplitude Vectors . . . . . . 103 4.2.4 Single Beam Solution for P and S-wave . . . . . . . . . . . . . . . . . 105 . . . . . . . . 107 4.3.1 Multiscale Gaussian Wavepacket Transform: Vector Functions . . . . 107 4.3.1.1 Decomposition of the Single Wavepacket . . . . . . . . . . . 108 4.3.1.2 Preprocessing the Initial Condition . . . . . . . . . . . . . . 109 4.4 Multiscale Gaussian Beam Method for Periodic Boundary Value Problem . . 111 4.5 Multiscale Gaussian beam method for Homogeneous Dirichlet Boundary Con- 4.3 Multiscale Gaussian Wavepacket Transform for Elastic Waves dition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112 4.5.1 P-wave Reflecting Beams: Ray Direction . . . . . . . . . . . . . . . . 112 4.5.2 P-wave Reflecting Beams: The Hessian of the Phase . . . . . . . . . . 114 4.5.3 P-wave Reflecting Beams: Amplitude Vector . . . . . . . . . . . . . . 116 4.5.4 S-wave Reflecting Beams: the Phase term . . . . . . . . . . . . . . . 117 S-wave Reflecting Beams: the Amplitude Vector . . . . . . . . . . . . 119 4.5.5 4.5.6 Method of Images for Boundary Conditions . . . . . . . . . . . . . . 120 4.6 Stationary Phase Analysis of Beams . . . . . . . . . . . . . . . . . . . . . . . 122 Stationary Phase Approximation with Respect to Spatial Variables . 123 Stationary Phase Approximation with Respect to Momentum Variables 126 4.7 Sharpening Beams by Reinitialization . . . . . . . . . . . . . . . . . . . . . . 129 4.7.1 The First Motivation for Developing a New Reinitialization Strategy 129 4.7.2 The Second Motivation for Developing a New Reinitialization Strategy 130 4.7.3 . . . . . . . . . . 133 Sharpened Wavepackets and Convergence Analysis 4.6.1 4.6.2 vii 4.8 Numerical Examples 4.8.1.1 4.8.1.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139 4.8.1 Beam Reinitialization . . . . . . . . . . . . . . . . . . . . . . . . . . . 139 SP Reflection V.S. PS Reflection . . . . . . . . . . . . . . . 139 Sharpened Beams V.S. Original Beams . . . . . . . . . . . . 141 4.8.2 Periodic Boundary Condition . . . . . . . . . . . . . . . . . . . . . . 142 4.8.2.1 Example 1: The Single Wavepacket in the Constant velocity 142 4.8.2.2 Example 2: General Initial Condition in the Constant velocity142 4.8.3 Reflection: P-wave . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145 4.8.3.1 Example 3: P-wave Reflection in the Constant velocity . . . 145 4.8.3.2 Example 4: P-wave Reflection in the Linear velocity . . . . 146 4.8.3.3 Example 5: P-wave Reflection in the Sinusoidal velocity . . 146 4.8.4 Reflection S-wave . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147 4.8.4.1 Example 6: S-wave Reflection with Orthogonal Hitting Angle 147 4.8.4.2 Example 7: S-wave Reflection with Non-Orthogonal Hitting Angle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149 4.8.4.3 Example 8: S-wave Reflection: Linear velocity . . . . . . . . 152 4.8.4.4 Example 9: S-wave Reflection: Sinusoidal velocity . . . . . . 153 4.8.5 General Initial Condition . . . . . . . . . . . . . . . . . . . . . . . . . 154 4.8.5.1 Example 10: The General Initial Condition in The Sinusoidal 4.8.5.2 Convergence Rate Analysis velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155 . . . . . . . . . . . . . . . . . . 157 Chapter 5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159 APPENDICES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161 Appendix A Proof in Inverse Process . . . . . . . . . . . . . . . . . . . . . . . . . 162 Appendix B Proof in Elastic Wave . . . . . . . . . . . . . . . . . . . . . . . . . . 193 BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201 viii LIST OF FIGURES Figure 1.1: Caustics Appearance in Gaussian Velocity . . . . . . . . . . . . . . . . . 2 Figure 3.1: (cid:80) l,i g2 l,i in different level l . . . . . . . . . . . . . . . . . . . . . . . . . . 29 Figure 3.2: A typical source gather in Gaussian slowness . . . . . . . . . . . . . . . . 30 Figure 3.3: Real Part of Single Wavepacket φl,i,k . . . . . . . . . . . . . . . . . . . . 49 Figure 3.4: Numerical test for fast approximation of Gaussian beam along the surface. Left: ω = -110 Right: ω = -100. . . . . . . . . . . . . . . . . . . . . . . . Figure 3.5: Constant Slowness: Theorem 3.3.1. Left: dx1 = dx2 Right: dx1 = 2dx2. Figure 3.6: General Speed: Theorem 3.3.1. Left: dx1 = dx2 Right: dx1 = 2dx2. . . Figure 3.7: Example 1: Constant Slowness with Dipped layer. Left: True Slowness . . . . . . . . . . . . . Right: Migration Result with single source trace. 85 86 87 88 Figure 3.8: Example 2: True Slowness Model with Multiple Layers . . . . . . . . . 88 Figure 3.9: Example 2: Constant Slowness with Multiple Layers. Left: Migration Result over the Whole Space Right: Migration Result V.S. True Value at x1 = 0. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 3.10: Example 3: Gradient Slowness Model. Left: True Slowness, Right: Smoothed Macro Slowness . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 3.11: Example 3: Gradient Slowness Model. Left: Migration Result over the . . . . Whole Space, Right: Migration Result V.S. True Value at x1 = 0. 89 90 90 Figure 3.12: Example 4: Constant Slowness with the Dipped Layer (Multiple Sources) 91 Figure 3.13: Example 5: Flat Layer in Lateral Background Velocity . . . . . . . . . 91 Figure 3.14: Example 6: Gaussian Slowness and its ray tracing. Left: Gaussian . . . . . . . . . . . . . Slowness with Flat Reflector Right: Ray Tracing Figure 3.15: Example 6: Migration Result in the Gaussian Slowness with Caustics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (Multiple Sources) 92 93 ix Figure 3.16: Example 7: True Trace V.S. Trace with Gaussian Error . . . . . . . . . 93 Figure 3.17: Example 7: Gaussian Slowness with polluted trace. Left: Migration result from Non-polluted Data; Right: Migration result from Polluted data 94 Figure 3.18: Example 7: Gaussian Slowness with polluted trace.Two Migration Re- sults at x2 = 0.65 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94 Figure 4.1: Periodic Boundary Problem: The Case Wavepacket leaving the Boundary. 111 Figure 4.2: Partially Reflected beams . . . . . . . . . . . . . . . . . . . . . . . . . . 121 Figure 4.3: Partially Reflected Beams with Odd Extension . . . . . . . . . . . . . . . 121 Figure 4.4: PS Reflection V.S. SP Reflection: Different Ratio Behaviors. PS Reflec- tion(’o’), SP Reflection (’-’) . . . . . . . . . . . . . . . . . . . . . . . . . 140 Figure 4.5: Sharpened Beams V.S. the Original Beam on Fixed (y, z). Left: y=0.1, z=0.1 Right: y=0.08, z=0.13. Beam Solution after Reinitialization(’o’), Original Beam Solution (’-’). . . . . . . . . . . . . . . . . . . . . . . . . 141 Figure 4.6: Example 1: Single Wavepacket Propagation with the Periodic Boundary Condition. Left: FDTD Solution Right: Gaussian Beam Solution. . . 143 Figure 4.7: Example 2: General Initial Value Propagation with Periodic Boundary Condition along x-axis FDTD Solution(’o’), Gaussian Beam Solution (’-’) 144 Figure 4.8: Example 2: General Initial Value Propagation with Periodic Boundary Condition. Left: FDTD Solution Right: Gaussian Beam Solution. . . 144 Figure 4.9: Example 3: P-wave Reflection in Constant velocity with Dirichlet Bound- ary Condition. Left: FDTD Solution Right: Gaussian Beam Solution. 146 Figure 4.10: Example 4: P-wave Reflection in the Linear velocity with Dirichlet Boundary Condition. FDTD Solution (’o’), Beam Solution(’-’) . . . . . . 147 Figure 4.11: Example 4: P-wave Reflection in Linear velocity with Dirichlet Bound- ary Condition. Left: FDTD Solution Right: Gaussian Beam Solution. 148 Figure 4.12: Example 5: P-wave Reflection in Sinusoidal velocity along x-axis. FDTD Solution(’o’), Beam Solution (’-’) . . . . . . . . . . . . . . . . . . . . . . 148 Figure 4.13: Example 5:P-wave Reflection in Sinusoidal velocity with Dirichlet Bound- ary Condition. Left: FDTD Solution, Right: Gaussian beam Solution . 149 x Figure 4.14: Example 6: S-wave Reflection with Orthogonal hitting Angle along y-axis FDTD Solution (’o’), GB Solution (’-’). . . . . . . . . . . . . . . 150 Figure 4.15: Example 6: S-wave Reflection with Orthogonal hitting Angle along x-axis. FDTD Solution (’o’), GB Solution (’-’). . . . . . . . . . . . . . . 150 Figure 4.16: Example 7: S-wave Reflection with Non-orthogonal Hitting Angle. Left: Gaussian Beam Solution without reinitialization, Right: Gaussian Beam Solution with reinitialization. FDTD Solution (’o’), GB Solution (’-’) . . 151 Figure 4.17: Example 8: S-wave Reflection in Linear velocity without Reinitializa- tion. Left: FDTD Solution Right: Gaussian Beam Solution without Reinitialization. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152 Figure 4.18: Example 8: S-wave Reflection in Linear velocity with Reinitialization. Left: FDTD Solution Right: Gaussian Beam Solution with Reinitializa- tion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153 Figure 4.19: Example 9: S-wave Reflection in Sinusoidal velocity. GB Solution(’o’), FDTD Solution(’-’) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154 Figure 4.20: Example 9: S-wave Reflection in Sinusoidal velocity with Reinitializa- tion. Left: FDTD Solution Right: Gaussian Beam Solution with Reini- tialization. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154 Figure 4.21: Example 10: General Initial Condition Propagation with Reinitializa- tion (First Component). Left: FDTD Solution Right: Gaussian Beam Solution with Reinitialization. . . . . . . . . . . . . . . . . . . . . . . . 156 Figure 4.22: Example 10: General Initial Condition Propagation with Reinitialization (Second Component). Left: FDTD Solution Right: Gaussian Beam Solution with Reinitialization. . . . . . . . . . . . . . . . . . . . . . . . 156 Figure 4.23: Example 10: General Initial Condition Propagation with Reinitialization (Second Component). FDTD Solution (’o’), GB Solution(’-’). . . . . . . 157 Figure 4.24: Log-log plot: Convergence Rate of the New Gaussian Beam Method GB Method Error Curve(’-*’), Line with the slope = 1 2 log(1.5)(’-’)158 Figure B.1: FDTD solution justify. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199 Figure B.2: Convergence Rate of FDTD algorithm . . . . . . . . . . . . . . . . . . . 200 xi LIST OF ALGORITHMS Algorithm 1 Discrete Gaussian Wavepacket Decomposition. . . . . . . . . . . . . . . 18 Algorithm 2 Multiscale Gaussian Beam True-Amplitude Migration . . . . . . . . . . 37 Algorithm 3 Discrete Vector-Valued Gaussian Wavepacket Transform. . . . . . . . 110 xii Chapter 1 Introduction Hyperbolic PDEs arise in a variety of practical applications, ranging from acoustics, elasticity, electromagnetics to geophysics. Therefore, it is desirable to develop fast and efficient algorithms to solve this family of PDEs. Moreover, efficient algorithms are also desirable in a lot of inverse problems modeled by hyperbolic PDEs. The term ’high frequency’ in the high frequency wave propagation is defined relative to the low-frequency background slowness in the model. Therefore, all medium appeared in this thesis is smooth, if not specified. It is very costly for the direct method, the finite difference method or the finite element method for example, to simulate the high frequency wave propagation, since fine grid mesh is required by these methods to capture the oscillation. Therefore, some alterna- tive methods have been developed, such as the traditional geometrical-optics method (WKBJ ansatz), which is required to solve a pair of equations. The first one, which is called eikonal equation, is a first-order nonlinear PDE. The second one is the trans- port equation relying on the differentiability of the result of eikonal equation. It yields faithful asymptotic solutions before caustics occur. However, the amplitude function governed by the transport equation breaks down around the caustics [6, 22, 36], where the phase function is multi-valued [36, 39]. The appearance of caustics is inevitable 1 even in the smooth medium [54]. Figure 1.1: Caustics Appearance in Gaussian Velocity One of the alternatives is the Gaussian beam method [49, 51, 55, 15]. To resolve the caustics problems automatically, the Gaussian beam method relaxes the restriction that the phase term is real-valued. The single Gaussian beam ansatz is made up of a complex-valued phase function and a complex-valued amplitude function near its central ray. Away from the ray path, the beam decays rapidly as a Gaussian profile. The superposition of many single beams will be an asymptotically correct solution of the hyperbolic PDE in the sense that both the initial condition and the boundary condition are satisfied asymptotically as well as the PDE itself. There are two methods based on Gaussian beam method presented in this thesis to resolve two related problems. The first one is about the simulation of the elastic wave in a bounded domain. The second one is using a novel Gaussian beam method to reconstruct reflectors under the surface by the data received along the boundary. 2 −0.4−0.200.200.20.40.60.8x1x2 1.1 Prestack Inversion Process In the first part, we propose a novel inversion and imaging procedure for finding the parameter of the medium by solving the linearized inverse scatter problem. There is a wide range of applications including seismic exploration, medical imaging and underwater acoustics, etc. There are many different migration methods existed to explore the geological structure. The very early one was using the one-way wave equation [16, 17] to recover the accurate travel time and locations of reflectors. However, the popular finite difference and finite element method require extremely refined grid mesh in the high frequency regime to prevent dispersion error. This is also found in the modeling process. Kirchhoff method [7, 8, 28], which bases on the asymptotic method, is mostly investi- gated and widely applied to resolve the high frequency pattern. The Kirchhoff method and all other similar ray-based methods are using the asymptotic approximation (high frequency) of the Green’s function for the acoustic wave equation. G(x, x0, ω) = Aeiωτ (x,x0), (1.1) where τ (x, x0) is the traveltime from the point x0 to the point x. There exist several other problems in the Kirchhoff migration, although it is efficient and flexible. The first obstacle is the presence of caustics points. It is not able to characterize the structure in the presence of caustics, since they rely on the first-arrival traveltime [18, 27] instead of multivalued traveltime. Its usefulness was questioned by 3 [25, 27, 43] since the first-arrival traveltime in complex media usually do not correspond to the most energetic traveltimes crucial for imaging complex structures. To overcome this issue, we use the Gaussian beam migration method [3, 45, 14]. The first obstacle is the way to initialize each beam solution efficiently, in other words, how to characterize the wave propagator by beam solution in a sparse form. Another closely related issue is the way to describe the wavefront of the perturbation. The wavefront defines the singularities of a function not only in spatial space but also with respect to its Fourier transform at each point. It is naturally adaptable to the high frequency wave propagation which can be considered as the propagation of singular- ity. From this point of view, the way to generate the decomposition should consider both the optimal representation of the wave propagator and the sufficient condition to reconstruct the target perturbation media. In some sense, some tradeoff should be obtained. The second obstacle is about the rigorous mathematical analysis of the imaging con- dition. In this paper, the spread loss has to be proved to be compensated for in our √ ω) new imaging method. It is well known that the Gaussian beam solution has O(1/ convergence rate as an asymptotic solution. We will take advantage of this convergence rate and the fact that the parabolic scaling principle is preserved along the propagation to conduct a rigorous error analysis. Besides, we assume that a set of geodesics (rays) have a consistent direction. The assumption simply means that the overturn rays don’t exist in our model and this assumption is natural in the practical applications. Our method enjoys several advantages. First, the flexibility of the Kirchhoff migration is preserved and imaging without losing multi-arrival in the general slowness is possible. 4 Second, due to the Multi-scale Gaussian wavepacket transform and the parabolic scal- ing principle employed, the sparse representation of the wave propagator is obtained. This decomposition also makes the reinitialization feasible during the propagation so that we have more control over the width of beam solution. Third, there is some re- dundancy in the data in the view of our imaging condition. In practice, the error is unavoidable, then this redundancy can help us to reduce the impact brought by the noise since the average value is employed to cancel the variation caused. Fourth, our imaging condition is performed in the time domain to avoid the extra Fourier trans- form on the data set. This feature makes our algorithm more applicable considering the large size of the trace dataset in the real world. The last feature is that the wavefront set being characterized by the Gaussian wavepacket enables us to image the subsurface partially in the sense of controlling the range of dipping angles. There are some other true amplitude migration methods different from the ray-based method described above. Zhang et al. [56, 57] developed true-amplitude common-shot migration for heterogeneous media. Again, only the geometric spreading loss is re- covered in these works. Other types of compensations, for example transmission loss compensation, are discussed in [21]. 1.2 Elastic Wave Equation The elastic wave equation is a good model to describe the seismic wave propagation in a uniform whole space and it has been used widely in the seismology community in both inverse and modeling problem [1]. Similar to its simpler form, i.e. the acoustic 5 wave equation, the elastic wave equation will propagate oscillations in space and time when the initial or boundary condition contains oscillation of the small wavelength. We will develop a multi-scale Gaussian beam method to simulate the elastic wave in a bounded domain. The first problem to apply the Gaussian Beam method is how to decompose any general initial condition to the form suitable to the beam profile. It is resolved by the Multiscale Gaussian Wavepacket transform developed in [48, 5] for the acoustic wave equation and single-scale transform for the Schrodinger equation [47]. Since the Hamiltonian of the wave equation is homogeneous of order one, a Gaussian beam should satisfy the parabolic scaling principle at any given time if it is satisfied at the first place. The wavepacket satisfying the parabolic scaling principle is defined as the wavelength of the typical oscillation of this wavepacket being equal to the square of the width of this wavepacket. To propagate each wavepacket, the dynamics system can be obtained in the typical way. In this paper, we extend this idea further to the decomposition of the vector- valued initial value and preserve all the optimal properties of the Multiscale Gaussian Wavepacket transform. The second problem is to derive the reflection dynamics. Unlike the Cauchy problem, we have to consider the reflected beams. Most recently in [4], a numerical method has been proposed for the acoustic wave equation. Other discussions can be found in [9, 49]. There will be more complex situations concerned in the elastic wave model as there are two different types of wave modes in the process. The difference between different wave modes requires extra efforts to preserve the accuracy. 6 The third problem is how to reinitialize Gaussian beams efficiently. Gaussian beam solutions behave well around the caustics, however, not so well in the long-term wave propagation. See more details illustrated in [48, 32, 42]. Therefore, there have been various methods developed to control the width of beams. One method, for example, is to call the Multiscale Gaussian wavepacket algorithm repeatedly during the propagation [48, 4]. A global time T is set at prior such that beams will be summarized after propagating for T and then decompose the resulting temporary wavefield to a new system of wavepackets. This process will be repeated several times. A new reinitialization method is proposed in this thesis which claims to have more freedom and remains to be the asymptotic solution to the elastic wave equation. Instead of decomposing the general wavefield after summation each time, we target on each single beam in this new reinitialization method. It will give us more freedom to choose which beams needed to be reinitialized rather than all of them. Moreover, with applying the reinitialization strategy to each single beam, there exists the explicit expression. 1.3 Related Work The fact that the Gaussian beam ansatz can be used to solve the wave equation has been known to the pure mathematical community since sometime in the 1960s [3]. Then it is applied to simulate the propagation of the singularity [49, 30]. A single Gaussian beam ansatz is an asymptotic solution concentrated on a single ray curve. The critical point is to have a global solution to the Hessian of the phase function so 7 that the transport equation is well-defined. Moreover, the fact that imaginary part of the Hessian remains to be S.P.D. leads to a well-localized solution. The localization is justified theoretically [49, 40, 52]. Considering its ability to resolve multi-valued phase function automatically, the Gaus- sian beam method was firstly introduced as a seismic imaging method by Hill in the form of the poststack [28] and then the prestack migration procedure [29]. The perfor- mance of the Gaussian beam migration is further tested by the common shot geometry [26]. Most recently, a purely Eulerian computational approach was proposed in [33] which improves the numerical method’s efficiency and its application in the semiclas- sical quantum mechanics has been proposed in [31]. See [9, 42] for other recent works. Besides the Gaussian beam method, there are several possible ways to construct global asymptotic solutions for the wave equation even in the presence of caustics. The first approach is based on Ludwigs uniform asymptotic expansions at caustics [35, 10] which requires that the caustic structure is given. The second approach is is based on the Maslov canonical operator theory [41], which requires to identify where the caustics are at prior. There is some recent advance in resolving the multi-valued traveltime problem. A new method called fast Huygens sweeping method has been proposed in [37, 38] to solve the Helmholtz equations in the inhomogeneous media and then it is used to solve Schrodinger equation [34]. They take advantage of the fact that eikonal equation is well-defined around the source point and Huygens-Kirchhoff secondary source principle. There have been some recent advance in the optimal representation of the wave propa- 8 gator [52, 11, 50]. It is closely related to the Fourier integral operator representation of the hyperbolic system and the special proposition of its phase function. The multiscale Gaussian Wavepacket transform is developed [48] for the wave equation and the single scale transform for the Schrodinger equation [47] which is also the fundamental basis of our algorithm. This difference comes from the different Hamiltonian for these two equations. Other paper also apply the Gaussian beam method in their true ampli- tude migration [46, 2], however, they do not require the parabolic scaling principle for Gaussian beams as our prestack inversion method does. The parabolic scaling principle provides the theoretical basis for our new reinitialization method as well as the proof of the correctness of the imaging operator in our new multi-scale inversion algorithm. Other methods using the similar idea can be found in [11, 50]. There are various types of such wavepackets, curvelet [13, 12] and wave atoms [19, 20] for example. However, the Multi-scale Gaussian beam method is different from these methods in that the single Gaussian wavepacket corresponds to the single Gaussian beam at final time T , while the curvelet frame does not have this one-to-one relationship. On the other hand, the Gaussian wavepacket transform has been proved a stable and efficient decomposition of the arbitrary function [48], equivalently, the wavepacket is a good characterization of the wavefront set. The Gaussian window function or Gabor frame [24] are both well-localized in the phase space as the Fourier transform of a Gaussian profile function is again a Gaussian profile function. The size of the Gaussian window function in the phase space can be determined by the Heisenberg Uncertainty Principle. 9 1.4 Contents The remainder of this thesis is organized as follows. In Chapter 2 we present a brief in- troduction of constructing and propagating a single Gaussian beam, which is the foun- dation to the following derivation. We then describe the original multi-scale Gaussian wavepacket transform [48] in Section 2.2. In Chapter 3, we propose a new prestack inversion process based on the multi-scale Gaussian beam method. We then modify the Gaussian wavepacket transform in Section 3.1 to adapt to the imaging operator. We then develop the new imaging operator with the help of Gaussian wavepacket transform and the Gaussian beam functions. Based on this operator, we propose the main inversion algorithm in Section 3.2.4. The next part is devoted to proving the correctness of this new algorithm in Section 3.3. In Section 3.4, we discuss the fast method to calculate the imaging operator. In the last section of this chapter, we select several well-designed numerical examples to justify the correctness of our analysis and the approximations mentioned earlier. In Chapter 4 we propose the Multiscale Gaussian beam method to solve the elastic wave equation with highly oscillated initial condition. We first extend the Multiscale Gaussian wavepacket transform to the vector-valued initial condition in Section 4.3 and develop a new propagating dynamics for each single beam. After proposing the decomposition scheme, the reflection dynamics for the homogeneous Dirichlet Bound- ary value is derived in Section 4.5. The difference among various types of reflection is analyzed in Section 4.7.2 and a new efficient reinitialization method is proposed to resolve the problem from S-wave reflection in Section 4.6 and Section 4.7.3. The reference solution to the elastic wave equation in general case is provided by the Finite- 10 Difference Time-Domain (FDTD) with staggered grid [53] and is justified in Section B. In Section 4.8, several numerical experiments are conducted to show the correctness and the convergence rate of our new multi-scale Gaussian beam method. 11 Chapter 2 Single Gaussian beam ansatz and Multiscale Gaussian wavepacket Transform 2.1 Single Gaussian beam ansatz The Gaussian beam solution itself is an asymptotic solution of the acoustic wave equa- tion even around caustics. 1 c2(x) t u(x, t) − ∆u(x, t) = 0, ∂2 (2.1) where x is the point coordinate in the space Rd and c(x) is smooth, positive and bounded away from zero. Similar to the Geometric-Optics ansatz, the Gaussian beam also assumes that the solution follows the form, u(x, t) = A(x, t)eiωτ (x,t), (2.2) 12 where ω is a large wavenumber, τ (x, t) is the phase function and A(x, t) is the amplitude function. The asymptotic solutions means u(x, t) (2.2) satisfies the wave equation (2.1) with small error when frequency ω is large. After inserting equation (2.2) into equation (2.1) and organize all terms according to the order of ω, there will be two equations obtained, which are eikonal and transport equations, governing τ (x, t) and A(x, t) respectively. They come from the leading orders in inverse power of the frequency ω. t (x, t) − c2|∇τ (x, t)|2 = 0 τ 2 2Atτt − 2c2∇A · ∇τ + A(τtt − c2trace(τxx)) = 0. Phase function τ : After factoring out equation (2.3), there are two branches generated, τt ± c(x)|∇τ| = 0. (2.3) (2.4) (2.5) Equation (2.6) is a Hamilton-Jacobi equation with the Hamiltonian G±(x, p) = ±c(x)|p|. We consider the generic situation for the eikonal equation, τt + G(x,∇τ (x, t)) = 0. (2.6) 13 We apply the method of the characteristics to solve the eikonal equation (2.6). dx dt dp dt x = Gp(x(t), p(t)), = −Gx(x(t), p(t)), = p0. (2.7) (cid:12)(cid:12)(cid:12)t=0 (cid:12)(cid:12)(cid:12)t=0 p = x0; where we define the ray trajectory γ = {(x(t), p(t)) : t ≥ 0}, whose initial point is (x0, p0) in the phase space. We have that the momentum p(t) = ∇τ (x(t), t) along the ray. To derive the dynamics of Hessian matrices, we first differentiate the eikonal equation (2.6) with respect to t and x: τtx(x, t) + Gx(x,∇τ (x, t)) + τxx(x, t)Gp(x(t),∇τ (x, t)) = 0, τtt(x, t) + Gp(x(t),∇τ (x, t)) · τxt(x, t) = 0, (2.8) (2.9) Differentiating equation (4.13) with respect to x yields τtxx + Gxx + τxxGxp + (Gxp)T τxx + τxxGppτxx + τxxxGp = 0. (2.10) Therefore, the Hessian M (t) = ∇∇τ (x(t), t) satisfies the following Riccati equation, dM dt + Gxx + M Gxp + GT xpM + M GppM = 0. M = iI (2.11) (cid:12)(cid:12)(cid:12)t=0 The size parameter  will be given after introducing the Gaussian wavepacket trans- form. 14 One of the most significant differences between the Gaussian beam and other ray-ansatz methods is that beams’ phase functions τ (x, t) are complex-valued. Complexifying the equation guarantees a well-defined Hessian and a well-defined transport equation as a result. This is not true in general case [49]. Furthermore, the positive definite imaginary part is always true throughout the propagation for smooth ray trajectories. Lemma 2.1.1. If the Hamiltonian G is smooth enough, then the Hessian M(t) along the ray path γ has a positive-definite imaginary part, provided that it initially does. Transport Equation A(x, t): With (x(t), p(t), M (t)) well-defined along the way, we can solve the transport equation (2.4). Taking advantage of the fact that dA(x(t), t) dt = At + ∇A · Gp(x(t), p(t)), equation (2.4) is reduced to dA dt + A 2G (v2trace(M ) − Gx · Gp − G T (cid:12)(cid:12)(cid:12)t=0 p MGp) = 0 . A = A0 (2.12) A single beam solution is in the following form, x0 p0 (x, t) = A(x, t)eiωτ (x,t), U (2.13) and the phase function is approximated by applying Taylor expansion around the 15 central trajectory {x(t) : t ≥ 0} at time t, τ (x, t) = p(t) · (x − x(t)) + (x − x(t))T M (t)(x − x(t)), 1 2 (2.14) and the amplitude is approximated by its value at trajectory γ at the same moment, A(x, t) = A(x(t), t) = A(t). (2.15) 2.2 Multiscale Gaussian Wavepacket Transform After constructing a single beam solution, the next problem is how to set up the initial condition for the ODE system. The answer is Multiscale Gaussian wavepacket transform, which will be introduced briefly in this section. More details about this phase space decomposition can be found in paper [48] and its single scale application can be found in paper [47]. The wavepacket transform is applied to L2 functions f in the Rd space. We first partition the Fourier space Rd into several Cartesian coronae Cl for l ≥ 1 as Cl = {ξ = (ξ1, ξ2,··· , ξd) : max 1≤i≤d |ξi| ∈ [4l−1, 4l]}. Now it is obvious to see that the L2 norm of ξ in Cl is O(4l). For each Cl, we can further partition it into multiple windows with width 2l, Bl,i = [2lis, 2l(is + 1)], d(cid:89) s=1 16 where the integer multiindex (i1, i2,··· , id) is any possible choice such that the box is in the lth layer, i.e. Bl,i ⊂ Cl. After defining these cell boxes Bl,i, we can define the Gaussian profile function gl,i associated with the box Bl,i by the following formula, (cid:32)|ξ−ξl,i| (cid:33)2 − gl,i(ξ) ≈ e σl , (2.16) where ξl,i is the center of the box Bl,i and σl = 2l is the width of the box Bl,i. The scale listed here is designed carefully following the parabolic-scaling principle. This is the key to the success of our multi-scale imaging process as it provides the theoretical justification of the size of each Gaussian wavepacket. The later proof and error analysis will rely on this conclusion heavily. To have a partition of unity, one also needs the conjugate filters hl,i, such that (cid:80) gl,i(ξ) l,i g2 l,i(ξ) hl,i(ξ) = , (2.17) The proof that the functions hl,i are well defined and well-localized can be found in paper [48]. It is easy to see that(cid:80) l,i gl,ihl,i = 1. By shifting the central point, ˆφl,i,k(ξ) = ˆψl,i,k(ξ) = −2πi kξ Ll gl,i(ξ), −2πi kξ Ll hl,i(ξ). e e 1 d/2 L l 1 d/2 L l 17 Taking the inverse Fourier transforms yields their definitions in the spatial domain: 2π(x− k Ll e )·ξ gl,i(ξ)dξ 2π(x− k Ll e )·ξ hl,i(ξ)dξ (2.18) (2.19) (cid:90) (cid:90) Rd Rd φl,i,k(x) = ψl,i,k(x) = 1 d/2 L l 1 d/2 L l (cid:18)(cid:114) π (cid:19)d σl Ll φl,i,k(x) ≈ The approximation expression of the wavepacket φl,i,k, 2πi(x− k Ll e )ξl,ie −σ2 l π2|x− k Ll |2 . (2.20) We list the lemma from paper [48] without proof to show that our decomposition is correct. Lemma 2.2.1. For any f ∈ L2(Rd), we have (cid:88) f (x) = (cid:104)ψl,i,k, f(cid:105)φl,i,k(x). (2.21) l,i,k The idea of decomposing discrete signals into wavepackets is very similar to the contin- uous case, therefore, we skip this part and provide the pseudo code below. The total Algorithm 1 Discrete Gaussian Wavepacket Decomposition 1. Apply the Fast Fourier Transform(FFT) to the discrete signal f . 2. Compute hl,i(ξ) ˆf (ξ). 3. Wrap the result to the domain [−2σl, 2σl]. 4. Apply the Inverse Fourier Transform to obtain the coefficients cl,i,k. cost of this algorithm is O(N d log(N )). 18 2.2.1 Multiscale Gaussian Beam Method With the Mutiscale Gaussian Wavepacket transform defined above, the initial con- dition (x0, p0, M0, A0) for a single beam solution will be defined corresponding to a wavepacket φl,i,k, = Gp(x(t), p(t)), x0 = k Ll ; = −Gx(x(t), p(t)), p0 = 2π ξl,i |ξl,i|, = −Gxx − M Gxp − GT = − A 2G (v2trace(M ) − Gx · Gp − G T dx dt dp dt dM dt dA dt and xpM − M GppM, M0 = i(2π2σ2 (cid:18)(cid:114) π (cid:19)d l /|ξl,i|)I, Ll N p MGp), A0 = σl . (2.22) u(x, t) = A(x(t), t)e i|ξl,i|τ (x,t) , (2.23) Then we can use the dynamic system (2.7)-(2.12) to propagate beams. The large wavenumber |ξl,i| serves as the key point of asymptotic methods. However, through- out the following derivation, we will combine this constant |ξl,i| into the phase for convenience. We argue that there’s no difference. Denote the beam solution using initial condition (x0, p0, M0, A0) as (x(t), p(t), M (t), At) and the one using initial condition (x0,|ξl,i|p0,|ξl,i|M0, A0) as (ˆx(t), ˆp(t), ˆM (t), ˆA(t)). With respect to the ray trajectory, dx(t) dt d(|ξl,i|p(t)) dt = v(x(t)) |ξl,i|p(t) |ξl,i||p(t)|, = ∇v(x(t))|ξl,i||p(t)|, x0 = k Ll |ξl,i|p0 = 2πξl,i 19 Since (x(t),|ξl,i|p(t)) is the solution of the same Hamiltonian system equipped with the same initial condition as (ˆx(t), ˆp(t)), by the uniqueness of the initial value problem, ˆp(t) = |ξl,i|p(t) and ˆx(t) = x(t). In terms of Hessian M , again we multiply |ξl,i| on both sides of equation (2.11). (cid:33)T d(|ξl,i|M ) dt = −∇∇v(x(t))|ξl,i||p(t)| − |ξl,i|M∇v (cid:32) |ξl,i|p(t) |ξl,i||p(t)| M (cid:32) |ξl,i|p(t) |ξl,i||p(t)| − |ξl,i|∇v (cid:33) (|ξl,i|M ) t xp(ˆx(t), ˆp(t)) ˆM v(x(t)) − (|ξl,i|M ) |ξl,i||p(t)| I − v(x(t)) |ξl,i||p(t)|3 p(t)pT = −∇∇v(ˆx(t))|ˆp(t)| − ˆM Gxp(ˆx(t), ˆp(t)) − GT − ˆM Gpp(ˆx(t), ˆp(t)) ˆM Again, by the uniqueness of the initial value problem, ˆMt = |ξl,i|M (t). The derivative term Gpp is given by Proposition 2.2.1. The second order derivative of the Hamiltonian about the momen- tum variable is (cid:16)|p|2Id − ppT(cid:17) , G± pp(x, p) = ±v(x) |p|3 (2.24) where Id is the identity matrix. The proof is easy, so we omit it here. d ˆA dt = − ˆA 2|ξl,i|G(x(t), p(t)) + ˆA 2|ξl,i|G(x(t), p(t)) (cid:16)|ξl,i|GT (v2|ξl,i|M (t) − |ξl,i|Gx(x(t), p(t)) · Gp(x(t), p(t))) (cid:17) p (x(t), p(t))M (t)Gp(x(t), p(t)) . (2.25) Apparently, A(t) and ˆA(t) share the same ODE and initial condition. 20 Chapter 3 Multiscale Gaussian-Beam Transforms for True Amplitude Prestack Migration 3.1 Modified Multiscale Gaussian Wavepacket Trans- form We have so far finished introducing the Gaussian wavepacket transform in [48]. To meet our inversion algorithm’s requirements, we have to substitute (cid:104)ψl,i,k, f(cid:105) in equation (2.21) with (cid:104)φl,i,k, f(cid:105). Several new notations are needed in the following arguments. We first define a set for each frequency ξ as its cover set S(ξ), S(ξ) = {(l, i) : gl,i(ξ) > 0}. (3.1) 21 Bl is defined as the border of the partitioning, Bl ≡ {ξ : max 1≤s≤d |ξs| ∈ [4l−1, 4l]} ∩ {ξ : max 1≤s≤d |ξs| − 4l−1 ≥ 2l} ∩ {ξ : 4l − max 1≤s≤d |ξs| ≤ 2l}. (3.2) The remainder part Cl \ Bl is defined as the major part of Cl. Obviously, the border part Bl is much smaller compared with the major part. We will focus on the major part first. All frequencies in the major part of Cl will not interact with those from different levels, that is, the cover sets of all frequency ξ in the major part Cl \ Bl will only contain the compact support functions gl,i from the same level. This allows us to prove the following claims. Proposition 3.1.1. If ξ ∈ Cl \ Bl, we have (cid:88) (cid:88) (cid:88) g2 (cid:48) l ,i (cid:48) (ξ) = (cid:48) l (cid:48) ,i g2 (cid:48) l (cid:48) (ξ) + ,i g2 (cid:48) l (cid:48) (ξ) + ,i (cid:48) l (cid:48) =l,i (cid:48) l (cid:48) l,i g2 (cid:48) l (cid:48) (ξ) = ,i (cid:88) (cid:48) l (cid:48) =l,i g2 (cid:48) l ,i (cid:48) (ξ). (3.3) Proof. We assume that the concerned frequency ξ = (ξ1, ξ2,··· , ξd) satisfies, ξ1 = max 1≤s≤d |ξs|. (3.4) 22 On the other hand, g l (cid:48) (cid:48) is a compact support function in the box centered at ξ l ,i (cid:48) (cid:48) ,i with side length 2l (cid:48) +1. Since (cid:48) 2l +1 ≤ 2l, (3.5) (3.6) (3.7) (3.8) We first check the first term in equation (3.3), where l (cid:88) (cid:48) l (cid:48) l. We denote, (cid:48) ξ l (cid:48) ,i ,s0 = max 1≤s≤d |ξ l (cid:48) (cid:48) ,i ,s |. Therefore, ||ξ − ξ l (cid:48) (cid:48)|| ≥ |ξs0 − ξ ,i (cid:48) (cid:48) ,i ,s0 | ≥ |ξ l (cid:48) (cid:48) ,i ,s0 | − |ξs0| ≥ |ξ l (cid:48) (cid:48) ,i ,s0 | − |ξ1| ≥ 2l+1. (3.9) l To summarize, for any ξ l (cid:48) ,i (cid:48) on the higher level, we have |ξs0 − ξ l (cid:48) (cid:48) ,i ,s0 | ≥ 2l+1, 23 (3.10) along the s0 direction. We notice that the compact support area of g ,i direction s0 has side length 2l+2, however, ξ must be on the left side of ξ l l (cid:48) (cid:48) along ,i (cid:48) (cid:48) along the the s0 direction. Then, ξ will be at most on the edge of support area of the function (cid:48) g l ,i (cid:48) . For any central frequency ξl,i ∈ Cl \ Bl, we define Jl,i Jl,i ≡(cid:88) (cid:48) (cid:48) ,i l g2 (cid:48) l (cid:48) (ξl,i). ,i g2 (cid:48) l ,i (cid:48) (ξl,i). (cid:48) By Proposition 3.1.1, Jl,i ≡ (cid:88) (cid:48) l =l,i Proposition 3.1.2. Jl,i is independent of the index i, that is Jl,i = Jl = g2 (cid:48) l (cid:48) (ξl,i), ,i ∀ξl,i ∈ Cl \ Bl. (cid:88) (cid:48) l (cid:48) =l,i (3.11) (3.12) (3.13) Proof. Suppose there are two different central frequencies ξl,i and ξl,i∗ in the same major part of Cl. By Proposition 3.1.1, we should only consider gl,i from the same level. On the other hand, in the fixed level l, the compact support area of each box function gl,i has side length 2l+1 along each dimension. In fact, each (l, i) will have overlapping support with only two other i-indexes in each direction, since the central frequency ξl,i in the level l is chosen as 2li with all possible integer multi index i. 24 (cid:32)|ξl,i,1 − ξ l,i | (cid:48) ,1 σl (cid:33)2 d Using Proposition 3.1.1, we have exp (cid:48) 1=i1−1 i  i1+1(cid:88) −2 (cid:32) −2 −2 d (cid:33)2 (cid:32)|ξl,i∗,1 − ξ (cid:48) i σl l,i σl l (cid:48) (cid:48) ,i (cid:48) ,i g2 (cid:48) l (cid:88) (cid:48) (ξl,i) ≈  i =1(cid:88)  i∗ 1+1(cid:88) (cid:48) 1=i∗ 1−1 =−1 exp = = (cid:48) i i exp (cid:33)2 d | (cid:48) ,1 ≈(cid:88) (cid:48) (cid:48) l ,i g2 (cid:48) l (cid:48) (ξl,i∗). ,i (3.14) Two summations are the same since Gaussian functions are only about the distance between two frequencies. Proposition 3.1.3. For all frequency ξ ∈ Cl \ Bl, i e = σl −2 (cid:33)2 (cid:33)2 (cid:88) (cid:48)∈Zd (cid:88) (cid:48)||∞≤4l−1 (cid:88) (cid:48)||∞≤4l ||ξ−ξ (cid:48)|| (cid:32)||ξ−2li (cid:48)|| (cid:32)||ξ−2li (cid:48)|| (cid:32)||ξ−2li (cid:48)|| 2 4l−1<||2li −2 e −2 e −2 σl σl + l,i σl ||2li + (cid:88) (cid:48) i = e (cid:32)||ξ−2li (cid:48)|| (cid:33)2 σl −2 e (cid:88) (cid:48)||∞>4l ||2li (cid:33)2 + , (3.15) where ||x||∞ ≡ max1≤s≤d |xs| and ξ l,i (cid:48) is the central frequency defined in the wavepacket transform at the level l.  is a small number which can be ignored. 25 Proof. −2 e (cid:32)||ξ−2li (cid:48)|| (cid:90) 4l−1−2l−1 (cid:88) (cid:48)||∞≤4l−1 (cid:90) σl s=d(cid:89) ··· 0 s=1 πerfc(3), ||2li l 2σ−d ≤ √ (cid:33)2 ≤ (cid:18)|ξs−cs| σl − e (cid:19)2 dc1dc2 ··· dcd (3.16) where erfc(·) is the complementary error function. Here we assume c is the central frequency at the lower level, therefore, c1 ≤ 4l−1 − 2l−1. Meanwhile, we assume that ξ1 = max1≤s≤d |ξs| ≥ 4l−1 + 2l, then ξ1 − c1 ≥ 2l + 2l−1. With σl = 2l−1, we have the upper bound specified in equation (3.16). (3.17) (3.18) Similarly, we have (cid:88) (cid:48) l >l,i e (cid:48) σ−d l ||ξ−ξ (cid:48) l σl − 2 (cid:48)|| ,i ≤ √ πerfc(5). (3.19) 26 Then (cid:88) (cid:48)∈Zd i e (cid:32)||ξ−2li (cid:48)|| σl − ||ξ−ξ (cid:48) l σl 2 (cid:48)|| ,i (cid:33)2 (cid:88) (cid:48) l (cid:48) =l,i − e = √ πerfc(3) + √ πerfc(5). + (3.20) Proposition 3.1.4. For any frequency ξ in the major part Cl \ Bl, we have (cid:88) (cid:48) l (cid:48) ,i (cid:88) (cid:48) l (cid:48) =l,i (cid:48) (ξ) ≈ Jl = g2 (cid:48) l ,i g2 (cid:48) l (cid:48) (ξl,i), ,i ∀ξ ∈ Cl \ Bl. (3.21) Proof. We have already proved in Proposition 3.1.1 that any frequency ξ in the major part Cl \ Bl satisfies, g2 (cid:48) l ,i (cid:48) (ξ) = g2 (cid:48) l ,i (cid:48) (ξ). (3.22) (cid:88) (cid:48) (cid:48) ,i l Meanwhile, according to the Poisson summation formula, (cid:88) (cid:48) l (cid:48) =l,i (cid:88) (cid:19)d (cid:88) (cid:18) a√ 2π − σ2 e 2 ||x+ma||2 = a ||2 2σ2||2π m − 1 e = const. (3.23) m∈Zd m∈Zd ||ξl,i−2li σl (cid:48)|| −2 2 . (3.24) Then we have, (cid:33)2 (cid:32)||ξ−2li (cid:48)|| σl (cid:88) (cid:48)∈Zd i −2 e (cid:88) (cid:48)∈Zd i e = 27 Using equation (3.15), (cid:48) (ξ) ≈ (cid:88) ||ξl,i−2li −2 =l,i e (cid:48) (cid:48) l σl −2 (cid:48)|| (cid:88) (cid:48) (cid:48) ,i l = g2 (cid:48) l ,i (cid:88) (cid:48)∈Zd i e 2 (cid:48)|| ,i (cid:48) l (cid:48) l σ ||ξ−ξ 2 ≈ (cid:88) (cid:48) l =l,i (cid:48) −2 e (cid:33)2 (cid:48)|| ≈ (cid:88) ||ξl,i−ξ (cid:48)∈Zd i e σl −2 (cid:48) ,i l σl (cid:32)||ξ−2li 2 (cid:48)|| ≈ Jl. (3.25) The proposition is proved. For the frequency ξ on the border of the partitioning, their sums satisfy Jl−1 ≤(cid:88) l,i(ξ) ≤ Jl. g2 (3.26) l,i Using Jl as an approximation will yield an overestimation, however, it is negligible since Bl is much smaller compared with the major area away from the border. To summarize, Then, Jl ≈(cid:88) l,i g2 l,i(ξ), ∀ξ ∈ Cl. hl,i(ξ) ≈ 1 Jl gl,i(ξ). We therefore have a modified inverse wavepacket transform as the following, Lemma 3.1.1. For any f ∈ L2(Rd), we have f (x) ≈(cid:88) l,i,k (cid:104)φl,i,k, f(cid:105)φl,i,k(x). 1 Jl 28 (3.27) (3.28) (3.29) To end this part, we would like to display some numerical results to justify Lemma 3.1.1. The denominator in the expression of hl,i (2.17) should be a step function about the index l, suggested by Lemma 3.1.1. The range of the frequency concerned is [64 : 320] × [64 : 320]. Figure 3.1: (cid:80) l,i g2 l,i in different level l 3.2 Multiscale Gaussian Wavepacket Inversion 3.2.1 Setup of the True Amplitude Migration Problem Let us suppose the wave propagation is governed by the scalar wave equation (2.1) with the wave propagation velocity decomposed as, c = v(1 + α), (3.30) where v is the macro velocity being responsible for the traveltime and amplitude. Moreover, it is assumed to be smooth and does not provide the significant energy back to the boundary data. The rapid perturbation α is small but reflects the wave signal back to the boundary data. In our inversion model, v is known at prior and our target 29 is to image α. We simplify our model as a constant density fluid occupying a half-space X ≡ {x ∈ Rd : xd ≥ 0}. The boundary data D(r, s, t) used in this paper is organized as the common-shot trace, for example, Fig. 3.2, r parametrizes receiver positions on the surface ∂X ≡ {x ∈ Rd : xd = 0}, while s parametrizes source positions. We also assume that the sources are contained in an open set Os ⊂ ∂X and receivers are in an open set Or ⊂ ∂X. Therefore, the boundary data D is a function defined at Os × Or × [0, T ]. Similar to [44], we make some assumptions about rays. Figure 3.2: A typical source gather in Gaussian slowness Assumption 3.2.1. There are no rays leaving points in the subsurface {x ∈ X : xd > 0} and returning to graze Os or Or. Moreover, there exists an universal lower bound b, such that |pd| ≥ b||p||, (3.31) for any rays hitting the surface where p is the momentum variable and pd is the dth component of p. 30 −20200.511.522.5xrt Assumption 3.2.2. Rays departing from a source in Os and traveling into the sub- surface do not return to receivers in Or. Assumption 3.2.3. There exists δ > 0, such that v(x) is a constant if 0 ≤ xd ≤ δ. 3.2.2 Born Approximation for the Trace Data Denote the wave propagator with background velocity v as L0 and the wave propagator with true velocity c as L. Meanwhile, the corresponding Green’s functions are written as G0 and G, respectively, G = −L−1; G0 = −L−1 0 , (3.32) and by some formal computations, G = G0 + G0(L − L0)G = G0 + G0V G, (3.33) where V = −L0 + L. The Born approximation assumes the whole scattering process as the following. The signal is initiated from the source and travel through a smooth medium afterwards. Then at some moment, it hits the reflector under the surface and is reflected back. Therefore, the boundary data D(r, s, t) is the reflection data, or the wavefield along the boundary minus the direct wave. During the whole process, we assume that the reflection or scattering only happens once so that we ignore multiple reflections. 31 Define the incident wavefield G0(x, t; s) generated by the source point s, L0G0(x, t; s) = −δ(t)δ(x − s), G0|t<0 = 0. (3.34) Then the reflection signal is obtained by total differentiation. We assume the true velocity c = v + αv = v + δv and the total wavefield u = u0 + δu, where L0u0 = 0, t u − ∆u ≈ ( 1 c2 ∂2 1 v2 − 2δv v3 )∂2 t (u0 + δu) − ∆(u0 + δu). Considering the first order term, we have L0δu = 2δv v3 ∂2 t u0. Now the perturbed wavefield δG [7, 46] satisfies, L0δG = 2α v2 ∂2G0 ∂t2 . (3.35) (3.36) (3.37) This perturbed wavefield δG is the data recorded along the surface based on the Born approximation, that is, (cid:20)2α (cid:21) v2 D (r, s, t) = δG = (cid:90) (cid:90) dx 2α v2 dh ˆG0(r, t − h; x) ∂2 ˜G0 ∂t2 (x, h; s), (3.38) where both ˜G and ˆG are Green’s functions. The Green’s function ˆG0(r, t − h; x) represents the perturbation received at the receiver r at the moment t − h, and its source is the subsurface point x. ˜G is about source points s and ˆG is about receivers 32 r. The same rules are applied to other functions. (cid:20)2α (cid:21) v2 D (cid:90) (cid:90) (r, s, t) = ∂2 ∂t2 dx 2α v2 dh ˆG0(r, t − h; x) ˜G0(x, h; s). (3.39) To make things easier, we would like to develop our algorithm in the frequency domain instead of the time domain so that we can simplify the convolution operator above. Applying the following Fourier transform in time, we have, Here we abuse the notation by writing the Fourier transform of the Green’s function ˆG0(r, t; x) about the time variable t by ˆG0(r, ω; x). In equation (3.40), the reciprocity of the Green’s function is involved as we replace ˜G0(x, ω; s) with ˜G0(s, ω; x). 3.2.3 Multiscale Gaussian Beam Approximation of the Green’s Function We then approximate the Green’s function in the high frequency regime by the summa- tion of Gaussian beams so that we can define the following multiscale Gaussian-beam transform of the perturbation of the velocity, (cid:20)2α (cid:21) v2 D (cid:90) (cid:90)(cid:90) (r, s, ω) = −ω2 dx 2α v2 dξdη ˆUGB(r, ω; x, ξ) ˜UGB(s, ω; x, η), (3.41) (cid:20)2α (cid:21) v2 D (r, s, ω) = −ω2 dx 2α v2 ˆG0(r, ω; x) ˜G0(s, ω; x). (3.40) (cid:90) 33 where without confusion, we sometimes shorten D (cid:105) (cid:104) 2α v2 (r, s, ω) to be D(r, s, ω) so that we use D(r, s, ω) to denote the Fourier transform of the boundary data D with respect to the time variable t. ˆUGB(r, ω; x, ξ) is the beam solution in the frequency domain starting at the point x with the initial momentum ξ. From now on, the following notation is used to the end, ˆUGB(r, ω; x, ξ) = ˆU x ξ (r, ω); ˜UGB(s, ω; x, η) = ˜U x η (s, ω). The Green’s function can be considered as the solution to the acoustic wave equation whose initial velocity is a Dirac-delta function by Duhamel’s principle and the mul- tiscale Gaussian wavepacket transform can be applied to decompose the Dirac-delta function. t G0(x, t; s) − ∆G0(x, t; s) = 0, G0 ∂2 1 v2(x) (3.42) (cid:12)(cid:12)(cid:12)t=0 (cid:12)(cid:12)(cid:12)t=0 = 0, ∂G0 ∂t = −δ(t)δ(x − s). Although the multiscale transform introduced in Section 2.2 and reference [48] is de- signed to decompose any general L2 functions, the Dirac-delta function can be approx- imated by some L2 functions. 3.2.4 True-Amplitude Migration Process A new operator Kpq applied to the perturbation α can be defined, which is correspond- ing to the certain pair of Gaussian beams, (cid:18) (cid:19) Kpq 2α v2 (r, s, ω) = dx 2α v2 ˆU x p (r, ω) ˜U x q (s, ω), (3.43) (cid:90) 34 which will be called the atomic Gaussian-beam transform. The operator Kpq maps the subsurface information to the boundary data. (cid:90)(cid:90) (K∗ pqg)(y, ω) = drds ¯ˆU y p ¯˜U y q g(r, s, ω), (3.44) which is the adjoint of the atomic Gaussian-beam transform. Applying K∗ pq to the boundary data D yields a single-frequency prestack angle-gather imaging function Ipq, (3.45) (3.46) (3.47) Ipq(y, ω) =(cid:0)K∗ pqD(cid:1) (y, ω), or to write it completely, Ipq (cid:20)2α (cid:21) v2 (y, ω) = (cid:18) K∗ pqD (cid:20)2α (cid:21)(cid:19) v2 (y, ω). Substitute equation (3.41) about the surface data D into equation (3.45), (cid:20)2α (cid:21) v2 Ipq (cid:90)(cid:90) (cid:90)(cid:90) (y, ω) = −ω2 (cid:90) dx 2α v2 drds ¯ˆU y p (r, ω) ¯˜U y q (s, ω) dξdη ˆU x ξ (r, ω) ˜U x η (s, ω). We will later show that the following approximation (3.47) is correct, − (cid:90)(cid:90)(cid:90)(cid:90) ≈ E(p, q, y)ei(p+q)(y−x)e(y−x)T iM0 dωdrdsdξdηω2 ¯ˆU y p (r, ω) ¯˜U y q (s, ω) ˆU x 2 (y−x), ξ (r, ω) ˜U x η (s, ω) where E(p, q, y) is a constant related to the parameters of the corresponding wavepacket, 35 and M0 is a symmetric matrix with a positive definite imaginary part as defined in the Multiscale Gaussian beam propagation. The integral (cid:82) dω should be interpreted as(cid:82) χ(ω)dω, where χ is an arbitrary C∞ function which is zero in the low frequency region and is 1 for the high frequency. If we integrate (3.46) further with respect to ω, then we will have the wavepacket transform about 2α v2 , (cid:90) 2α dx Ipq v2 (cid:20)2α (cid:21) (y, ω)dω = E(p, q, y) (cid:90) As we can see in equation (3.48),(cid:82) dωIpq is essentially the Gaussian wavepacket trans- v2 in the direction of p + q. The integral (cid:82) dωIpq(y, ω) can be v2 ei(p+q)(y−x)ei(y−x)T M0 form of the function 2α 2 (y−x). (3.48) considered as taking the inverse Fourier transform to obtain Ipq(y, t) at t = 0. Accord- ing to Claerbout imaging principle [29], Ipq(y, t) at t = 0 yields the initial state of the subsurface that we want to image. By using the modified wavepacket transform (3.29), we can reconstruct perturbation 2α v2 through the imaging function (3.48). (cid:88) (cid:88) l,i,k 2α v2 = = 1 Jl v2 , φl,i,k(cid:105)φl,i,k(x) (cid:104)2α (cid:104) 2α (cid:105) (cid:82) Ipq (cid:88) ( k Ll v2 , ω)dω l,i,k p+q=ξl,i EJl Based on equation (3.49) , 36 φl,i,k(x). (3.49) Algorithm 2 Multiscale Gaussian Beam True-Amplitude Migration 1.Run the multiscale Gaussian wavepacket transform to get the dictionary of central points y and central momentum p + q. 2. Separate p+q into two different wavepackets and shoot them from the subsurface point y to the acquisition surface. 3. Angle-gather prestack image by calculating Ipq(y, ω) =(cid:82)(cid:82) dsdr ¯ˆU y p ¯˜U y q D(r, s, ω). 4. Stack all partial image Ipq with the same p + q to reduce the noise. 5. Use equation (3.48) to get the coefficient of each wavepacket expansion of the perturbation 2α v2 . 6. Run the inverse multiscale Gaussian wavepacket transform to recover the rapid pertur- bation 2α v2 . 3.2.5 Motivation for Inverting the multiscale Gaussian-beam transform Hereby we provide some intuitive justifications of the inversion process of the multiscale Gaussian-beam transform, which may provide some theoretical guideline for carrying out further analysis of our new methodology and extending our methodology to other applications. We start with considering the function b(r, s, t) defined by the linear operator D, D[f ](r, s, t) = b(r, s, t), (3.50) where the function f is defined at subsurface points. To solve this linear operator 37 equation rapidly and efficiently, we would like to diagonalize the operator D by carrying out a certain frame representation. Since the argument f of the operator D does not sit in the same space as the right-hand side b, we first apply the adjoint operator D∗ to both sides, so that we have D∗D[f ] = D∗b. (3.51) Let W be the multiscale Gaussian wavepacket transform defined in Section 2.2, which satisfies W∗W = I. Then we apply W to both sides of equation (3.51), yielding W(D∗D)W∗Wf = WD∗b. (3.52) Our results in Section 3.2.4 indicate that the above diagonalization is justified. The operator WD∗b is essentially equation (3.46), and the operator W(D∗D)W∗ is essen- tially captured by the diagonal factor E(p, q, y) in equation (3.47). The overall effects of the diagonalization process are epitomized in equation (3.48). Therefore, we have Wf = (W(D∗D)W∗) −1 WD∗b, f = W∗ (W(D∗D)W∗) −1 WD∗b, (3.53) which results in our fast reconstruction formula (3.49). To establish the theoretical foundation for our new methodology in terms of the Fourier Integral Operator (FIO) theory, we need to carry out symbolic calculus to establish several facts by following the works in [50, 23]: 38 1. the forward operator D belonging to a certain class of FIO; 2. the conjugation process W(D∗D)W yielding a diagonal operator in the frame de- fined by W. On the other hand, since the FIO theory is originated from the asymptotic of geometric optics and a Gaussian-beam solution provides a globally defined asymptotic solution for wave equations, we will carry out brute-force calculations to justify our new methodol- ogy by heavily relying on the structure of multiscale Gaussian-beam transform, which is composed of Gaussian beams and multiscale Gaussian wavepacket transform. 3.3 Theoretical Validation: The Proof about The Imaging Operator 3.3.1 Road Map of The Theoretical Analysis We will prove the approximation (3.47) holds. Since it involves the interaction of four beams in time, we will carry out the analysis essentially in two main steps. The first step consists of analysis of the following two integrals dealing with the beam interactions on the receiver side and the source side respectively, (cid:90)(cid:90) drdξ ¯ˆU y p (r, ω) ˆU x ξ (r, ω), (cid:90)(cid:90) dsdη ¯˜U y q (s, ω) ˜U x η (s, ω). Since receivers and sources are reciprocal in wave propagation, we just need to focus on analysis of beam interaction on the receiver side, and the analysis of the source side 39 is essentially analogous. Our analysis of the two beams’ interaction yields the following approximations (cid:90)(cid:90) (cid:90)(cid:90) drdξ ¯ˆU y p (r, ω) ˆU x dsdη ¯˜U y q (s, ω) ˜U x ξ (r, ω) ≈ eip·(y−x) ˆH(x, ξ, ω; y, p), η (s, ω) ≈ eiq·(y−x) ˜H(x, η, ω; y, q). where the functions ˆH and ˜H will be defined later. Based on the first step, the second step consists of analyzing the left-hand side of approximation (3.47) so that we have (cid:90)(cid:90)(cid:90)(cid:90) ξ (r, ω) ˜U x ≈ ei(p+q)(y−x) ˆH(x, ξ, ω; y, p) ˜H(x, η, ω; y, q). ¯ˆU y p (r, ω) ¯˜U y q (s, ω) ˆU x drdsdξdη η (s, ω) After carrying out the integral about ω, the above approximation reduces to (cid:90)(cid:90)(cid:90)(cid:90) − q (s, ω) ˆU x ei(p+q)(y−x)K(p, q, y) ˆH(x, y, p, q) ˜H(x, y, p, q). dωdrdsdξdηω2 ¯ˆU y p (r, ω) ¯˜U y ξ (r, ω) ˜U x η (s, ω) = (3.54) (3.55) where K(p, q, y), ˆH and ˜H are defined later. After making an essential assumption on the invertibility of the imaging operator, we 40 are able to approximate functions ˆH and ˜H by ˆH(x, y, p, q) ≈ e ˜H(x, y, p, q) ≈ e i 4||y−x||2 ˆM (0) ˆK(y, p, q), i 4||y−x||2 ˜M (0) ˜K(y, p, q), where ˆK(y, p, q) and ˜K(y, p, q) can be easily computed. These latter approximations allow us to obtain our main theorem, (cid:90) (cid:90) (cid:90) −ω2dω drds ¯ˆU y p (r, ω) ¯˜U y q (s, ω) dξdη ˆU x ξ (r, ω) ˜U x η (s, ω) ≈ K(p, q, y) ˆK(y, p, q) ˜K(y, p, q)ei(p+q)·(y−x)e i||y−x||2 M0/2, which says that the four-beam interaction in time yields a weighted Gaussian wavepacket centered at the scattering point y in the direction p + q, where (y, p) is the ray param- eter for the beam from the scattering point y to the boundary receiver in the direction p, and (y, q) is the ray parameter for the beam from the scattering point y to the boundary source in the direction q. The rest of the analysis will follow the above road map. We will prove equation (3.47) in this section. Throughout the proof, the amplitude function A is not involved as it is more smooth compared with the phase function part. Denote the beam as (ˆy(t), ˆp(t), ˆM (t), ˆA(t)), whose initial value is (y, p, M, A), and (ˆx(t), ˆξ(t), ˆN (t), ˆC(t)) whose initial value is (x, ξ, N, C). Moreover, ˆΞ and ˆκ are 41 defined according to a fixed beam ˆU y p , ˆξ(t) = ˆκ(t; x, ξ, y, p)ˆp(t) + ˆΞ(t; x, ξ, y, p), (3.56) and ˆΞ(t; x, ξ, y, p) · ˆp(t) = 0. (3.57) On the source side, we have similar notation (˜y(t), ˜q(t), ˜M (t), ˜A(t)), whose initial value is (y, q, M, A), and (˜x(t), ˜η(t), ˜N (t), ˜C(t)) whose initial value is (x, η, N, C). 3.3.2 Approximation of Gaussian Beams along the Surface The beam functions ˆU y p (r, ω) and ˜U y p (s, ω) are used in the inversion process to link the data to the unknown perturbation. We would like to explore more about the beam function’s structure to build the foundation for the future proof and calculation. ||x||2 M denotes xT M x in this paper for all vectors x and all symmetric matrices M . For each beam ˆU y p , we define the hitting time ˆt0 = ˆt0(y, p) and hitting point ˆr0 = ˆy(ˆt0(y, p)) according to the arrival time of its central ray at the boundary so that ˆr0 = ˆy(ˆt0(y, p)) ∈ {x ∈ Rd : xd = 0}. In [49], the single beam is propagating by treating the time variable same as the spatial components, see also [24]. The complete Hessian matrix is S.P.D. along the directions transversal to the ray direction. Therefore, by Assumption 3.2.1, we have a Gaussian profile if intersecting the beam solution at the surface xd = 0. Proposition 3.3.1. The Gaussian beam ˆU y p (r, t) along the surface for r ∈ {x = 42 (x1,··· ,d) : xd = 0} can be represented as p (r, t) ≈ ˆA(ˆt0) exp(cid:0)i(cid:0)ˆτx(ˆt0; y, p) · (r − ˆr0) + ˆτt(ˆt0; y, p)(t − ˆt0)(cid:1)(cid:1)× ||r − ˆr0||2 i ˆτtt(ˆt0; y, p)(t − ˆt0)2 + (t − ˆt0)ˆτ T + ˆU y exp 1 2 ˆM (ˆt0) 2 tx(ˆt0; y, p)(r − ˆr0)  . (3.58) where ˆt0 = ˆt0(y, p) and all partial derivatives about the phase function τ are on the central ray. The proposition is equivalent to intersecting the complete (t, x) beam ansatz at the surface xd = 0. We first introduce the way to compute the terms ˆτtx and ˆτtt, which can be obtained by inserting corresponding ray parameters into equations (4.13)-(4.14). ˆτt(t; y, p) = − ˆG(ˆy(t), ˆp(t)), ˆτtx(t; y, p) = − ˆGx(ˆy(t), ˆp(t)) − ˆM (t; y, p) ˆGp(ˆy(t), ˆp(t)), ˆτtt(t; y, p) = − ˆGp(ˆy(t), ˆp(t)) · ˆτtx(t; y, p), (3.59) (3.60) (3.61) where (ˆy(t), ˆp(t)) = (ˆy(t; y, p), ˆp(t; y, p)). We also denote ˆM∗ as the complete Hessian 43 matrix about (t, x) at (t; y, p) and t = ˆt0(y, p),  ˆτtt(t; y, p), ˆτ T ˆτtx(t; y, p), tx(t; y, p) ˆM (t; y, p)  , ˆM∗(t; y, p) = ˆF (r, t; y, p) = ˆτt(t; y, p) + Re(ˆτtx(t; y, p)) · (r − ˆy(t)), ˆθ(r, t; y, p) = ˆp(t) · (r − ˆy(t)) + ||r − ˆy(t)||2 ˆQ(r, t; y, p) = −(Im(ˆτtx)(t; y, p))T (r − ˆy(t)) Re( ˆM (t)) 2 . , Im(ˆτtt)(t; y, p) Similarly, we denote (ˆx(t), ˆξ(t)) = (ˆx(t; x, ξ), ˆξ(t; x, ξ)), ˆN∗(t; x, ξ), ˆF (r, t; x, ξ), ˆθ(r, t; x, ξ) and ˆQ(r, t; x, ξ). The following proposition is needed when taking the Fourier transform of ˆU y p (r, t) about time t, Proposition 3.3.2. For any complex number γ with positive real part, i.e. Re(γ) > 0, (cid:90) ∞ −∞ e−γt2 (cid:114) π γ e−iωtdt = − ω2 4γ . e (3.62) Lemma 3.3.1. The Fourier transform of ˆU y p (r, t) with respect to time t is ˆU y p (r, ω) = ˆU y p (r, ω; t) = eiˆ(r,t;y,p)e−iωtei ˆβ(t;y,p)|ω−ˆτt(t;y,p)−ˆζ(t;y,p)T (r−ˆy(t))|2 (cid:115) −||r−ˆy(t)||2 ˆM(t;y,p,M0) i2π ˆτtt(t; y, p) 2 , (3.63) (cid:12)(cid:12)(cid:12)t=ˆt0(y,p) ˆA(t)e 44 where ˆM(t; y, p) = Im( ˆM )(t; y, p) − Im(ˆτtx)Im(ˆτtx)T Im(ˆτtt(t; y, p)) ˆβ(t; y, p) = − 1 2ˆτtt(t; y, p) = − 1 2iIm(ˆτtt(t; y, p)) + 2Re(ˆτtt(t; y, p)) (3.64) , (3.65) ˆζ(t; y, p) = Re(ˆτtx(t; y, p)) − Re(ˆτtt(t; y, p)) Im(ˆτtt(t; y, p)) Im(ˆτtx(t; y, p)), (3.66) ˆ (r, t; y, p) = ˆθ(r, t; y, p) + ( ˆF (r, t; y, p) − ω) ˆQ(r, t; y, p) + Re(ˆτtt(t; y, p)) 2 ˆQ(r, t; y, p)2. (3.67) Here t in ˆU y p (·; t) serves as a fixed parameter, since all terms defined above is defined at this fixed moment. Proof. We still abbreviate ˆt0(y, p) as ˆt0 in this proof since we only concern the single beam ˆU y p here. From Proposition 3.3.1, (cid:90) p (r, ω) ≈ ˆA(ˆt0)eiˆθ(r,ˆt0;y,p) ˆU y −||(t−ˆt0,r−ˆy(ˆt0))||2 × e e−iωtei ˆF (r,ˆt0;y,p)(t−ˆt0)ei 1 Im( ˆM∗(ˆt0;y,p,M0))/2dt, 2 Re(ˆτtt(ˆt0;y,p))(t−ˆt0)2 After expanding the term e −||(t−ˆt0,r−ˆy(ˆt0))||2 Im( ˆM∗/2), we have −||(0,r−ˆy(ˆt0))||2 Im( ˆM∗)(ˆt0;y,p,M0) 2 2 Re(ˆτtt(ˆt0;y,p))(t−ˆt0)2 − Im(ˆτtt(ˆt0;y,p)) e 2 (t−ˆt0)2 (cid:90) p (r, ω) ≈ ˆA(ˆt0)eiˆθe−iωˆt0e ˆU y e−i(− ˆF +ω)(t−ˆt0)ei 1 e−(t−ˆt0)(Im(ˆτtx)(ˆt0;y,p))T (r−ˆy(ˆt0))dt. 45 (3.68) (3.69) To make a complete square term in equation (3.69), we have − Im(ˆτtt)(ˆt0; y, p) |(Im(ˆτtx)(ˆt0; y, p))T (r − ˆy(ˆt0))|2 2 2Im(ˆτtt(ˆt0; y, p)) (cid:32) (t − ˆt0)2 − (t − ˆt0)Im(ˆτtx(ˆt0; y, p))T (r − ˆy(ˆt0)) = − Im(ˆτtt) 2 t − ˆt0 + (Im(ˆτtx))T (r − ˆy(ˆt0)) Im(ˆτtt) (cid:33)2 . Since ˆQ(r, ˆt0; y, p) = − (Im(ˆτtx)(ˆt0;y,p))T (r−ˆy(ˆt0)) Im(ˆτtt)(ˆt0;y,p) , p (r, ω) = ˆA(ˆt0)eiˆθe−iωˆt0ei( ˆF−ω) ˆQei 1 ˆU y |(Im(ˆτtx)(ˆt0;y,p))T (r−ˆy(ˆt0))|2 −||r−ˆy(ˆt0)||2 2 Re(ˆτtt) ˆQ2 e 2Im(ˆτtt)(ˆt0;y,p) (cid:90) ei( ˆF−ω+Re(ˆτtt) ˆQ)(t−ˆt0− ˆQ) Im( ˆM )(ˆt0) 2 e 2 2 Re(ˆτtt)(t−ˆt0− ˆQ)2 ei 1 e − Im(ˆτtt)(ˆt0;y,p) ≈ ei ˆF ˆQei 1 2 Re(ˆτtt)( ˆQ)2 e−iω(ˆt0+ ˆQ)e (cid:115) (t−ˆt0− ˆQ)2 dt + O(|t − t0|2) |(Im(ˆτtx)(ˆt0;y,p))T (r−ˆy(ˆt0))|2 2Im(ˆτtt)(ˆt0;y,p) −||r−ˆy(ˆt0)||2 Im( ˆM )(ˆt0) 2 e × i2π ˆτtt(ˆt0; y, p) ˆA(ˆt0)eiˆθe−ˆ|−ˆτt+ω−Re(ˆτtx)(ˆt0;y,p)·(r−ˆy(ˆt0))−Re(ˆτtt)(ˆt0;y,p) ˆQ|2 . (3.70) Equation (3.70) is obtained by Proposition 3.3.2 where ˆ(ˆt0; y, p) = 2Im(ˆτtt)(ˆt0; y, p) − i2Re(ˆτtt)(ˆt0; y, p) 1 . (3.71) After replacing ˆβ = iˆ, we have the lemma proved. 46 Corollary 3.3.1. The Fourier transform of ˆU x ξ (r, t) with respect to time t is ˆU x ξ (r, ω) = ˆU x ξ (r, ω; t) (cid:12)(cid:12)(cid:12)t=ˆt0(x,ξ) = eiˆ(r,t;x,ξ)e−iωteiˆγ(t;x,ξ)|ω−ˆτt(t;x,ξ)− ˆϑ(t;x,ξ)T (r−ˆx(t))|2 (cid:115) −||r−ˆx(t)||2 ˆN (t;x,ξ,N0) i2π ˆC(t)e 2 , (3.72) ˆτtt(t; x, ξ) where ˆN (t; x, ξ) = Im( ˆN )(t; x, ξ) − Im(ˆτtx)Im(ˆτtx)T Im(ˆτtt(t; x, ξ)) ˆ (r, t; x, ξ) = ˆθ(r, t; x, ξ) + ( ˆF (r, t; x, ξ) − ω) ˆQ(r, t; x, ξ) + 1 2 ˆγ(t; x, ξ) = − Re(ˆτtt(t; x, ξ)) ˆQ(r, t; x, ξ)2, 1 2iIm(ˆτtt(t; x, ξ)) + 2Re(ˆτtt(t; x, ξ)) , (3.73) (3.74) ˆϑ(t; x, ξ) = Re(ˆτtx(t; x, ξ)) − Re(ˆτtt(t; x, ξ)) Im(ˆτtt(t; x, ξ)) Im(ˆτtx(t; x, ξ)). (3.75) We have the same conclusion for ˜U y q (s, ω) and ˜U x η (s, ω) as Lemma 3.3.1 and all terms involved are defined accordingly. 3.3.3 Asymptotic Analysis of Two Beams’ Interaction In this section, we would like to explore the interaction between two Gaussian beams, that is (cid:90)(cid:90) (cid:90)(cid:90) dsdη ¯˜U y q (s, ω) ˜U x η (s, ω). drdξ ¯ˆU y p (r, ω) ˆU x ξ (r, ω), 47 The rest of Section 3.3.3 is organized as the following. We will first discuss the distance between two beams ˆU y p and ˆU x space. In Section 3.3.3.2, we will evaluate the error caused by replacing ˆU x ξ satisfying the parabolic scaling principle in the phase ξ (·; ˆt0(x, ξ)) ξ (·; ˆt0(y, p)). After that, the difference between the exponents at different times with ˆU x is evaluated and it will allow us to map the phase term to the initial moment. In Section 3.3.3.4 and 3.3.3.5, we will compute the integral about r and show that there exists a Gaussian profile centered at ˆy(ˆt0(y, p)). And the analysis can all be applied to the source side similarly. Here we require that there is a significant interaction between two beams concerned, which means the distance between two central rays, (ˆy(t), ˆp(t)) and (ˆx(t), ˆξ(t)) is less than the width of the beam ˆU y √ p and the width of a beam is defined as 1/ , where  is the smallest eigenvalue of Im( ˆM (t)). 3.3.3.1 Parabolic Scaling Principle A wavepacket satisfying the so-called parabolic scaling principle means the wavelength of the typical oscillation of the wavepacket being equal to the square of the width of the wavepacket, and a Gaussian beam will satisfy parabolic scaling principle at any given time if it does initially. The following graph shows a single Gaussian wavepacket in R2 satisfying the parabolic scaling principle. The following asymptotic analysis is needed throughout the proof. Lemma 3.3.2. Consider two scattering beams (ˆy(t), ˆp(t), ˆM (t), ˆA(t)) and 48 Figure 3.3: Real Part of Single Wavepacket φl,i,k (ˆx(t), ˆξ(t), ˆN (t), ˆC(t)) satisfying the parabolic scaling principle at the initial time, then || ˆM (t)|| ∼ O(||p||), || ˆN (t)|| ∼ O(||ξ||). (3.76) If there exists significant interaction effects between two beams, then (cid:33) 1(cid:112)||p|| Moreover, if we have ||p − ξ|| ∼ O((cid:112)||p||) at the beginning, ||ˆy(t) − ˆx(t)|| ∼ O (cid:32) . |ˆκ(t)| ∼ 1 + O(||p||− 1 2 ), ||ˆΞ(t)|| ∼ O((cid:112)||p||). (3.77) (3.78) Proof. The assumption ||p− ξ|| ∼ O((cid:112)||p||) is a reasonable assumption, as we will see later ||p − ξ|| will be controlled by a Gaussian profile, which means the value will be exponentially decaying when p and ξ are far from each other. First, the Hamiltonian G(x, p) = v(x)||p|| remains as a constant along the ray. This implies that the order of the momentum ||p|| will not change as the velocity v is a smooth function and bounded away from zero, that is ||ˆp(t)|| ∼ ||p||. 49 -2101x220.5x10.500 Second, we will evaluate ˆM (t). The homogeneous of degree one Hamiltonian guarantees that Gaussian wavepackets satisfy the parabolic scaling principle at any given time. This implies that the size of Hessian ˆM is around O(||ˆp(t)||) and equation (3.76) is correct. Due to the fact that ˆU y p is well-localized around ˆy(t) in the physical space, ˆU x ξ will have small interaction with ˆU y p , if ˆx(t) is beyond this localized region. And equation (3.77) is correct. Third, (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)d(ˆp(t) − ˆξ(t)) dt (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)∇∇v(ˆy(t))(ˆy(t) − ˆx(t))||ˆp(t)||(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:19)T (cid:18) ˆp(t) ||ˆp(t)|| + ||∇v(ˆy(t)) (ˆp(t) − ˆξ(t))|| ≤ C2||ˆy(t) − ˆx(t)||||ˆp(t)|| + C1||ˆp(t) − ˆξ(t)||, where C1 and C2 are the upper bound of ||∇v|| and ||∇∇v|| respectively. Moreover, d||ˆp(t) − ˆξ(t)||2 dt d(ˆp(t) − ˆξ(t)) dt = 2 · (ˆp(t) − ˆξ(t)) ≤ 2||d(ˆp(t) − ˆξ(t)) dt ||||ˆp(t) − ˆξ(t)||, by Cauchy-Schwartz inequality. We further get, d||ˆp(t) − ˆξ(t)||2 dt ≤ C1||ˆp(t) − ˆξ(t)||2 + C2(||ˆy(t) − ˆx(t)||)(||ˆp(t)||)(||ˆp(t) − ˆξ(t)||) ||ˆp(t) − ˆξ(t)||2 2 . ≤ C1||ˆp(t) − ˆξ(t)||2 + 1 2 (C2||ˆp(t)||||ˆy(t) − ˆx(t)||)2 + 50 By Gronwall inequality, we have ||ˆp(t) − ˆξ(t)||2 ∼ O(||p||), since C2||ˆp(t)||||ˆy(t) − ˆx(t)|| is uniformly bounded. Consequently, equation (3.78) is correct. 3.3.3.2 Difference between Two Interacted Beams’ Traveltime In this part, we would like to calibrate the beam ˆU x ξ (r, ω; ˆt0(x, ξ)) according to the beam ˆU y p (r, ω) by shifting time ˆt0(x, ξ) to time ˆt0(y, p). By Corollary 3.3.1, ξ (r, ω) = eiˆ(r,ˆt0(x,ξ);x,ξ)e−iωˆt0(x,ξ) ˆU x eiˆγ(ˆt0(x,ξ);x,ξ)|ω−ˆτt(ˆt0(x,ξ);x,ξ)− ˆϑ(ˆt0(x,ξ);x,ξ)T (r−ˆx(ˆt0(x,ξ)))|2 (cid:115) −||r−ˆx(ˆt0(x,ξ))||2 ˆN (ˆt0(x,ξ);x,ξ,N0) i2π ˆτtt(ˆt0(x, ξ); x, ξ) ˆC(ˆt0(x, ξ))e The difference |ˆt0(x, ξ) − ˆt0(y, p)| is around O bound can be obtained using Lemma 3.3.2, 2 . (3.79) (cid:18) 1(cid:112)||p|| (cid:19) . To see this, the following |ˆyd(t) − ˆxd(t)| ≤ ||ˆy(t) − ˆx(t)|| ∼ O( 1(cid:112)||p||). (3.80) 51 By Assumption 3.2.1, we have |ˆt0(x, ξ) − ˆt0(y, p)| ≤ |ˆyd(t) − ˆxd(t)| b ∼ O (cid:32) (cid:33) . 1(cid:112)||p|| (3.81) The following proof is essentially comparing each term of equation (3.79) at two dif- ferent times ˆt0(x, ξ) and ˆt0(y, p). First, e−iωˆt0(x,ξ) becomes, e−iωˆt0(x,ξ) = e−iωˆt0(y,p)e−iω(ˆt0(x,ξ)−ˆt0(y,p)). (3.82) Second, we will discuss eiˆγ(ˆt0(x,ξ);x,ξ)|ω−ˆτt(ˆt0(x,ξ);x,ξ)− ˆϑ(ˆt0(x,ξ);x,ξ)T (r−ˆx(ˆt0))|2 . Proposition 3.3.3. iˆγ(ˆt0(x, ξ); x, ξ)|ω − ˆτt(ˆt0(x, ξ); x, ξ) − ˆϑ(ˆt0(x, ξ); x, ξ)T (r − ˆx(ˆt0(x, ξ)))|2 = (r − ˆx(ˆt0(y, p)))|2 iˆγ(ˆt0(y, p); x, ξ)|ω − ˆτt(ˆt0(y, p); x, ξ) −(cid:16) ˆϑ(ˆt0(y, p); x, ξ)) (cid:17)T (cid:18) 1 (cid:19) (3.83) (cid:33) (cid:32) 1(cid:112)||p|| . (3.84) + O ||p|| . Proof. See Appendix A. Third, we will discuss ||r − ˆx(ˆt0(x, ξ))||2 ˆN . Proposition 3.3.4. ||r − ˆx(ˆt0(x, ξ))||2 ˆN (ˆt0(x,ξ)) = ||r − ˆx(ˆt0(y, p))||2 ˆN (ˆt0(y,p)) + O 52 Proof. According to Lemma A.0.1, we have, ˆN (ˆt0(x, ξ); x, ξ) = Im( ˆN )(ˆt0(x, ξ); x, ξ) − Im( ˆN ) ˆξ ˆξT Im( ˆN ) ˆξT Im( ˆN ) ˆξ (ˆt0(x, ξ); x, ξ)). Then, ˆN (ˆt0(x, ξ); x, ξ) ˆξ(ˆt0(x, ξ)) = Im( ˆN )(ˆt0(x, ξ); x, ξ) ˆξ(ˆt0(x, ξ)) − Im( ˆN )(ˆt0(x, ξ); x, ξ) ˆξ(ˆt0(x, ξ)) = 0. (3.85) This is also correct for any vector parallel to ˆξ(ˆt0(x, ξ)). On the other hand, r − ˆx(ˆt0(x, ξ)) = r − ˆx(ˆt0(y, p)) + ˆx(ˆt0(y, p)) − ˆx(ˆt0(x, ξ)) = r − ˆx(ˆt0(y, p)) + G± = r − ˆx(ˆt0(y, p)) ± v(ˆx(ˆt0(y, p)))(ˆt0(y, p) − ˆt0(x, ξ)) p (ˆx(ˆt0(y, p)), ˆξ(ˆt0(y, p)))(ˆt0(y, p) − ˆt0(x, ξ)) ˆξ(ˆt0(x, ξ)) || ˆξ(ˆt0(x, ξ))||. (3.86) Consequently, ||r − ˆx(ˆt0(x, ξ))||2 = ||r − ˆx(ˆt0(y, p))||2 ˆN (ˆt0(x,ξ)) ˆN (ˆt0(y,p)) = ||r − ˆx(ˆt0(y, p))||2 ˆN (ˆt0(y,p)) + O + ||r − ˆx(ˆt0(y, p))||2 (cid:32) (cid:33) 1(cid:112)||p|| = ||r − ˆx(ˆt0(y, p)) + λ ˆξ(ˆt0(x, ξ))||2 ˆN (ˆt0(x,ξ)) ˆN (ˆt0(x,ξ))− ˆN (ˆt0(y,p)) . (3.87) Finally, about ˆ(r, ˆt0(x, ξ); x, ξ) defined in equation (3.74). All other terms are at 53 constant order, except for ˆξ(ˆt0(x, ξ)) · (r − ˆx(ˆt0(x, ξ))), ˆξ(ˆt0(x, ξ)) · (r − ˆx(ˆt0(x, ξ))) = ˆξ(ˆt0(y, p)) · (r − ˆx(ˆt0(x, ξ))) = ˆξ(ˆt0(y, p)) ·(cid:0)r − ˆx(ˆt0(y, p))(cid:1) ± v(ˆx(ˆt0(y, p)))(ˆt0(y, p) − ˆt0(x, ξ))|| ˆξ(ˆt0(y, p))||) = ˆξ(ˆt0(y, p)) ·(cid:0)r − ˆx(ˆt0(y, p))(cid:1) ± v||ˆp(ˆt0(y, p))||(ˆt0(y, p) − ˆt0(x, ξ)) + O(1) = ˆξ(ˆt0(y, p)) · (r − ˆx(ˆt0(y, p)) − ˆτt(ˆt0(y, p); y, p)(ˆt0(y, p) − ˆt0(x, ξ)) + O(1). (3.88) The asymptotic analysis in the last step comes from Lemma 3.3.2, that is ||ˆp(t) − ˆξ(t)|| ∼ O((cid:112)||p||). To expedite the discussion, we will use the following notations: ˆtc = ˆt0(y, p), ∆ˆt0(x, ξ; y, p) = ˆt0(y, p) − ˆt0(x, ξ), ˜tc = ˜t0(y, q), ∆˜t0(x, η; y, q) = ˜t0(y, q) − ˜t0(x, η). Lemma 3.3.3. By Proposition 3.3.3, Proposition 3.3.4 and equation (3.82) and (3.88), i2π ˆτtt(ˆtc; x, ξ) ˆU x ξ (r, ω; ˆt0(x, ξ)) = ei(ω−ˆτt(ˆtc;x,ξ))∆ˆt0(x,ξ;y,p)ei ˆξ(ˆtc)·(r−ˆx(ˆtc))eiˆγ(ˆtc;x,ξ)|ω−ˆτt(ˆtc;x,ξ)− ˆϑ(ˆtc;x,ξ)T (r−ˆx(ˆtc;x,ξ))|2 −||r−ˆx(ˆtc;x,ξ)||2 e ˆC(ˆtc)eiˆ(r,ˆtc;x,ξ)−i ˆξ(ˆt0(x,ξ))·(r−ˆx(ˆt0(x,ξ)))e−iωˆtc (cid:33) ˆN (ˆtc) 2 + O (3.89) (cid:115) (cid:32) 1(cid:112)||p|| . Now, since two beams are using the same traveltime, we will abbreviate parameter ˆt0 54 in ˆU x ξ (r, ω; ˆt0(x, ξ)). On the source side, Corollary 3.3.2. When we calibrate the beam ˜U x η (s, ω; ˜t0(x, η)) according to the beam ˜U y q (s, ω) by shifting time ˜t0(x, η) to time ˜tc, we have (cid:115) ˜U x η (s, ω; ˜t0(x, η)) = i2π ˜τtt(˜tc; x, η) ei(ω−˜τt(˜tc;x,η))∆˜t0(x,η;y,q)ei˜η(˜tc)·(s−˜x(˜tc)) ei˜γ(˜tc;x,η)|ω−˜τt(˜tc;x,η)− ˜ϑ(˜tc;x,η)T (s−˜x(˜tc;x,η))|2 ˜C(˜tc)ei˜(s,˜tc;x,η)−i˜η(˜t0(x,η))·(s−˜x(˜t0(x,η)))e−iω˜tc (cid:33) (cid:32) −||s−˜x(˜tc;x,η)||2 e ˜N (˜tc) 2 + O . 1(cid:112)||q|| (3.90) 3.3.3.3 Difference between Two Interacted Beams’ Phase and Hessians We will compare the difference between ˆM (t) − ˆN (t) in this section. We first have the following inequality Proposition 3.3.5. Consider two scattering beams (ˆy(t), ˆp(t), ˆM (t), ˆA(t)) and (ˆx(t), ˆξ(t), ˆN (t), ˆC(t)), and assume that there exists significant interaction effects be- tween these two beams. There exists two bounded positive constants C∗ 1 and C∗ 2 related to the background velocity, such that d|| ˆM (t) − ˆN (t)|| dt ≤ C∗ 1 (cid:112)||p|| + C∗ 2|| ˆM (t) − ˆN (t)||. (3.91) where || ˆM (t) − ˆN (t)|| is defined as the matrix norm induced by the vector 2-norm. Proof. See Appendix A. 55 Lemma 3.3.4. Consider two scattering beams (ˆy(t), ˆp(t), ˆM (t), ˆA(t)) and (ˆx(t), ˆξ(t), ˆN (t), ˆC(t)), and there exists significant interaction effects between these two beams. Then || ˆM (t) − ˆN (t)|| ∼ O (cid:16)(cid:112)||p||(cid:17) , ∀t ∈ [0, T ], (3.92) Proof. First, ˆM (t)− ˆN (t) is zero at t = 0, since they satisfy the same initial condition. Using Proposition 3.3.5 and the fact that the norm || ˆM (t)− ˆN (t)|| is positive and both C∗ 1 and C∗ 2 are positive, || ˆM (t) − ˆN (t)|| ≤ C∗ 1 C∗ 2|| ˆM (s) − ˆN (s)||ds, (3.93) (cid:112)||p(0)||T + (cid:90) t 0 since the boundary data D(r, s, t) is measured in the time interval [0, T ]. With Gronwall inequality, || ˆM (t) − ˆN (t)|| ≤ C∗ 2 t(cid:112)||p||. 1 T eC∗ (3.94) Moreover, we have the same conclusion for other related terms, Corollary 3.3.3. || ˆM(t) − ˆN (t)|| ∼ O((cid:112)||p||), ||ˆτtx(ˆtc; y, p) − ˆτtx(ˆtc; x, ξ)|| ∼ O((cid:112)||p||), |ˆτtt(ˆtc; y, p) − ˆτtt(ˆtc; x, ξ)| ∼ O((cid:112)||p||), || ˜M(t) − ˜N (t)|| ∼ O((cid:112)||q||); ||˜τtx(˜tc; y, q) − ˆτtx(˜tc; x, η)|| ∼ O((cid:112)||q||); |˜τtt(˜tc; y, q) − ˜τtt(˜tc; x, η)| ∼ O((cid:112)||q||). Consequently, by Lemma A.0.1, both ˆU y p (r, ω) and ˆU x ξ (r, ω) are well-localized along 56 the boundary, and we have 1 2 ||r − ˆx(t)||2 ˆM− ˆN ∼ O(||p||− 1 2 ), which will be used in equation (3.101) when we replace ˆN with ˆM. Similarly, ||s − ˜x(t)||2 ˜M− ˜N ∼ O(||q||− 1 2 ). 1 2 Lemma 3.3.5. Consider two scattering beams (ˆy(t), ˆp(t), ˆM (t), ˆA(t)) and (ˆx(t), ˆξ(t), ˆN (t), ˆC(t)), and there exists significant interaction effects between these two beams. Suppose the function g(t) is g(t) = ˆp(t) · (ˆy(t) − ˆx(t)), g(t) = g(0) + O(1), (cid:32) (cid:33) . ||ˆΞ(t)||2 ˆκ(t)2||ˆp(t)|| 1 2 (cid:48) g (t) = v(ˆx(t)) then we have and Proof. See Appendix A. (3.95) (3.96) (3.97) Lemma 3.3.6. Consider two scattering beams (ˆy(t), ˆp(t), ˆM (t), ˆA(t)) and (ˆx(t), ˆξ(t), ˆN (t), ˆC(t)), and there exists significant interaction effects between these two beams. Suppose the pure imaginary matrix ˆM (0) has a symmetric positive definite 57 imaginary part and is the initial condition of the Hessian for the beam, then (y − x)T ˆM (0)(y − x) = (ˆy(t) − ˆx(t))T ˆM (t)(ˆy(t) − ˆx(t)) + O(1). (3.98) Proof. See Appendix A. 3.3.3.4 Approximation of Two Beams’ Interaction explicit formula of(cid:82)(cid:82) dξdr In this section, we will use the conclusion obtained in previous sections to get the ¯ˆU y p (r, ω) ˆU x ξ (r, ω). We first have the proposition below which will be used in approximation, Proposition 3.3.6. The real-valued phase terms, ˆ(r, ˆtc; y, p)−ˆθ(r, ˆtc; y, p) and ˆ(r, ˆtc; x, ξ)− ˆθ(r, ˆtc; x, ξ), can be ignored since they are constant order terms with respect to the large wavenumber ||ξl,i|| = ||p + q||. Proof. See Appendix A. 58 According to equation (3.89), (3.63), ¯ˆU y p (r, ω) ˆU x ξ (r, ω) = eiO(1) (cid:90)(cid:90) (cid:90)(cid:90) eiˆ(r,ˆtc;x,ξ)−iˆθ(r,ˆtc;x,ξ)eiˆθ(r,ˆtc;x,ξ)eiˆγ|ω−ˆτt(ˆtc;x,ξ)− ˆϑ(r−ˆx(ˆtc))|2 −||r−ˆx(ˆtc)||2 e dξdr drdξeiωˆtc−iωˆtce−iˆ(r,ˆtc;y,p)+iˆθ(r,ˆtc;y,p)e−iˆθ(r,ˆtc;y,p)e(i ˆβ)|ω−ˆτt(ˆtc;y,p)−ˆζ(r−ˆy(ˆtc))|2 ei(ω−ˆτt(ˆtc;y,p))∆ˆt0(x,ξ;y,p) −||r−ˆy(ˆtc)||2 ˆM(ˆtc) ˆN (ˆtc) e 2 2 (cid:90) (cid:16) × e where 2||ˆy(ˆtc)−ˆx(ˆtc)||2 i 1 ˆM(ˆtc)e Re( ˆM )(ˆtc) eiˆγ|ω−ˆτt(ˆtc;x,ξ)− ˆϑT (r−ˆx(ˆtc))|2 − 1 4||ˆy(ˆtc)−ˆx(ˆtc)||2 i ˆβ dre = eiO(1) dξe−iωˆtc+iωˆtce (cid:17)|ω−ˆτt(ˆtc;y,p)−ˆζT (r−ˆy(ˆtc))|2 (cid:90) × × ei ˆξ(ˆtc)·(r−ˆx(ˆtc))−iˆp(ˆtc)·(r−ˆy(ˆtc))ei(ω−ˆτt(ˆtc;y,p))∆ˆt0(x,ξ;y,p) 4||ˆy(ˆtc)−ˆx(ˆtc)||2 − 1 2||r−ˆx(ˆtc)||2 − 1 2||r−ˆy(ˆtc)||2 1 ˆM(ˆtc)e ˆN (ˆtc)e ˆM(ˆtc), ei ˆχ(r,ˆtc;y,p,x,ξ) (3.99) ||ˆy(ˆtc) − ˆx(ˆtc)||2 ˆχ(r, ˆtc; y, p, x, ξ) = −1 2 ||r − ˆx(ˆtc)||2 1 2 + Re( ˆN )(ˆtc) Re( ˆM )(ˆtc) − 1 2 ||r − ˆy(ˆtc)||2 Re( ˆM )(ˆtc) . (3.100) Here we neglect some constant order real-valued phase terms by Proposition 3.3.6, which can be considered as a smooth residual. 59 Then we replace ˆN with ˆM, and the inner integral becomes, (cid:90) dre(i ˆβ)|ω−ˆτt(ˆtc;y,p)−ˆζT (r−ˆy(ˆtc))|2 eiˆγ|ω−ˆτt(ˆtc;x,ξ)− ˆϑT (r−ˆx(ˆtc))|2 − 1 2||r−ˆy(ˆtc)||2 ˆM(ˆtc) 1 × ei ˆξ(ˆtc)·(r−ˆx(ˆtc))−iˆp(ˆtc)·(r−ˆy(ˆtc))ei ˆχe 4||ˆy(ˆtc)−ˆx(ˆtc)||2 − 1 2||r−ˆx(ˆtc)||2 ˆN (ˆtc)e × e −i = ei(ω−ˆτt(ˆtc;y,p))∆ˆt0(x,ξ;y,p)eiˆp(ˆtc)·(ˆy(ˆtc)−ˆx(ˆtc))e (cid:90) ˆξ(ˆtc)−ˆp(ˆtc) ·(2r−ˆx(ˆtc)−ˆy(ˆtc))e ei ˆχei 2 − 1 4||2r−ˆx(ˆtc)−ˆy(ˆtc)||2 ˆM(ˆtc) ˆM(ˆtc)ei(ω−ˆτt(ˆtc;y,p))∆ˆt0(x,ξ;y,p) (3.101) ˆp(ˆtc)− ˆξ(ˆtc) 2 ·(ˆy(ˆtc)−ˆx(ˆtc)) e(i ˆβ)|ω−ˆτt(ˆtc;y,p)−ˆζT (r−ˆy(ˆtc))|2 eiˆγ|ω−ˆτt(ˆtc;x,ξ)− ˆϑT (r−ˆx(ˆtc))|2 dr + (cid:32) (cid:33) 1(cid:112)||p|| . (3.102) The difference of replacing Hessian matrix has been evaluated in Lemma 3.3.4 and its Corollary. Obviously, now the integral about the receiver variable r is well-defined. We denote its result as ˆB(x, ξ, ω; y, p), ei ˆχei ˆB(x, ξ, ω; y, p) = × e(i ˆβ)|ω−ˆτt(ˆtc;y,p)−ˆζT (r−ˆy(ˆtc))|2 2 ˆξ(ˆtc)−ˆp(ˆtc) ·(2r−ˆx(ˆtc)−ˆy(ˆtc))e − 1 4||2r−ˆx(ˆtc)−ˆy(ˆtc)||2 ˆM(ˆtc) eiˆγ|ω−ˆτt(ˆtc;x,ξ)− ˆϑT (r−ˆx(ˆtc))|2 dr. (3.103) (cid:90) (cid:90) Similarly, ei ˜χei ˜B(x, η, ω; y, q) = × e(i ˜β)|ω−˜τt(˜tc;y,q)−˜ζT (s−˜y(˜tc))|2 2 ˜η(˜tc)−˜q(˜tc) ·(2s−˜x(˜tc)−˜y(˜tc))e − 1 4||2s−˜x(˜tc)−˜y(˜tc)||2 ˜M(˜tc) ei˜γ|ω−˜τt(˜tc;x,η)− ˜ϑT (s−˜x(˜tc))|2 ds. (3.104) We can see that the integral about r and s is accounted for in the computation of ˆB and ˜B. 60 3.3.3.5 Integral about Boundary Points r and s In this section, we will evaluate ˆB and ˜B defined in equation (3.103) and (3.104) to show that they are essentially Gaussian functions. Continuing from the expression (3.103) of ˆB(x, ξ, ω; y, p), we first simplify the exponent ˆχ(r, ˆtc; y, p, x, ξ), ||ˆy(ˆtc) − ˆx(ˆtc)||2 ˆχ = −1 2 ||ˆy(ˆtc) − ˆx(ˆtc)||2 = −1 2 Re( ˆM )(ˆtc) Re( ˆM )(ˆtc) + 1 2 ||r − ˆy(ˆtc)||2 − 1 Re( ˆM )(ˆtc) 2 (ˆy(ˆtc) − ˆx(ˆtc))T Re( ˆM )(ˆtc)(2r − ˆx(ˆtc) − ˆy(ˆtc)), ||r − ˆx(ˆtc)||2 Re( ˆM )(ˆtc) 1 2 + so that we have −i 1 2||ˆy(ˆtc)−ˆx(ˆtc)||2 Re( ˆM )(ˆtc) − Re( ˆM )(ˆtc) (ˆy(ˆtc)−ˆx(ˆtc)) 2 (cid:19) ·(2r−ˆx(ˆtc)−ˆy(ˆtc)) ˆB(x, ξ, ω; y, p) = e (cid:18) ˆp(ˆtc)− ˆξ(ˆtc) (cid:90) −i 2 e − 1 4||2r−ˆx(ˆtc)−ˆy(ˆtc)||2 × e ˆM(ˆtc) ˆF(r, x, ξ, ω; y, p)dr, (3.105) (3.106) where ˆF(r, x, ξ, ω; y, p) = e−Im( ˆβ)|ω−ˆτt(ˆtc;y,p)−ˆζT (r−ˆy(ˆtc))|2 e−Im(ˆγ)|ω−ˆτt(ˆtc;x,ξ)− ˆϑT (r−ˆx(ˆtc))|2 eiˆg(r,x,ξ,ω;y,p), (3.107) 61 and ˆg(r, x, ξ, ω; y, p) = −Re( ˆβ)|ω − ˆτt(ˆtc; y, p) − ˆζT (r − ˆy(ˆtc))|2 + Re(ˆγ)|ω − ˆτt(ˆtc; x, ξ) − ˆϑT (r − ˆx(ˆtc))|2. (3.108) Proposition 3.3.7. The sum of first two terms in the exponent of ˆF in (3.107) satisfy, − Im( ˆβ)|ω − ˆτt(ˆtc; y, p) − ˆζT (r − ˆy(ˆtc))|2 − Im(ˆγ)|ω − ˆτt(ˆtc; x, ξ) − ˆϑT (r − ˆx(ˆtc))|2 = −||r − ˆx(ˆtc) + ˆy(ˆtc) 2Im( ˆβ)ˆζ ˆζT − Im( ˆβ) ||2 ||ˆy(ˆtc) − ˆx(ˆtc)||2 ˆζ ˆζT 2 2 − Im( ˆβ)|ω − ˆτt(ˆtc; y, p)|2 − Im(ˆγ)|ω − ˆτt(ˆtc; x, ξ)|2 + O , (3.109) (cid:32) (cid:33) 1(cid:112)||p|| where ˆβ, ˆγ, ˆζ and ˆϑ are all defined at ˆtc. Similarly, ˆg in equation (3.108) is an O(1) term. Proof. See Appendix A. Lemma 3.3.7. The integral ˆB(x, ξ, ω; y, p) defined in (3.106) has a Gaussian profile centered at ˆy(ˆtc) and ˆp(ˆtc), ˆB(x, ξ, ω; y, p) = e − i 2||ˆy(ˆtc)−ˆx(ˆtc)||2 Re( ˆM )(ˆtc)−iIm( ˆβ)ˆζ ˆζT (cid:118)(cid:117)(cid:117)(cid:116) e−Im( ˆβ)|ω−ˆτt(ˆtc;y,p)|2−Im(ˆγ)|ω−ˆτt(ˆtc;x,ξ)|2 −||(cid:16) (cid:17)||2 ˆp(ˆtc)− ˆξ(ˆtc)−Re( ˆM )(ˆtc)(ˆy(ˆtc)−ˆx(ˆtc)) (2π)d−1 det( ˆM(ˆtc) + Im( ˆβ) 2 Im( ˆβ)ˆζ ˆζT )−1 2 ( ˆM(ˆtc)+ 1 ˆζ ˆζT ) eiO(1), e where we ignore the real-valued phase term ˆg. 62 Proof. Using Proposition 3.3.7, the integral ˆB in (3.106) becomes the Fourier transform of Gaussian functions, so that we have ˆB(x, ξ, ω; y, p) = e −i 1 2||ˆy(ˆtc)−ˆx(ˆtc)||2 Re( ˆM )(ˆtc)−iIm( ˆβ)ˆζ ˆζT (cid:118)(cid:117)(cid:117)(cid:116) (2π)d−1 det( ˆM(ˆtc) + Im( ˆβ) 2 Im( ˆβ)ˆζ ˆζT )−1 2 . ( ˆM(ˆtc)+ 1 ˆζ ˆζT ) (3.110) e−Im( ˆβ)|ω−ˆτt(ˆtc;y,p)|2−Im(ˆγ)|ω−ˆτt(ˆtc;x,ξ)|2 −||(cid:16) (cid:17)||2 ˆp(ˆtc)− ˆξ(ˆtc)−Re( ˆM )(ˆtc)(ˆy(ˆtc)−ˆx(ˆtc)) e Corollary 3.3.4. Similarly, on the source side, ˜B(x, η, ω; y, q) = e −i 1 2||˜y(˜tc)−˜x(˜tc)||2 Re( ˜M )(˜tc)−iIm( ˜β)˜ζ ˜ζT e−Im( ˜β)|ω−˜τt(˜tc;y,q)|2−Im(˜γ)|ω−˜τt(˜tc;x,η)|2 (cid:17)||2 −||(cid:16) ˜q(˜tc)−˜η(˜tc)−Re( ˜M )(˜tc)(˜y(˜tc)−˜x(˜tc)) e (cid:118)(cid:117)(cid:117)(cid:116) (2π)d−1 det( ˜M)(˜tc) + Im( ˜β) 2 ˜ζ ˜ζT ) ( ˜M(˜tc)+ 1 2 Im( ˜β)˜ζ ˜ζT )−1 eiO(1). (3.111) 3.3.3.6 Conclusion of Two Beams’ Interaction To summarize, (cid:90) (cid:90) (cid:90) dξe dξ ˆU x ξ (r, ω) ≈ − 1 4||ˆy(ˆtc)−ˆx(ˆtc)||2 2||ˆy(ˆtc)−ˆx(ˆtc)||2 i 1 ˆM(ˆtc)e Re( ˆM )(ˆtc) ¯ˆU y p (r, ω) dr ei(ω−ˆτt(ˆtc;y,p))∆ˆt0(x,ξ;y,p)eiˆp(ˆtc)·(ˆy(ˆtc)−ˆx(ˆtc))ei ˆξ(ˆtc)−ˆp(ˆtc) 2 ·(ˆy(ˆtc)−ˆx(ˆtc)) ˆB(x, ξ, ω; y, p). (3.112) 63 By Lemma 3.3.5 and Lemma 3.3.6, equation (3.112) reduces to (cid:90)(cid:90) drdξ ¯ˆU y p (r, ω) ˆU x ξ (r, ω) = eip·(y−x) ˆH(x, ω; y, p), (3.113) where ˆH(x, ω; y, p) = (cid:90) 2 ( ˆξ(ˆtc)−ˆp(ˆtc))·(ˆy(ˆtc)−ˆx(ˆtc)) ˆB(x, ξ, ω; y, p)ei(ω−ˆτt(ˆtc;y,p))∆ˆt0(x,ξ;y,p), ei 1 dξei( ˆψ1(ˆtc)− ˆψ1(0))ei( ˆψ2(ˆtc)− ˆψ2(0))e − 1 4||ˆy(ˆtc)−ˆx(ˆtc)||2 ˆM(ˆtc) (3.114) and ˆψ1(t; x, ξ, y, p) = ˆp(t) · (ˆy(t) − ˆx(t)); ˆψ1(0; x, ξ, y, p) = p · (y − x); ˆψ2(t; x, ξ, y, p) = ||ˆy(t) − ˆx(t)||2 Re( ˆM )(t) 1 2 ; ˆψ2(0; x, ξ, y, p) = ||y − x||2 Re( ˆM )(0) . 1 2 (3.115) The extra term e Re( ˆM )(0) = 1, since ˆM (0) is a pure imaginary matrix in −i 1 4||y−x||2 Gaussian wavepacket transform. We denote ˆL(x, ξ, ˆtc; y, p) = ˆψi(ˆtc) − ˆψi(0) − 1 2 (ˆp(ˆtc) − ˆξ(ˆtc)) · (ˆy(ˆtc) − ˆx(ˆtc)). (3.116) (cid:88) i=1,2 (cid:90)(cid:90) Lemma 3.3.5 and Lemma 3.3.6 guarantees ˆL is O(1). Lemma 3.3.8. The receiver-side beam interaction reduces to, drdξ ¯ˆU y p (r, ω) ˆU x ξ (r, ω) ≈ eip·(y−x) ˆH(x, ξ, ω; y, p), 64 where ˆH(x, ξ, ω; y, p) = (cid:90) ˆM(ˆtc)e × ˆB(x, ξ, ω; y, p)ei(ω−ˆτt(ˆtc;y,p))∆ˆt0(x,ξ;y,p). − 1 4||ˆy(ˆtc)−ˆx(ˆtc)||2 dξei ˆL(x,ξ,ˆtc;y,p)e −i 1 2||y−x||2 Re( ˆM )(0) Corollary 3.3.5. The source-side beam interaction reduces to, (cid:90)(cid:90) dsdη ¯˜U y q (s, ω) ˜U x η (s, ω) ≈ eiq·(y−x) ˜H(x, η, ω; y, q), where ˜H(x, η, ω; y, q) = (cid:90) ˜M(˜tc)e × ˜B(x, η, ω; y, q)ei(ω−˜τt(˜tc;y,q))∆˜t0(x,η;y,q), − 1 4||˜y(˜tc)−˜x(˜tc)||2 dηei ˜L(x,η,˜tc;y,q)e −i 1 2||y−x||2 Re( ˜M )(0) where ˜L is defined accordingly. 65 3.3.4 Asymptotic Analysis of Four Beams’ Interaction Using Lemma 3.3.8 and Corollary 3.3.5, the left-hand side of equation (3.47) now becomes, p (r, ω) ¯˜U y ξ (r, ω) ˜U x η (s, ω) dωdrdsdξdηω2 ¯ˆU y (cid:90)(cid:90)(cid:90)(cid:90) − q (s, ω) ˆU x ≈ −ei(p+q)(y−x) ˆH(x, ξ, ω; y, p) ˜H(x, η, ω; y, q) = −ei(p+q)(y−x) (cid:90)(cid:90) (cid:90) dξdηei ˆL(x,ξ,ˆtc;y,p)+i ˜L(x,η,˜tc;y,q)e ω2 ˆB ˜Bei(ω−ˆτt(ˆtc;y,p))∆ˆt0(x,ξ;y,p)ei(ω−˜τt(ˆtc;y,q))∆˜t0(x,η;y,q)dω. 4||ˆy(ˆtc)−ˆx(ˆtc)||2 − 1 ˆM(ˆtc)e 4||˜y(˜tc)−˜x(˜tc)||2 − 1 ˜M(˜tc) (3.117) We will discuss the interaction between four beams in this subsection. 3.3.4.1 Integral about Wavenumber ω We will evaluate the first layer of integral (3.117) about frequency ω in this subsection. Before that, we define a function K(p, q, y), (cid:32) Im( ˆβ)(cid:0)ˆτt(ˆtc; y, p)(cid:1) + Im( ˜β)(cid:0)˜τt(˜tc; y, q)(cid:1) (cid:33)2 K(p, q, y) ≡ − Im( ˆβ) + Im( ˜β) Im( ˆβ)Im( ˜β)(ˆτt(ˆtc;y,p)−˜τt(˜tc;y,q))2 Im( ˆβ)+Im( ˜β) −2 e , (3.118) 66 e−Im( ˆβ)|ˆτt(ˆtc;y,p)−ˆτt(ˆtc;x,ξ)|2 − Im( ˆβ) e 2 ||ˆy(ˆtc)−ˆx(ˆtc)||2 ˆζ ˆζT (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)2(cid:16) ˆM(ˆtc)+2Im( ˆβ)ˆζ ˆζT(cid:17)−1 , (3.119) e−Im( ˜β)|˜τt(˜tc;y,q)−˜τt(˜tc;x,η)|2 − Im( ˜β) 2 e ||˜y(˜tc)−˜x(˜tc)||2 ˜ζ ˜ζT (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)2(cid:16) ˜M(˜tc)+2Im( ˜β)˜ζ ˜ζT(cid:17)−1 . (3.120) a function ˆB(x, ξ; y, p, q) on the receiver side, ˆB = e −|∆ˆt0(x,ξ;y,p)|2 4Im( ˆβ+ ˜β) (2π)d−1 (cid:118)(cid:117)(cid:117)(cid:116) −(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)ˆp(ˆtc)− ˆξ(ˆtc)−Re( ˆM )(ˆtc)(ˆy(ˆtc)−ˆx(ˆtc)) det( ˆM(ˆtc) + Im( ˆβ) ˆζ ˆζT ) 2 × and a function ˜B(x, η; y, p, q) on the source side, ˜B = e −|∆˜t0(x,η;y,q)|2 4Im( ˆβ+ ˜β) (2π)d−1 (cid:118)(cid:117)(cid:117)(cid:116) −(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)˜q(˜tc)−˜η(˜tc)−Re( ˜M )(˜tc)(˜y(˜tc)−˜x(˜tc)) det( ˜M(˜tc) + Im( ˜β) ˜ζ ˜ζT ) 2 × e e (cid:90) − Lemma 3.3.9. The result after taking the integral about ω can be approximated, ei(ω−ˆτt(ˆtc;y,p))∆ˆt0(x,ξ;y,p)ei(ω−˜τt(ˆtc;y,q))∆˜t0(x,η;y,q)ω2 ˆB ˜Bdω − i 2||ˆy(ˆtc)−ˆx(ˆtc)||2 = eiO(1)K(p, q, y) ˆB(x, ξ; y, p, q) ˜B(x, η; y, p, q)e 2||˜y(˜tc)−˜x(˜tc)||2 − i Re( ˜M (˜tc)), e Re( ˆM (ˆtc)) (3.121) where both ˆB and ˜B have phase functions with pure imaginary part only. Proof. See Appendix A. 67 3.3.4.2 Integral about Momentum ξ and η: Evaluation of Real Part of Phase We can define the following integral directly from equation (3.117), (cid:90) ei ˆL(x,ξ,ˆtc;y,p) ˆH(x; y, p, q) = −i 1 e 2||ˆy(ˆtc)−ˆx(ˆtc)||2 Re( ˆM )(ˆtc) ˆB(x, ξ; y, p, q)e − 1 4||ˆy(ˆtc)−ˆx(ˆtc)||2 ˆM(ˆtc)dξ, where ˆL(x, ξ, ˆtc; y, p) = (cid:88) i=1,2 and recall that ˆψi(ˆtc) − ˆψi(0) − 1 2 (ˆp(ˆtc) − ˆξ(ˆtc)) · (ˆy(ˆtc) − ˆx(ˆtc)), ˆψ1(t; x, ξ, y, p) = ˆp(t) · (ˆy(t) − ˆx(t)); ˆψ1(0; x, ξ, y, p) = p · (y − x); ˆψ2(t; x, ξ, y, p) = ||ˆy(t) − ˆx(t)||2 Re( ˆM )(t) 1 2 ; ˆψ2(0; x, ξ, y, p) = ||y − x||2 Re( ˆM )(0) . 1 2 (3.122) Define functions ˆφ1 and ˆφ2 as derivatives of equation (3.122), ˆφ1(t; x, ξ, y, p) = d ˆψ1(t; x, ξ, y, p) dt , ˆφ2(t; x, ξ, y, p) = d ˆψ2(t; x, ξ, y, p) dt . (3.123) In this subsection, we will explore the real part of the phase function in ˆH, i.e., ˆL − 2||ˆy(ˆtc) − ˆx(ˆtc)||2 1 . Re( ˆM ) 68 Proposition 3.3.8. The function ˆφ1 satisfies ˆφ1(t; x, ξ, y, p) = ||ˆp(t) − ˆξ(t)||2 Gpp + O 1 2 (cid:32) (cid:33) . 1(cid:112)||p|| Proof. See Appendix A. Define a 2d by 2d real-valued matrix ˆR,  0, − 1 ˆR(t; y, p) =  . − 1 4 I 4I, tGpp(ˆy(t), ˆp(t)) Proposition 3.3.9. ˆH(x, y, p, q) satisfies (cid:90) ˆH(x, y, p, q) = eiO(1) Proof. First, dξ ˆB(x, ξ; y, p, q)e − 1 4||ˆy(ˆtc)−ˆx(ˆtc)||2 ˆMe i||(ˆy−ˆx,ˆp− ˆξ)( (3.124) (3.125) 2 )||2 ˆtc ˆR (cid:19) (cid:18) ˆtc 2 . ˆL − 1 2 ||ˆy(ˆtc) − ˆx(ˆtc)||2 Re( ˆM )(ˆtc) = ˆψ1(ˆtc) − ˆψ1(0) − ˆψ2(0) − 1 2 = ˆψ1(ˆtc) − ˆψ1(0) − 1 2 (ˆp(ˆtc) − ˆξ(ˆtc)) · (ˆy(ˆtc) − ˆx(ˆtc)) (ˆp(ˆtc) − ˆξ(ˆtc)) · (ˆy(ˆtc) − ˆx(ˆtc)), since Re( ˆM )(0) = 0. We approximate ˆL as, ˆψ1(ˆtc) − ˆψ1(0) = (cid:18) ˆtc (cid:19) 2 ˆtc. ˆφ1(t)dt ≈ ˆφ1 (cid:90) ˆtc 0 69 (3.126) (3.127) (cid:18) ˆp (cid:18) ˆtc 2 (cid:19) (cid:18) ˆtc (cid:19)(cid:19) 2 (cid:18) · ˆy (cid:18) ˆtc (cid:19) 2 − ˆξ (cid:18) ˆtc (cid:19)(cid:19) 2 − ˆx + O(1). Moreover, by using Lemma 3.3.5 twice, (ˆp(ˆtc) − ˆξ(ˆtc)) · (ˆy(ˆtc) − ˆx(ˆtc)) = (3.128) To summarize, ˆL − 1 2 ˆtc 2 ˆφ1 2 = ˆtc 2 ||ˆp ||ˆy(ˆtc) − ˆx(ˆtc)||2 (cid:18) ˆtc (cid:19) (cid:18) ˆtc 2 2 ≈ (cid:19) (cid:18) ˆtc (cid:18) (cid:18) ˆtc (cid:19) Re( ˆM )(ˆtc) − ˆξ (cid:18) ˆtc ||2 Gpp − 1 ˆp 2 2 2 2 (cid:19) − 1 2 − ˆξ = ||(ˆy − ˆx, ˆp − ˆξ)||2 ˆR( ˆtc 2 ) (cid:19)(cid:19) (cid:18) (cid:18) ˆtc (cid:18) · (cid:18) ˆtc (cid:19) ˆy (cid:19) (cid:18) ˆtc 2 − ˆξ ˆp (cid:18) ˆtc (cid:19)(cid:19) 2 · − ˆx 2 2 (cid:19)(cid:19) (cid:18) (cid:18) ˆtc ˆy 2 (cid:19) − ˆx (cid:18) ˆtc (cid:19)(cid:19) 2 . (3.129) Corollary 3.3.6. (cid:90) ˜H(x, y, p, q) ≈ dη ˜B(x, η; y, p, q)e − 1 4||˜y(˜tc)−˜x(˜tc)||2 ˜M exp (cid:32) i||(˜y − ˜x, ˜q − ˜η)( (cid:33) . ˜tc 2 )||2 ˜R( ˜tc 2 ) All terms are defined accordingly. Now the right hand side of equation (3.117) is equal to ei(p+q)(y−x) ˆH ˜H. Therefore, we will focus on ˆH and ˜H next. 70 3.3.4.3 Integral about Momentum ξ and η: Evaluation of Imaginary Part of Phase After seeing the real part of the phase function in Proposition 3.3.9 is a complete quadratic term, we will explore more about the imaginary part in this section. Similar to the real part, we will prove the imaginary part is a complete quadratic term as well as a non-degenerate quadratic term. We start with rewriting ˆB(x, ξ; y, p, q) in (3.119), (cid:118)(cid:117)(cid:117)(cid:116) ˆB(x, ξ; y, p, q) = (2π)d−1 det( ˆM(ˆtc) + Im( ˆβ) 2 −||(ˆy(ˆtc)−ˆx(ˆtc),ˆp(ˆtc)− ˆξ(ˆtc))||2 e ˆI(ˆtc;y,p) ˆζ ˆζT ) −|∆ˆt0(x,ξ;y,p)|2 4Im( ˆβ+ ˜β) e e−Im( ˆβ)|ˆτt(ˆtc;y,p)−ˆτt(ˆtc;x,ξ)|2 . where ˆI(ˆtc; y, p) is a symmetric matrix depending on the fixed beam’s parameters (y, p). −Re( ˆM )  ( ˆM + 2Im( ˆβ)ˆζ ˆζT )−1 (cid:20) I ˆI(ˆtc; y, p) = (cid:21)  Im( ˆβ) 2 ˆζ ˆζT , 0 0, 0  . −Re( ˆM ) I + (3.130) Using Proposition 3.3.9, we have 71 Lemma 3.3.10. ˆH(x; y, p, q) = (cid:118)(cid:117)(cid:117)(cid:116) (2π)d−1 det( ˆM(ˆtc) + Im( ˆβ) 2 ˆζ ˆζT ) (cid:90) − 1 4||ˆy(ˆtc)−ˆx(ˆtc)||2 ˆM(ˆtc) dξe (cid:21) (cid:21) 2 ) − ˆx( ˆtc 2 ) − ˆξ( ˆtc ˆy( ˆy(ˆtc) − ˆx(ˆtc) ˆp(ˆtc) − ˆξ(ˆtc) ˆp( ˆtc 2 ) ˆtc 2 )     . ˆy(ˆtc) − ˆx(ˆtc), ˆp(ˆtc) − ˆξ(ˆtc) ˆI (3.131) × e−Im( ˆβ)|ˆτt(ˆtc;y,p)−ˆτt(ˆtc;x,ξ)|2 −|∆ˆt0(x,ξ;y,p)|2 4Im( ˆβ+ ˜β) e 2 ) − ˆx( ˆtc ˆtc 2 ), ˆp( 2 ) − ˆξ( ˆtc ˆtc 2 ) ˆy( ˆR i (cid:20) − (cid:20) × exp × exp Before proving ˆH has a non-degenerate Gaussian profile, we need one extra assumption. Transformation U between Two Phase Spaces: Suppose there is a transfor- mation U between two phase spaces governed by the certain Hamiltonian flow. The ξ is initially in the phase space P1 = {(x, ξ), x ∈ Rd, ξ ∈ bi-characteristic of the beam ˆU x Rd}, then after propagating to the surface, the bi-characteristic is in the phase space P2 = {((t, x∗), (ω, ξ∗)), t, ω ∈ R, x∗, ξ∗ ∈ Rd−1}. (cid:16) (cid:17)(cid:17) (cid:17) (cid:16) (ˆt0(x, ξ), ˆx∗(ˆt0(x, ξ))), ˆx(ˆt0(x, ξ)), ˆξ(ˆt0(x, ξ)) , ˆξ∗(ˆt0(x, ξ)) , (cid:16)−G U(x, ξ) = (3.132) where ˆt0(x, ξ) is the hitting time defined before, and ˆx(ˆt0(x, ξ)) = (ˆx∗(ˆt0(x, ξ)), 0) is the corresponding hitting point on the boundary. G(x, ξ) is the associated Hamiltonian for the central ray and ˆξ∗ is the component of the ray direction corresponding to the tangential direction of the surface {x ∈ Rd, xd = 0}, 72 that is ˆξ∗(ˆt0(x, ξ)) = ( ˆξ1(ˆt0(x, ξ)),··· , ˆξd−1(ˆt0(x, ξ))). (3.133) The component ˆξd corresponding to the normal direction of the surface can be uniquely defined by (−G, ˆξ∗) according to eikonal equation (2.6), so the degree of freedom won’t change. We need an extra assumption about the bi-characteristic, Assumption 3.3.1. U is invertible. The Gaussian profile is only related to the imaginary part of the phase function, there- fore, we first ignore the term associated with ˆR in (3.131). Lemma 3.3.11. There exists a full-rank 2d by 2d S.P.D. matrix ˆE, such that − 1 4 ||ˆy(ˆtc) − ˆx(ˆtc)||2 ˆM(ˆtc) (cid:20) − − Im( ˆβ)|ˆτt(ˆtc; y, p) − ˆτt(ˆtc; x, ξ)|2 − |∆ˆt0(x, ξ; y, p)|2 (cid:21)  = −||(y − x, p − ξ)||2 4Im( ˆβ + ˜β) ˆy(ˆtc) − ˆx(ˆtc) ˆp(ˆtc) − ˆξ(ˆtc) ˆI ˆE . (3.134) ˆy(ˆtc) − ˆx(ˆtc), ˆp(ˆtc) − ˆξ(ˆtc) Proof. There are three steps to justify the non-degenerate Gaussian profile. First, we will use some approximations to move the left-hand side of equation (3.134) to the phase space P2, since ˆx(ˆtc) is not on the boundary. Second, we will prove that equation (3.134) in the phase space P2 is non-degenerate. Finally, we will use the transformation defined in Assumption 3.3.1 to move the left-hand side term from the phase space P2 to P1. 73 Recall that ˆtc = ˆt0(y, p), − 1 4 ||ˆy(ˆtc) − ˆx(ˆtc)||2 ˆM(ˆtc) (cid:20) − ˆy(ˆtc) − ˆx(ˆtc), ˆp(ˆtc) − ˆξ(ˆtc) − Im( ˆβ)|ˆτt(ˆtc; y, p) − ˆτt(ˆtc; x, ξ)|2 − |∆ˆt0(x, ξ; y, p)|2 (cid:21) ˆy(ˆtc) − ˆx(ˆtc)  4Im( ˆβ + ˜β) ˆI ˆp(ˆtc) − ˆξ(ˆtc) = −1 4 ||ˆy(ˆt0(y, p)) − ˆx(ˆt0(x, ξ))||2 ˆM(ˆt0(y,p)) − |∆ˆt0(x, ξ; y, p)|2 4Im( ˆβ + ˜β) − ˆy(ˆtc) − ˆx(ˆtc), ˆp(ˆtc) − ˆξ(ˆtc) (cid:20) (cid:21) − Im( ˆβ)|ˆτt(ˆt0(y, p); y, p) − ˆτt(ˆt0(x, ξ); x, ξ)|2 (cid:33) ˆy(ˆtc) − ˆx(ˆtc)  + O 1(cid:112)||p|| (cid:32) ˆI ˆp(ˆtc) − ˆξ(ˆtc) = −1 4 ||ˆy(ˆt0(y, p)) − ˆx(ˆt0(x, ξ))||2 ˆM(ˆt0(y,p)) − |∆ˆt0(x, ξ; y, p)|2 4Im( ˆβ + ˜β) − Im( ˆβ)|ˆτt(ˆt0(y, p); y, p) − ˆτt(ˆt0(x, ξ); x, ξ)|2 −(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:20) (3.135) (cid:32) (cid:33) , 1(cid:112)||p|| (3.136) (cid:21)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)2 ˆy(ˆt0(y, p)) − ˆx(ˆt0(x, ξ)) ± ∆ˆt0v ˆξ || ˆξ||, ˆp(ˆt0(y, p)) − ˆξ(ˆt0(x, ξ)) ˆI + O where v in equation (3.136) is defined at ˆy(ˆt0(y, p)). The first step (3.135) is due to the following derivation. Using Corollary 3.3.3, ||ˆy(ˆtc) − ˆx(ˆtc)||2 ˆM(ˆtc) = ||ˆy(ˆtc) − ˆx(ˆtc)||2 ˆN (ˆtc) + O (cid:32) (cid:33) . 1(cid:112)|p|| Using Proposition 3.3.4, ||ˆy(ˆtc) − ˆx(ˆtc)||2 ˆN (ˆtc) = ||ˆy(ˆt0(y, p)) − ˆx(ˆt0(x, ξ))||2 ˆN (ˆtc) + O (cid:32) (cid:33) , 1(cid:112)|p|| 74 since ˆy(ˆt0(y, p)) is on the boundary. Use Corollary 3.3.3 again, ||ˆy(ˆt0(y, p)) − ˆx(ˆt0(x, ξ))||2 ˆN (ˆt0(y,p)) ||ˆy(ˆt0(y, p)) − ˆx(ˆt0(x, ξ))||2 ˆM(ˆt0(y,p)) (cid:32) = + O (cid:33) . 1(cid:112)|p|| The second step (3.136) is due to Assumption 3.2.3 as the ray direction ˆξ remains as a constant near the boundary. Now we will prove equation (3.136) is non-degenerate in P2. If ˆU y p and ˆU x ξ have different hitting points ˆy(ˆtc) (cid:54)= ˆx(ˆt0(x, ξ)), then equation (3.136) is obviously negative. This is also true for the nonzero difference of travel time ∆ˆt0 and ˆτt(ˆt0(y, p); y, p) − ˆτt(ˆt0(x, ξ); x, ξ). If we have all difference mentioned above is zero and ˆp− ˆξ is nonzero along the tangential directions of the boundary, then equation (3.136) becomes, ||(0, ˆp(ˆt0(y, p)) − ˆξ(ˆt0(x, ξ)))||2 ˆI = ||ˆp(ˆtc) − ˆξ(ˆtc)||2 ( ˆM+2Im( ˆβ)ˆζ ˆζT )−1 > 0, (3.137) since ˆM and its pseudo-inverse ˆM−1 are S.P.D. if restricted to the tangential directions by Lemma A.0.1. Finally, due to the existence of ˆU−1, we can define a non-degenerate Gaussian profile about (x, ξ) accordingly, that is ˆE = ˆU−T ˆC ˆU−1, (3.138) 75 where ˆC makes the following term (cid:20) || (∆ˆt0, ˆy(ˆtc) − ˆx(ˆt0(x, ξ))), (cid:21) (ˆτt(ˆtc; y, p) − ˆτt(ˆtc; y, p), ˆp(ˆtc) − ˆξ(ˆt0(x, ξ))) ||2 ˆC equal to the equation (3.136). Remark 3.3.1. Although the real part of the phase function in equation (3.131) is ignored in this part of the computation, it will not affect the above computation essen- tially, especially the existence of the Gaussian profile. 3.3.4.4 Integral about Momentum ξ and η After justifying the existence of a non-degenerate Gaussian profile centered at (y, p), the next question is how to calculate ˆE in Lemma 3.3.11 numerically. We first define a matrix ˆA(ˆtc; y, p, q) depending only on fixed beam’s parameters (y, p)  1 4 ˆM(ˆtc) + ||ˆp(ˆtc)||2edeT 4(cid:0)v(ˆy(ˆtc))ˆpd(ˆtc)(cid:1)2Im( ˆβ+ ˜β) d (cid:19)(cid:18) (cid:18) , 0 (cid:19)T 0, Im( ˆβ)v2(ˆy(ˆtc)) ˆp(ˆtc) ||ˆp(ˆtc)|| ˆp(ˆtc) ||ˆp(ˆtc)||  , (3.139) ˆA(ˆtc; y, p, q) = ˆI(ˆtc; y, p) + where ed = (0,··· , 0, 1) ∈ Rd. Lemma 3.3.12. There exists a S.P.D. 2d by 2d matrix ˆA, such that ||(y − x, p − ξ)||2 ˆE = ||(ˆy(ˆtc) − ˆx(ˆtc), ˆp(ˆtc) − ˆξ(ˆtc))||2 ˆA, (3.140) and ˆA is defined in equation (3.139). Proof. Firstly, we would like to approximate ∆ˆt0(x, ξ; y, p) and ˆτt(ˆtc; y, p)− ˆτt(ˆtc; x, ξ). 76 By the definition of hitting time and Assumption 3.2.3, 0 = ˆyd(ˆtc) − ˆxd(ˆt0(x, ξ)) = ˆyd(ˆtc) − ˆxd(ˆtc) ± v(ˆy(ˆtc)) ˆpd(ˆtc) ||ˆp(ˆtc)||∆ˆt(x, ξ; y, p). (3.141) Here ± is determined by the sign of the Hamiltonian. To summarize, −|∆ˆt0(x, ξ; y, p)|2 4Im( ˆβ + ˜β) 4(cid:0)v(ˆy(ˆtc))ˆpd(ˆtc)(cid:1)2 = − ||ˆp(ˆtc)||2|eT d (ˆy(ˆtc) − ˆx(ˆtc))|2 Im( ˆβ + ˜β) . (3.142) Similarly, −Im( ˆβ)|ˆτt(ˆtc; y, p) − ˆτt(ˆtc; x, ξ)|2 = −Im( ˆβ)v2(ˆy(ˆtc)) (cid:12)(cid:12)(cid:12)||ˆp(ˆtc)|| − || ˆξ(ˆtc)||(cid:12)(cid:12)(cid:12)2 . (3.143) Furthermore, it can be approximated by, −Im( ˆβ)|ˆτt(ˆtc; y, p) − ˆτt(ˆtc; x, ξ)|2 ≈ −Im( ˆβ)v2(ˆy(ˆtc)) . (3.144) (cid:12)(cid:12)(cid:12) ˆp(ˆtc) (cid:12)(cid:12)(cid:12)2 ||ˆp(ˆtc)|| · (ˆp − ˆξ) The Lemma is proved. The integrand (3.131) now is a quadratic term about (ˆy(ˆtc) − ˆx(ˆtc), ˆp(ˆtc) − ˆξ(ˆtc)). Furthermore, we have the following proposition. Proposition 3.3.10. There exists a linear map ˆJ (ˆtc; y, p), such that ˆy(ˆtc) − ˆx(ˆtc) ˆp(ˆtc) − ˆξ(ˆtc)  ≈ ˆJ (ˆtc; y, p) y − x p − ξ  . (3.145) 77 Proof. See Appendix A. Similarly, the matrix ˜A(˜tc; y, p, q) and the map ˜J (˜tc; y, q) on the source side are defined accordingly. Now ˆH becomes, ˆH(x; y, p, q) = (cid:90) dξe (cid:118)(cid:117)(cid:117)(cid:116) (2π)d−1 det( ˆM(ˆtc) + Im( ˆβ) 2 ˆζ ˆζT ) i|| ˆJ ( −|| ˆJ (ˆtc;y,p)(y−x,ξ−p)||2 ˆA(ˆtc;y,p,q)e 2 ;y,p)(y−x,ξ−p)||2 ˆtc ˆR(ˆtc;y,p). (3.146) Therefore, ˆE under Assumption 3.2.3 is ˆE ≈ ˆJ (ˆtc; y, p)T ˆA(ˆtc; y, p, q) ˆJ (ˆtc; y, p). (3.147) The above equation provides an efficient way to approximate ˆE. Now let’s compute the integral about ξ. We suppose the S.P.D. matrix ˆE,  ˆE11, ˆE T 12,  . ˆE12 ˆE22 ˆE = (3.148) 78 Now we would like to show that there’s a Gaussian profile about x centered at y, −||(y−x,p−ξ)||2 ˆE dξe −||y−x||2 −||y−x||2 −||y−x||2 ||y−x||2 ˆE12 ˆE−1 22 ˆET 12 ||y−x||2 ˆE12 ˆE−1 22 ˆET 12 ||y−x||2 ˆE12 ˆE−1 22 ˆET 12 ˆE11 e ˆE11 e ˆE11 e (cid:90) (cid:90) (cid:90) −||p−ξ+ ˆE−1 22 ˆET 12(y−x)||2 ˆE22 dξe −||p−ξ||2 ˆE22 dξe −||(0,p−ξ)||2 ˆE . dξe (3.149) (cid:90) = e = e = e (cid:90) ˆE11 − ˆE12 ˆE−1 displays that ˆH contains a Gaussian profile about y − x. ˆE T 12 is S.P.D by the fact that it is the Schur complement of ˆE22 in ˆE. This 22 Together with equation (3.147), we have −||(y−x,p−ξ)||2 ˆE dξe −||y−x||2 ||y−x||2 ˆE12 ˆE−1 22 ˆET 12 ˆE11 e = e (cid:90) −|| ˆJ (ˆtc;y,p)(0,p−ξ)||2 ˆA. dξe Together with equation (3.146), we obtain a way to calculate ˆH(x; y, p, q). The real part of the phase function associated with ˆR will not affect the final result essentially and it can be compensated by a constant order phase term. ˆH(x; y, p, q) = eiO(1) (2π)d−1 (cid:90) det( ˆM(ˆtc) + Im( ˆβ) 2 i|| ˆJ ( i|| ˆJ (ˆtc;y,p)(0,p−ξ)||2 ˆAe dξe i 4||y−x||2 ˆM (0) e ˆζ ˆζT ) 2 ;y,p)(y−x,p−ξ)||2 ˆtc ˆR(ˆtc;y,p). (3.150) (cid:118)(cid:117)(cid:117)(cid:116) Here we use 1 4M0 to approximate the Schur complement of ˆE. 79 Lemma 3.3.13. ˆH can be approximated by the following equation, ˆH(x; y, p, q) ≈ ˆK(y, p, q)e i 4||y−x||2 ˆM (0), (3.151) where (cid:118)(cid:117)(cid:117)(cid:116) (2π)d−1 det( ˆM(ˆtc) + Im( ˆβ) 2 ˆK(y, p, q) = −|| ˆJ (ˆtc;y,p)(0,ξ−p)||2 ˆA(ˆtc;y,p,q)e 2 ;y,p)(0,ξ−p)||2 ˆtc ˆR(ˆtc;y,p). (3.152) dξe ˆζ ˆζT ) i|| ˆJ ( Similarly, ˜H can be approximated by the following equation, (cid:90) (cid:90) Corollary 3.3.7. where ˜H(x; y, p, q) ≈ ˜K(y, p, q)e i 4||y−x||2 ˜M (0) (3.153) (cid:118)(cid:117)(cid:117)(cid:116) (2π)d−1 det( ˜M(˜tc) + Im( ˜β) 2 ˜K(y, p, q) = ˜ζ ˜ζT ) i|| ˜J ( −|| ˜J (˜tc;y,q)(0,η−q)||2 ˜A(˜tc;y,p,q)e 2 ;y,q)(0,η−q)||2 ˜tc ˜R(˜tc;y,q). (3.154) dηe The real part ˆR(ˆtc; y, p) and ˜R(˜tc; y, q) can be compensated by a constant phase term and will not affect the result essentially. Remark 3.3.2. In equation (3.151), we essentially approximate the Schur complement ˆE11 − ˆE12 ˆE−1 22 ˆE T 12 by M0 4 . However, it is costly to use the exact value (3.146) since 80 we have to store all matrices ˆA(ˆtc; y, p, q), ˜A(˜tc; y, p, q) generated by different pairs of {(p, q) : p + q = ξl,i} to apply the inverse Gaussian wavepacket transform (3.49). On the other hand, the difference between equation (3.151) and equation (3.146) will be at constant order guaranteed by Lemma 3.3.6. To summarize, the central direction and central point of the wavepacket will not be affected by approximation (3.151) and the width of wavepacket is at the same scale. The exact Hessian information (3.146) can be covered but it is costly to compute. Theorem 3.3.1. (cid:90) −ω2dω (cid:90) (cid:90) drds ¯ˆU y p (r, ω) ¯˜U y q (s, ω) dξdη ˆU x ξ (r, ω) ˜U x η (s, ω) ≈ K(p, q, y) ˆK(y, p, q) ˜K(y, p, q)ei(p+q)·(y−x)e i||y−x||2 M0/2, (3.155) where K(p, q, y) is defined in equation (3.118), ˆK and ˜K are defined in equation (3.152)-(3.154). The distance between ˆτt(ˆtc; y, p) and ˜τt(˜tc; y, q) is controlled by K(p, q, y). 3.4 Implementation of the Prestack Imaging Oper- ator By equations (3.46), (3.47) and (3.48), we conclude that the partial imaging function Ipq(y, ω) is related to the Gaussian wavepacket transform of 2α v2 centered at y in the 81 direction p + q. (cid:90) Ipq(y, ω)dω = E(p, q, y) (cid:90) dx 2α v2 ei(p+q)·(y−x)e −||y−x||2 M0/2, (3.156) where E(p, q, y) = ˆK(y, p, q) ˜K(y, p, q)K(p, q, y). The numerical scheme to calculate E(p, q, y) is given by equation (3.118) and equations (3.152)-(3.154). Therefore, this section will be devoted to illustrating how to compute the integral of the imaging function Ipq(y, ω) efficiently. (cid:90) (cid:90) (cid:90) dω Ipq(y, ω)dω = drds ¯˜U y q (s, ω) ¯ˆU y p (r, ω)D(r, s, ω). (3.157) We start with the integral about the wavenumber ω. Using Corollary 3.3.1, ˆΦ(r, ˆtc; y, p) = ˆτt(ˆtc; y, p) + ˆζT (r − ˆy(ˆtc)), ˜Φ(s, ˜tc; y, q) = ˜τt(˜tc; y, q) + ˜ζT (s − ˜y(˜tc)). By considering the terms containing ω in ˜U y q (s, ω) and ˆU y p (r, ω) only, (cid:90) eiO(1) dω ¯˜U y (cid:90) ¯ˆU y p (r, ω)D(r, s, ω) = q (s, ω) D(r, s, ω)eiω(ˆtc+˜tc)e−Im( ˆβ)|ω− ˆΦ|2 e−Im( ˜β)|ω− ˜Φ|2 dω. (3.158) Here we neglect the real part of the exponent, Re( ˆβ)|ω − ˆΦ|2 and Re( ˜β)|ω − ˜Φ|2, since they are constant order terms by Proposition 3.3.7. Consequently, they are small compared with the term ω(ˆtc + ˜tc). 82 The integral now can be considered as an inverse Fourier transform about wavenumber ω, (cid:90) e−Im( ˜β)|ω− ˜Φ|2 dω D(r, s, ω)eiω(ˆtc+˜tc)e−Im( ˆβ)|ω− ˆΦ|2 (cid:90) = − Im( ˆβ)Im( ˜β)( ˜Φ− ˆΦ)2 e Im( ˆβ+ ˜β) eiS(ˆtc+˜tc) D(r, s, ω)ei(ω−S)(ˆtc+˜tc)e−Im( ˆβ+ ˜β)(ω−S)2 dω, (3.159) where S(r, s, ˆtc, ˜tc; y, p, q) = Im( ˆβ) ˆΦ(r, ˆtc; y, p) + Im( ˜β) ˜Φ(s, ˜tc; y, q) Im( ˜β + ˆβ) . (3.160) The integral in equation (3.159) is indeed a convolution, − Im( ˆβ)Im( ˜β)( ˜Φ− ˆΦ)2 Im( ˆβ+ ˜β) (cid:90) π eiS(ˆtc+˜tc)e (cid:114) = − Im( ˆβ)Im( ˜β)( ˜Φ− ˆΦ)2 e Im( ˆβ + ˜β) Im( ˆβ+ ˜β) (cid:90) eiS(ˆtc+˜tc) D(r, s, ω)ei(ω−S)(ˆtc+˜tc)e−Im( ˆβ+ ˜β)(ω−S)2 dω − e h2 4Im( ˆβ+ ˜β) eiShD(r, s, ˆtc + ˜tc − h)dh. (3.161) Notice the data D in the above formula is in the time domain, which means there is no need to apply the Fourier transform to the data at prior. Moreover, the convolution (cid:32) (cid:113)|ˆτtt(ˆtc;y,p)|+|ˆτtt(˜tc;y,q)| 1 (cid:33) . integral is conducted in a small range, i.e. O 83 Naturally, use Corollary 3.3.1 and equation (3.157) (cid:114) (cid:90) (cid:90) drdse (cid:115) π ˆA(ˆtc) ˜A(˜tc) Ipq(y, ω)dω = Im( ˆβ + ˜β) −||r−ˆy(ˆtc)||2 ˆM 2 e −||s−˜y(˜tc)||2 ˜M 2 e (cid:90) × 2π 1 ˆτtt(ˆtc; y, p)˜τtt(˜tc; y, q) − Im( ˆβ)Im( ˜β)( ˆΦ− ˜Φ)2 Im( ˆβ+ ˜β) eiS(ˆtc+˜tc)e−i(ˆ(r)+˜(s)) − e h2 4Im( ˆβ+ ˜β) eiShD(r, s, ˆtc + ˜tc − h)dh. (3.162) The integral about r and s is easy due to the existence of Gaussian profiles which will constrain the integral range. Therefore, the regular integration scheme is enough. Here we notice that the value of the imaging function is controlled by | ˜Φ − ˆΦ| ∼ |˜τt(˜tc; y, q)− ˆτt(ˆtc; y, p)|. Therefore, an efficient way is needed to avoid computing pairs of beams with little illumination or imaging function (cid:82) dωIpq(y, ω) (3.162) which is small. Therefore, we should select pairs of beams (y, p) and (y, q) such that | ˆΦ(r, ˆtc) − ˜Φ(s, ˜tc)| ≤ 2(cid:112)||p|| + 2(cid:112)||q||. (3.163) Fortunately, the time derivative of the phase function does not change along the ray, which means we can estimate equation (3.163) without propagating beams. 84 3.5 Numerical Results 3.5.1 Approximation of Beams along the surface The numerical examples in this subsection are provided to justify the approximation in Section 3.3.2. The following tests are done by comparing our approximation results to the results obtained by the numerical integration method. The velocity used here is v = 0.8 + 0.4x2 and the initial beam is initiated at the subsurface point (0, 0.5). The initial momentum is set as 30π[cos(0.1π), sin(0.1π)] and the initial amplitude is 1 + i. Figure 3.4: Numerical test for fast approximation of Gaussian beam along the surface. Left: ω = -110 Right: ω = -100. The blue line represents the result achieved from the numerical integration method and the red star is the one from our fast approximation algorithm. 85 −101−0.0500.050.10.15x1u−101−0.2−0.100.10.20.3x1u 3.5.2 The Correctness of Prestack Imaging Operator We conduct the numerical test to justify Theorem 3.3.1. We first fix a direction e and the point y so that x is acquired by moving along this certain fixed direction, i.e. x = y + ∆he. We compare our proposed result with the numerical integration result after removing the highly oscillated term eip·(y−x). The initial subsurface point is the point y = (0, 0.5) and the initial ray direction is p = (6π,−30π). Two different directions are picked here. The first one is e = (1, 1), while the second is chosen as (1, 2). The x-axis in each plot is ∆h , y-axis in each plot Figure 3.5: Constant Slowness: Theorem 3.3.1. Left: dx1 = dx2 Right: dx1 = 2dx2. is the value of integral. As we can see in the constant slowness, the approximation proposed in Theorem 3.3.1 has only a small amount of error. The second velocity used for test is 1 + 0.1x2 + 0.1x1. Other setups are the same. As we can see in Fig. 3.6, the imaging operator does not perform as well as it does in the 86 −0.0500.053.544.555.5dx1y−0.0500.051.522.53dx1y Figure 3.6: General Speed: Theorem 3.3.1. Left: dx1 = dx2 Right: dx1 = 2dx2. constant slowness. However, the central momentum is captured correctly as stated in Remark 3.3.2. 3.5.3 Single Source Migration Test We will recover the reflector by the single-source data trace in this section. 3.5.3.1 Example 1: Constant Background Slowness Fig. 3.7 is the true slowness we employ, and there is a dipped layer. The source point here is at x = (0, 0). As the Fig. 3.7 shows, the migration result shows the ability of our algorithm to detect the correct location and dipped angle. 87 −0.0500.05012345dx1y−0.0500.05−20246dx1y Figure 3.7: Example 1: Constant Slowness with Dipped layer. Left: True Slowness Right: Migration Result with single source trace. 3.5.3.2 Example 2: Multiple Flat Reflectors In this numerical example, our migration algorithm is tested by two flat reflectors at different depth, The migration result is, The red dashed line is the true value while Figure 3.8: Example 2: True Slowness Model with Multiple Layers the blue line the migration result. The deeper layer is not captured as well as the first layer. The error here is due to the Born approximation assumption. 88 x1x2 −0.500.5−0.4−0.200.20.40.60.80.780.790.80.810.82x1x2 −0.500.5−0.4−0.200.20.40.60.8−0.04−0.0200.020.040.06x1x2 −0.500.5−0.4−0.200.20.40.60.480.490.50.510.52 Figure 3.9: Example 2: Constant Slowness with Multiple Layers. Left: Migration Result over the Whole Space Right: Migration Result V.S. True Value at x1 = 0. 3.5.3.3 Example 3: Linear Background Slowness To see the amplitude information, we plot the slowness at x1 = 0 and compare it with the true value. And the red dashed line is the true value while the blue line is our migration result. 3.5.4 Multiple Source Migration Test We will use multiple-source data trace in this section. 3.5.4.1 Example 4: Constant Slowness with Dipped Layer The background slowness is same as the one in Example 1. Sources are a series of points along the surface − 1 4. This is applied to all multiple-source tests. The 4 : 1 16 : 1 dipped layer is displayed correctly in Fig. 3.12. 89 x1x2 −0.500.5−0.4−0.200.20.40.6−0.04−0.0200.020.040.06−1−0.500.51−0.06−0.04−0.0200.020.040.06x2u Figure 3.10: Example 3: Gradient Slowness Model. Left: True Slowness, Right: Smoothed Macro Slowness Figure 3.11: Example 3: Gradient Slowness Model. Left: Migration Result over the Whole Space, Right: Migration Result V.S. True Value at x1 = 0. 90 x1x2 −0.500.5−0.4−0.200.20.40.40.450.50.55x1x2 −0.500.5−0.4−0.200.20.40.40.450.50.55x1x2 −0.4−0.200.20.40.6−0.4−0.3−0.2−0.100.10.20.30.4−0.0500.05−101−0.0500.05x2u Figure 3.12: Example 4: Constant Slowness with the Dipped Layer (Multiple Sources) 3.5.4.2 Example 5: Flat Layer in Lateral Background Velocity We add some lateral variation to the background slowness, i.e. v = 0.8+0.1 sin(0.5πy) sin(3π(x+ 0.05)). The reflector is a horizontal reflector. Figure 3.13: Example 5: Flat Layer in Lateral Background Velocity 91 x1x2 −0.4−0.200.20.40.6−0.4−0.200.20.40.60.81−0.05−0.04−0.03−0.02−0.0100.010.020.030.040.05x1x2 −1−0.500.5−1−0.500.5−0.06−0.04−0.0200.02 3.5.4.3 Example 6: Slowness with Caustics The next two examples in the section are both using the Gaussian velocity as the macro velocity shown in Fig. 3.14 (a). This is more complex as the caustics will show up. As we can see in Fig. 3.14, there is a caustics around the level x2 = 0.46. Our Figure 3.14: Example 6: Gaussian Slowness and its ray tracing. Left: Gaussian Slowness with Flat Reflector Right: Ray Tracing flat reflector is below this caustics at x2 = 0.6. The multi-value problem caused by caustics is resolved automatically by the Gaussian beam solution in Fig. 3.15. 3.5.4.4 Example 7: Polluted Trace Data In the end, we would like to test our inversion process using the polluted data. We add 5% Gaussian noise into the synthetic data. See Fig. 3.16 for more details. The red dot line is the trace with extra Gaussian error, while the blue line is the original trace. 92 x1x2 −0.4−0.200.20.40.600.10.20.30.40.50.60.70.620.640.660.680.70.720.740.760.780.8−0.4−0.200.200.20.40.60.8x1x2 Figure 3.15: Example 6: Migration Result in the Gaussian Slowness with Caustics (Multiple Sources) Figure 3.16: Example 7: True Trace V.S. Trace with Gaussian Error The migration result is displayed in Fig. 3.17. There is no much difference in resulting images, especially around the reflector. To see more details, we compare two results at x2 = 0.65 in Fig. 3.18, The red dot line comes from the non-polluted data while the blue line is from the polluted boundary data. 93 x1x2 −1−0.500.5−1−0.8−0.6−0.4−0.200.20.40.60.8−0.08−0.06−0.04−0.0200.020.040.0600.511.522.5−0.8−0.6−0.4−0.200.20.40.60.8tu Figure 3.17: Example 7: Gaussian Slowness with polluted trace. Left: Migration result from Non-polluted Data; Right: Migration result from Polluted data Figure 3.18: Example 7: Gaussian Slowness with polluted trace.Two Migration Results at x2 = 0.65 94 x1x2 −1−0.500.5−1−0.500.5−0.08−0.06−0.04−0.0200.020.040.06x1x2 −1−0.500.5−1−0.500.5−0.1−0.0500.05−1−0.500.51−0.0200.020.040.060.080.1x1u Chapter 4 Fast Multiscale Gaussian Beam Method for Elastic Wave Equations in Bounded Domains 4.1 Asymptotic Method for the Elastic Wave equa- tion The problem considered in this paper is the initial-boundary value problem of the elastic wave equation. 0 = ρ¨u − ∇λ(∇ · u) − ∇µ · (∇u + ∇uT ) − (λ + µ)∇(∇ · u) − µ∆u, (4.1) where the parameters λ and µ are known as the Lame parameters. ˙u = ∂u ∂t ; ¨u = ∂2u ∂t2 (4.2) 95 and this notation is applied to all functions. They are assumed to be smooth, positive and bounded away from zero. The initial condition is defined as the following, u(x, 0) = f; ut(x, 0) = g, (4.3) where the functions f and g are compactly supported vector-valued functions in the space L2(Rd). We are looking for the asymptotic solution for the elastic wave equation (4.1) with two different types of boundary conditions, the periodic boundary condition and the homogeneous Dirichlet boundary condition, i.e. (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)∂Ω×[0,T ] u(x, t) = 0. (4.4) 4.2 The Asymptotic Ansatz Solution to the Elastic Wave We firstly derive the eikonal and the transport equation for the elastic wave equation. Same as the Geometrical-optics form, we consider the solution as the following series expansion, u(x, t) = eiωτ (t,x) ∞(cid:88) n=0 A(n)(t, x)(iω)−n, (4.5) where the wavenumber ω is assumed to be a large parameter relative to the elastic moduli λ and µ’s changing rate, i.e. ωL min(λ,µ) >> 1, where L is the characteristic distance defined as the scale over which the velocity changes slowly. The asymptotic solution for equation (4.1) is defined in the sense that both the equation itself and 96 initial-boundary conditions are satisfied approximately with a small error when ω is large. To construct two equations governing the phase function τ and amplitude func- tion A(0) respectively, we substitute the ansatz form A(0)eiωτ into equation (4.1). For convenience, we write A(0) as A. The jth component of the elastic wave equation will then be (cid:18) 1 ω O (cid:19) = eiωτ(cid:8)λ,j(Ak,k + iωAkτ,k) + µ,k[Ak,j + Aj,k + iω(τ,jAk + Ajτ,k)] + (λ + µ)[(cid:0)Ak,k (cid:1) − ω2τ,jAkτ,k] (cid:1) ,j + iω(cid:0)(Akτ,k),j + τ,jAk,k (cid:9), + µ[(Aj,kk) + iω(2Aj,kτ,k + Ajτ,kk) − Ajω2τ 2 ,k] + ρAjω2( ˙τ )2 − 2iωρ ˙τ ˙Aj − iωρAj ¨τ − ρ ¨Aj (4.6) here | · | denotes the Euclidean norm in Rd and  .  ∂ ∂x1 ... ∂ ∂xd ∇ = Other notations used frequently in this paper are · representing the inner product between two column vectors and vT representing the transpose of vector v. Other notations we use in the above equation (4.6) are, τ,k = ∂τ /∂xk. To make our derivation simpler, we let ρ = 1 without losing any generality. 97 4.2.1 P-wave and S-wave’s Eikonal Equations To cancel out the leading term ω2 in equation (4.6), we set its coefficient to be zero, 0 = ˙τ 2Aj −(cid:0)(λ + µ)τ,jAkτ,k + µAjτ,kτ,k (cid:1) . After concatenating j = 1, 2,··· , d as a vector, we will get (λ + µ)∇τ∇τ T A = ( ˙τ 2 − µτ,kτ,k)A, (4.7) (4.8) so the amplitude vector A will be the eigenvector of the matrix ∇τ∇τ T . Notice that this simple matrix is rank one matrix by some linear algebra calculations, and the eigenvector corresponding to the single nonzero eigenvalue must be parallel to the vector ∇τ P , while other two associated with the zero eigenvalues are orthogonal to ∇τ S. The superscript here represents the category of their wave modes. We have ( ˙τ P )2 − (λ + 2µ)(τ P ,k τ P ,k ) = 0. (4.9) This is the eikonal equation for the P-wave whose amplitude vector AP is parallel to the ray direction ∇τ P . Another two eigenvectors are corresponding to the S-wave, ( ˙τ S)2 − µ(τ S ,kτ S ,k) = 0. (4.10) whose the amplitude vector AS is perpendicular to the ray direction ∇τ S. With the eikonal equation at our disposal, we can apply the method of characteristics 98 to the nonlinear eikonal equations (4.9) and (4.10). These two eikonal equations (4.9) and (4.10) are essentially the same, consequently, we consider them in the generic situation as a Hamilton-Jacobi equation. ˙τ P,S + GP,S(x,∇τ ) = 0, (4.11) where the Hamiltonian of the P-wave GP (x, p) = ±√ of the S-wave GS(x, p) = ±√ µ|p|. We consider the P-wave case for the illustration, λ + 2µ|p| and the Hamiltonian  dx dt = GP dt = −GP dp p (x(t), p(t)), x (x(t), p(t)), x(0) = x0; p(0) = p0. (4.12) where t is the running parameter of a bicharacteristic. Solving this ODE system yields a bicharacteristic in the phase space {(x(t), p(t)) : t ≥ 0} and the associated ray γ = {x(t) : t ≥ 0}, which is its x-component. Moreover, it is noticed that we have the equation p(t) = ∇τ (t, x(t)) along the ray γ due to the method of characteristics. One of the most significant difference between the Gaussian beam and other ray-ansatz methods is that beams’ phase functions are complex-valued. To be more specific, its second order derivative is complex-valued. To derive the dynamics of the Hessian matrix, we first differentiate the eikonal equation (4.11) with respect to t and x near 99 the ray γ: ˙τ P x + GP x + τ P xxGP p = 0, ¨τ P + (GP p )T ˙τ P x = 0. Differentiating the first equation above (4.13) with respect to x again yields ˙τ P xx + GP xx + τ P xxGP xp + (GP xp)T τ P xx + τ P xxGP p = 0. ppτ p xx + τ P xxxGP (4.13) (4.14) (4.15) Since the equations (4.14), (4.13) and (4.15) are all valid everywhere in the phase space, it will still be valid if we concentrate them along the ray. Let M P (t) be the Hessian of the phase function along the ray dM P dt + GP xx + M P GP xp + (GP xp)T M P + M P GP ppM P = 0. (4.16) And the same rule can be applied to the S-wave with GS(x, p) = ±√ µ|p|. One in- teresting property of the Gaussian beam solution is that it will remain well-localized throughout the propagation, which means the imaginary part of the Hessian M should always be symmetric positive definite. The following lemma [49] guarantees this prop- erty throughout the propagation for all smooth ray trajectories, Lemma 4.2.1. If the Hamiltonian G is smooth enough, then the Hessian M(t) along the ray path γ has a positive-definite imaginary part, provided that it initially does. Accordingly, the Hessian of beam ansatz’s phase functions is well-defined at all points even the caustics. 100 4.2.2 Transport Equation Governing P-wave’s Amplitude Vec- tors Setting the coefficient of O(ω) term in equation (4.6) equal to zero will yield the transport equation about the amplitude vector A. We first see the jth component, 0 = (2 ˙Aj ˙τ + ¨τ Aj) − λ,j(Akτ,k) − µ,k(τ,jAk + Ajτ,k) − (λ + µ)(cid:0)(Akτ,k),j + Ak,kτ,j (cid:1) − µ(2Aj,kτ,k + Ajτ,kk). (4.17) Although it is complex at first glance, especially compared with the transport equation of the scalar wave equation, the complexity can be reduced by properties of the P-wave and S-wave. We start with deriving the amplitude vector AP for the P-wave, which is parallel to the ray direction ∇τ P . Therefore, the P-wave’s amplitude can be separated as A = a∇τ P . To make derivation more readable, we write τ P as τ in this part. We will yield the jth component by inserting A = a∇τ into equation (4.17) (cid:16) (cid:17) ,k − 2a(µ,kτ,k)τ,j ,k + 2aτ,kτ,kj 0 = (2 ˙aτ,j ˙τ + 2a ˙τ ˙τ,j + a¨τ τ,j) − λ,jaτ 2 − (λ + µ) − µ(cid:0)aτ,jτ,kk + 2τ,ja,kτ,k + 2aτ,kτ,kj τ,ja,kτ,k + aτ,jτ,kk + a,jτ 2 (cid:1) . 101 Then we multiply the above equation with τ,j and then sum over j, (2 ˙a ˙τ + a¨τ )τ 2 ,kτ,kk ,k + a ˙τ|∇τ|2 t = a|∇τ|2((λ + 2µ),kτ,k) + a(λ + 2µ)τ 2 + 2(λ + 2µ)τ 2 ,k(τ,ka,k) + 2a(λ + 2µ)(τ,jτ,jkτ,k). The term a ˙τ|∇τ|2 t on the left hand side of the above equation is equal to a ˙τ|∇τ|2 t = 2a ˙τ (τ,k ˙τ,k) = aτ,k( ˙τ )2 ,k = aτ,k((λ + 2µ)|∇τ|2),k = a|∇τ|2τ,k (λ + 2µ),k + 2a(λ + 2µ)τ,jτ,jkτ,k, (4.18) since we are talking about the P-wave mode now and its phase τ satisfies the Hamiltonian- Jacobi equation ˙τ 2 − (λ + 2µ)|∇τ|2 = 0. Consequently, the transport equation about the norm of the P-wave’s amplitude vector A is ˙a + (λ + 2µ)a,kτ,k G + a 2G ((λ + 2µ)trace(M ) − ¨τ ) = 0. (4.19) Notice the second order derivatives of the phase function τ is involved, and the trans- port equation will be undefined if the phase function is not smooth. Lemma 4.2.1 guarantees a well-defined transport equation, while the classical Geometrical-Optical ansatz fails at the caustics region. Following [48], the ODE about the norm of ampli- 102 tude can be added into the P-wave’s dynamics by using equation (4.19) and dx dt = Gp, da(t, x(t)) dt + a 2G ((λ + 2µ)trace(M ) − Gx · Gp − GT p M Gp) = 0. (4.20) 4.2.3 Transport Equation Governing S-wave’s Amplitude Vec- tors Now let’s see the S-wave’s case. We abbreviate τ S as τ until the end of this section. Again, we first separate the amplitude vector as AS = aD, where D is a vector which is orthogonal to the ray direction ∇τ S and its norm is fixed to be a constant. After substituting A = aD into equation (4.17), we will have the following equation for the amplitude’s jth component, 0 = 2( ˙aDj + a ˙Dj) ˙τ + a¨τ Dj − a(µ,kDk)τ,j − a(µ,kτ,k)Dj − (λ + µ)((aD)k,kτ,j) − 2µ(aDj),kτ,k − µaDjτ,kk, (4.21) for j = 1, 2, 3. After multiplying Dj with the above equation (4.21), we sum over the index j. (cid:0)aµ,kτ,k + aµτ,kk + 2µa,kτ,k (cid:1)|D|2 +2µaτ,k|D|2 ,k = a¨τ|D|2 +2 ˙a ˙τ|D|2 +a ˙τ ˙|D|2 , (4.22) since D is orthogonal to ∇τ . The last term in the above equation is zero as the norm of D is fixed, we have (∇τ · ∇)|D|2 = 0. (4.23) 103 Therefore, the above equation can be simplified as, aµ,kτ,k + aµτ,kk + 2µa,kτ,k = 2 ˙a ˙τ + a¨τ . (4.24) We fix the norm of D to be one for convenience and the above equation (4.24) provides the way to calculate the amplitude’s norm a. It is not the same equation as the transport equation in the scalar wave equation. To yield the same equation, we divide √ µ on both sides of equation (4.24), 2 ˙a ˙τ + a¨τ√ µ = a µ = 2τ,k( √ µ,kτ,k + a µτ,kk + 2 √ µa)τ,kk √ µa),k + ( √ µa,kτ,k √ If we set ˜a = µa as new amplitude, then 2 ˙˜a ˙τ + ˜a¨τ = µ(2τ,k˜a,k + ˜aτ,kk) (4.25) After yielding the same transport equation, we can obtain the same ODE as the P-wave (4.20). Unlike the P-wave, we still need one more equation in the S-wave’s ray system to describe the amplitude vector AS’s direction D. To obtain the equation about the amplitude direction D, we would like to plug equation (4.24) into equation (4.21) and the coeeficients in front of the direction D is zeros suggested by equation (4.24), 2a ˙τ ˙Dj − 2µa(Dj,iτ,i) = τ,j (cid:0)aµ,iDi + (λ + µ)(aD)k,k (cid:1) (4.26) 104 The left hand side of equation (4.26) is equal to the term dDj (t,x(t)) dt , since ˙τ dDj dt dxi dt = ˙τ ˙Dj + ˙τ Dj,i = ˙τ ˙Dj − G(x(t), p(t))DT Gp(x(t), p(t)) = ˙τ ˙Dj − µDj,iτ,i(t, x(t)) (4.27) We know that the left hand side of equation (4.26) is 2a ˙τ of equation (4.26) is parallel to ∇τ . Therefore, dD dt dDj dt , and the right hand side is parallel to ∇τ . Together with the fact that the amplitude’s direction D is always perpendicular to the ray direction p(t) = ∇τ (t, x(t)), dDkpk(t) 0 = dt (cid:18) dpj(t) dDk 0 = dt = − pk(t) + Dj dt Dk dpk(t) (cid:19) pk(t) dt |p(t)|2 dDk dt (4.28) 4.2.4 Single Beam Solution for P and S-wave To summarize the ODE dynamics generated by the method of characteristics, we have dx dt dp dt dM dt da dt = Gp(x(t), p(t)), x(0) = x0 = −Gx(x(t), p(t)), p(0) = p0 = −(Gxp)T M − M Gpx − M GppM − Gxx, M (0) = iI = − a 2G (c2trace(M ) − Gx · Gp − GT p M Gp, A(0) = A0 (4.29) 105 where the velocity term c2 = λ+2µ for the P-wave and c2 = µ for the S-wave. The term G is the corresponding Hamiltonian. There is one extra equation about the direction D in the S-wave’s dynamics, = − dDk dt (cid:18) dpj(t) dt Dj (cid:19) pk(t) |p(t)|2 , D(0) = D0. The initial condition of the system above will be given by the Multiscale Gaussian Wavepacket transform, which will be specified in the later section. Now the way of propagating the phase functions τ P,S and the amplitude vectors AP,S is provided, and it allows us to finish the construction of a single-beam asymptotic solution, ΦP (t, x) = a(t)∇τ (t, x(t))eiωτ (t,x) ΦS(t, x) = a(t)Deiωτ (t,x), (4.30) (4.31) and the phase function is approximated by the Taylor expansion near the central ray, τ P,S(t, x) = ∇τ P,S · (x − x(t)) + = p(t) · (x − x(t)) + 1 2 (x − x(t))T M P,S(t)(x − x(t)) 1 2 (x − x(t))T M P,S(t)(x − x(t)). (4.32) The Gaussian profile is offered by the imaginary part of the Hessian matrix M (cid:16)−ω 2 (cid:17) exp (x − x(t))T Im(M (t))(x − x(t)) . (4.33) Suggested by Lemma 4.2.1, a beam ansatz will be always well localized throughout the propagation. 106 4.3 Multiscale Gaussian Wavepacket Transform for Elastic Waves The initial condition of the elastic wave equation (4.3) can be any general L2 vector- valued function, and it is not necessary to take the exact form like, (cid:19)(cid:19) . (4.34) (cid:18) (cid:18) A exp iω p(0)T (x − x0) + 1 2 (x − x0)T M (0)(x − x0) The problem here is that how to decompose any general L2 function to multiple Gaus- sian wavepackets like the above form (4.34) efficiently and make the total number of beams to be calculated as small as possible. We will provide a very brief introduction of the Multiscale Gaussian Wavepacket Trans- form [48] for the scalar functions first in this section. More details can be found in [48]. Then its extension designed for the vector-valued initial conditions is presented afterwards. 4.3.1 Multiscale Gaussian Wavepacket Transform: Vector Func- tions After proposing the Multiscale Gaussian Wavepacket transform for the scalar function in Section 2.2, we would like to extend this idea to the vector-valued function f . Here we assume that each component of the vector fj is a L2 function. 107 4.3.1.1 Decomposition of the Single Wavepacket Suppose we have already applied the wavepacket transform to each component of the initial condition f , i.e. f = =   f1 f2 ... fd (cid:88) κl,i,kφl,i,k. l,i,k The idea here is to decompose each single Gaussian wavepacket into the sum of the P-wave and the S-wave. Let the unit vector vl,i,k be κl,i,k = κT l,i,kvl,i,k vl,i,k + κl,i,k. (4.35) The first term on the right hand side of equation (4.35) is the P-wave component, and the initial condition for its amplitude vector can then be written as (cid:16) (cid:16) (cid:17) (cid:17) ξl,i |ξl,i|, then (cid:16) Id − vl,i,k(vl,i,k)T(cid:17) (cid:32) (cid:33) ξl,i |ξl,i|2 κT l,i,kvl,i,k vl,i,k = κT l,i,k ξl,i. (4.36) To initialize the S-wave, we have to specify initial directions which are orthogonal to each other and all of them are supposed to be orthogonal to vl,i,k. 108 Therefore, we choose the first direction D(1) as the unit vector of the first column (cid:16) Id − vl,i,k(vl,i,k)T(cid:17) (cid:16) Id − vl,i,k(vl,i,k)T(cid:17) vector of the matrix and apply the Gram-Schmidt process to generate the rest D(m) for m = 2,··· , d − 1. And we notice that every column vector of the matrix is orthogonal to vl,i,k, therefore, so are their linear combinations. The corresponding amplitude norm am, (cid:16) Id − vl,i,k(vl,i,k)T(cid:17) am = κT l,i,k D(m), m = 1, 2,··· , d − 1. (4.37) 4.3.1.2 Preprocessing the Initial Condition Before applying the method described above, we pre-process the initial condition first. Following the same technology employed in [47], there are supposed to be two different l,i,k and κ− branches κ+ Hamiltonian ±c(x)|p|, where c(x) is the corresponding velocity. l,i,k for each wavepacket corresponding to different signs of the To satisfy both the initial wavefield f and the initial velocity g, we define l,i,k + κ− (κ+ l,i,k)φl,i,k = κl,i,kφl,i,k = f . Taking the derivative of the wavefield about the time variable t yields l,i,k)(cid:0)i|ξl,i|G+(x0, p0)(cid:1) φl,i,k ≈ Ξl,i,kφl,i,k = g, (κ− l,i,k − κ+ (4.38) (4.39) where Ξl,i,k is the coefficients generated from decomposing the initial velocity g. Here 109 the left hand side of equation (4.39) is not the complete form of the beam’s time derivative, instead, we pick the leading order term to approximate. After solving the coefficients κ+ l,i,k and κ− l,i,k based on equation (4.38) and equa- tion (4.39), we can apply the decomposition described by equations (4.35) - (4.37) to κ+ l,i,k and κ− l,i,k respectively. The summary of the vector-version Multiscale Gaus- sian Wavepacket is provided below. Algorithm 3 Discrete Vector-Valued Gaussian Wavepacket Transform 1.Call the Discrete Gaussian Wavepacket Transform for each component in the discrete signal f and g (cid:18) l,i,k and κ− 2. Use equations (4.38) and (4.39) to compute κ+ κ± l,i,k · ξl,i |ξl,i|2 3. Generate P-wave with the amplitude vector 4. Generate S-waves amD(m) by equation (4.37), for m = 1, 2,··· , d − 1. (cid:19) ξl,i l,i,k The above process defines the initial amplitude vectors for the P-wave and the S-wave, and the initial value of the phase function and its derivatives are given by the Multiscale Gaussian Wavepacket transform of the scalar form, that is dx dt dp dt dM dt da dt dDj dt = Gp(x(t), p(t)), x(0) = k Ll = −Gx(x(t), p(t)), p(0) = 2π ξl,i |ξl,i|, = −(Gxp)T M − M Gpx − M GppM − Gxx, M (0) = i(2π2σ2 = − a 2G dτ,k dt τ,j |∇τ|2 D(m)(0) = D(m). (c2trace(M ) − Gx · Gp − GT (cid:18)(cid:114) π p M Gp, a(0) = LlN (cid:19)d l /|ξl,i|)I. σl . = ( Dk) 110 4.4 Multiscale Gaussian Beam Method for Periodic Boundary Value Problem In the above section, we have demonstrated the way to decompose the vector-valued initial conditions in the L2 space. To solve the periodic boundary problem,each param- eters and solutions are assumed to be periodic functions. Meanwhile, the central ray in the periodic boundary problem will be smooth along the propagation in the sense of modules. The principle shown in Figure 4.1 will be employed to solve the periodic boundary value problem. The red dashed line represents the wavepacket leaving the domain [0, 0.5], as Figure 4.1: Periodic Boundary Problem: The Case Wavepacket leaving the Boundary. the left half goes beyond the domain, while the right half still shows up. The missing left half of the beam solution will enter from the other side with the same shape, which is the blue line in the graph suggested by the periodic boundary condition. The cubic region [0, 0.5]3 is chosen to test the correctness of our algorithm. We show numerical results in Section 4.8. 111 00.10.20.30.40.500.10.20.30.40.50.60.70.80.91xu 4.5 Multiscale Gaussian beam method for Homo- geneous Dirichlet Boundary Condition In this section, we would like to explore the solution to the homogeneous Dirichlet Boundary condition. From now on, we use the 3-D space {x = (x, y, z) : x, y, z ∈ R} as our model. The wavefield on the boundary is assumed to be zero in this section. When the reflection happens, the sum of all wavefields at time tr and the central point x(tr) of the ray should vanish, i.e. u(tr, x(tr)) = 0. The time when the central point x(tr) of the ray is on the boundary is defined as the reflection time. From now on, all equations below in this section are defined on the point (tr, x(tr)), if not specified. The Hamiltonian used in this section is assumed to be positive G = c(x)|p|, and the negative Hamiltonian will be treated similarly. 4.5.1 P-wave Reflecting Beams: Ray Direction When the P-wave reflection happens, the total wavefield is made up by three different sources, the original P-wave Gaussian beam, the new P-wave beam after the reflection (PP-wave) and the new S-wave beam (PS-wave). At the reflection point x(tr), −aP eiτ P ∇τ P = aP P eiτ P P ∇τ P P + aP Seiτ P S DP S. (4.40) Both P and S-wave will be generated after reflection. All phase functions should have the same value at the reflection point, τ P = τ P P = 112 τ P S. Otherwise, if we change the value of large wavenumber |ξl,i|, the homogeneous boundary condition will be violated. The principle to derive new phase functions τ P P and τ P S is to take advantage of the continuity conditions, i.e. the continuity of the tangential components of the first order derivatives of τ , so τ P y = τ P P y = τ P S y τ P z = τ P P z = τ P S z , , (4.41) where we assume the reflection happens along the surface {x = (x, y, z) : x = 0}. Besides the spatial variables, the partial derivative of the phase function with respect to the time variable t should also follow, ˙τ P = ˙τ P P = ˙τ P S ⇒ cP|∇τ P| = cS|∇τ P S|, cP|∇τ P| = cP|∇τ P P|. (4.42) √ where cP = λ + 2µ is the velocity of the P-wave and cS = √ µ is the velocity of the S-wave. The partial derivatives along the tangential directions of the boundary can be obtained directly from equation (4.41). To obtain the momentum along the reflection direction or the normal direction of the boundary, one needs to use equations (4.42) and (4.41) collectively. x = −sign(τ P τ P S x ) (cid:115)(cid:18) cP cS (cid:19)2 |∇τ P|2 − (τ P S y 113 )2 − (τ P S z )2. (4.43) Here we focus our derivation on the reflection from P-wave to S-wave, while the rule of the reflection between the same mode can follow the same way, so we skip it here. The only ambiguity left here is the case when a beam hits the boundary at a corner since it causes diffraction and the above derivation does not apply any more. Here we simply ignore the situation when a beam hits a corner of the domain since the Gaussian method is asymptotic. The numerical accuracy will not be degraded without those beams as those diffractions have exponentially small effects. 4.5.2 P-wave Reflecting Beams: The Hessian of the Phase To illustrate the derivation of the second order derivative terms, we pick three entries among six distinct entries in the Hessian for explanation, τyy, τxy and τxx, since all other entries can be classified into one of these three types. Again, the reflection is assumed to happen along the surface {x = (x, y, z) : x = 0} and all terms without arguments are defined at the reflection point. To start with the first type τyy, which is tangential component τ P yy = τ P S yy = τ P P yy . (4.44) τzz and τyz will also stay the same. To derive the second type of terms τ P S xy , we use the partial derivatives about the time 114 variable t, ˙τ P y = cP ˙τ P S y = cS y |∇τ P| + cP (∇τ P y )T∇τ P |∇τ P| y |∇τ P S| + cS (∇τ P S y |∇τ P S| )T∇τ P S The notation used here is ˙τy = ∂2τ ∂t∂y . We have ˙τ P y = ˙τ P S y , , (4.45) (4.46) (4.47) due to the continuity of the tangential and time component. Now substitute equation (4.45) and equation (4.46) into equation (4.47), cP y |∇τ P| + cP (∇τ P (∇τ P S y )T∇τ P |∇τ P| )T∇τ P S = y = cS τ P S xy = y y |∇τ P S| + cS (∇τ P S |∇τ P S| )T∇τ P S |∇τ P S| y |∇τ P| + cP (∇τ P y |∇τ P| + cP ∇τ P (cid:32) (cid:32) cS |∇τ P S| cSτ P S x cP cP − cS y )T∇τ P |∇τ P| y · ∇τ P |∇τ P| − cS (cid:33) y |∇τ P S| (cid:33) y |∇τ P S| − 1 τ P S x (τ P S y τ P S yy + τ P S z τ P S yz ). (4.48) To obtain the last type of the term τ P S xx , we need to derive the formula of τ P S tx first. ˙τ P = cP|∇τ P| ⇒ ¨τ P = cP (∇ ˙τ P )T∇τ P ˙τ P S = cS|∇τ P S| ⇒ ¨τ P S = cS (∇ ˙τ P S)T∇τ P S |∇τ P| , |∇τ P S| (4.49) (4.50) . 115 We then have the following equation from equation (4.50) and equation (4.49), cP (∇ ˙τ P )T∇τ P (∇ ˙τ P S)T∇τ P S = |∇τ P| = cS (∇ ˙τ P S)T∇τ P S |∇τ P S| cP (∇ ˙τ P )T∇τ P |∇τ P S| (cid:32)|∇τ P S| cS ˙τ P S x = 1 τ P S x |∇τ P| cP (∇ ˙τ P )T∇τ P |∇τ P| cS (cid:33) (4.51) − ˙τ P S y y − ˙τ P S τ P S z τ P S z With the formula about the term ˙τ P S x given above, the term τ P S xx can be obtained by solving the following equation, ˙τ P S = cS|∇τ P S| ⇒ ˙τ P S x = cS x|∇τ P S| + cS (∇τ P S x )T∇τ P S |∇τ P S| . (4.52) To remark, τ P S x will not be zero as we assumed our initial conditions are compactly supported. 4.5.3 P-wave Reflecting Beams: Amplitude Vector We have so far already derived the initial condition of all terms involved with the phase function for our new ODE dynamic system after P-wave reflection. Since the phase function itself does not change after reflection in the center of the beam, we will have the following equation about the amplitude to satisfy the homogeneous boundary condition at the reflection point x(tr), −aP∇τ P = aP P∇τ P P + aP SDP S. (4.53) 116 We know that the PS-wave’s amplitude direction DP S is orthogonal to its ray direction ∇τ P S by the definition and with unit norm by our restriction. Consequently, the following equation can be obtained if we project both sides of equation (4.53) to the vector ∇τ P S at the same time. −aP (∇τ P )T∇τ P S = aP P (∇τ P P )T∇τ P S, aP P = −aP (∇τ P )T∇τ P S (∇τ P P )T∇τ P S . (4.54) Like the initial condition, aP SDP S is the summation of two S-waves. (cid:88) aP SDP S = αP S i D(i). (4.55) i Similar to the initial condition, we pick the first direction D(1) to be the first column vector of the matrix I3 − vvT and v = ∇τ P S |∇τ P S|. Then, the second vector D2 will be D(2) = D(1) × v, where × represents the cross product between two vectors. After normalizing each direction, we can project the residual −aP∇τ P − aP P∇τ P P to each direction to obtain the amplitude αP S i . 4.5.4 S-wave Reflecting Beams: the Phase term Similar to the P-wave reflection, we have the following equation to satisfy the homo- geneous boundary condition for the S-wave wavepacket at the reflection point x(tr). −aSeiτ S DS = aSP eiτ SP ∇τ SP + aSSeiτ SS DSS. (4.56) 117 After taking first glance of the above equation, the S-wave reflection dynamics seems to be the same as the P-wave reflection developed in the last section. However, we will find that there will be some significant difference between the S-wave reflection and the P-wave reflection. To begin with, let us still use the case hitting the surface {x = (x, y, z) : x = 0} as an example, τ S y = τ SS y = τ SP y τ S z = τ SS z = τ SP z , , τ S t = τ SS t = τ SP , ⇒ cS|∇τ S| = cP|∇τ SP|. t With all these equations combined, we will get x = −sign(τ S τ SP x ) (cid:19)2 − (τ S y )2 − (τ S z )2, (cid:115)(cid:18) cS|∇τ S| cP (4.57) (4.58) (4.59) (4.60) (4.61) √ λ + 2µ. This and the S-wave’s velocity cS = √ µ is less than the P-wave’s cP = leads to the possibility that the part inside the square root in equation (4.61) will be negative, or equivalently, τ SP x can be in general a pure imaginary number. It will be a disaster, since there will be some exponentially increasing wave on one side of the boundary. Let’s consider the regular situation first, in which case, equation (4.61) is a real number. It is same as the one employed in the P-wave reflection as illustrated above, so we skip it here. 118 The second case is when the term τ SP is a pure imaginary number, i.e. cS|∇τ S| cP + (τ S y )2 + (τ S z )2 ∇τ sp = (4.62) (cid:115) − i x (cid:18)  (cid:19)2 τ S y τ S z  . This phenomenon is called the evanescent wave and the energy will fade away quickly around the boundary in this case. Therefore, there is no need to derive its Hessian due to its small energy. 4.5.5 S-wave Reflecting Beams: the Amplitude Vector The evanescent wave fades away quickly, however, we will still include SP-wave’s am- plitude vector in our derivation so as to make the derivation easier. Moreover, we need the nonzero amplitude vector ASP to make the homogeneous boundary assumption true at the reflection point x(tr). To summarize, −aSDS = aSP (Re(∇τ sp) + iIm(∇τ sp)) + aSSDSS. (4.63) The way we get the amplitude vector ASP is still from the same idea by using the fact that the SS wave’s amplitude direction DSS is orthogonal to its ray direction v, (cid:0)Re(ASP )Re(∇τ SP ) − Im(ASP )Im(∇τ SP )(cid:1) · v = −Re(aS)DS · v; (cid:0)Re(ASP )Im(∇τ SP ) + Im(ASP )Re(∇τ SP )(cid:1) · v = −Im(aS)DS · v. (4.64) (4.65) 119 The amplitude of the SP-wave’s amplitude ASP can be obtained by solving the above system. Consequently, the residual −aSDS − aSP∇τ SP is now well defined. Following the same process defined in the PS wave case, we can set up the amplitudes and directions easily for SS-wave. Notice that the ray after reflection is no longer smooth, which means that Lemma 4.2.1 is not applicable when reflections happen. Naturally, one needs to show that the imaginary part of the Hessian after reflections defined above will still be positive definite, especially for the PS wave and the SP wave case. In [4], authors have already proved this is true for the PP wave and the SS wave, i.e. the conversion between same wave modes. The proof about the conversion between different wave modes is provided in Appendix B. 4.5.6 Method of Images for Boundary Conditions In [4], authors have proposed a method to tackle the problem caused by partially reflected beams. The partial reflection problem means the frontier part of a beam is needed to be reflected back even when its central ray has not hit the boundary yet and consequently the reflection dynamics has not been called. This is due to the fact that beams have nonzero width and illustrated in the following graph. Therefore, some modifications should be added to these partial reflection cases so that the homogeneous boundary condition is always satisfied as well as our wavefield remains to be continuous. Our strategy presented here is that the outer part is considered to be reflected back, which is carried by some artificial beams. So we essentially apply the odd extension to those beams, like what Figure 4.3 shows. The trajectory of the 120 Figure 4.2: Partially Reflected beams Figure 4.3: Partially Reflected Beams with Odd Extension blue dashed wavepacket in the graph is completely determined by its associates, the red solid wavepacket in the graph. It implies that we don’t use any extra assumption of the velocity outside the domain. The blue dashed wavepacket only serves as a supplementary beam to satisfy the vanishing boundary condition. 121 −101−1−0.500.51xu−1−0.500.51−1−0.500.51xu 4.6 Stationary Phase Analysis of Beams To reinitialize or sharpen the single Gaussian beam ansatz, we need to apply some stationary phase analysis to the single beam. So we first list some lemmas and com- putations here which are needed to implement the reinitialization process in the next section. For any function u in L2(Rd), there is a phase space decomposition method such that, (cid:16) ω (cid:17)3d/2(cid:90) 2d/2eiω(p(x−x (cid:48) (cid:48) )−p( ˜x−x − ω|x−x 2 ))e (cid:48)|2 (cid:48)|2 − ω| ˜x−x 2 e u( ˜x)d ˜xdpdx (cid:48) . u(x) = (4.66) 2π R3d Here x, ˜x and x (cid:48) are points in the spatial space Rd and p is the dual momentum variable in the frequency space. ω is a fixed parameter determining the size of Gaussian window functions.Here we would like to explore how to apply representation (4.66) to Gaussian beams without considering its amplitude’s direction. Using P-wave as an example, consider ΦP (t, x) = a(t)eiτ (t,x) instead of a(t)eiτ (t,x)p(t). Moreover, all beam functions considered here are treated as single-variable functions by assuming a principal variable while other variables are fixed. In all the following derivations, the principal variable is assumed to be the variable y, while other variables x and z are fixed. 122 4.6.1 Stationary Phase Approximation with Respect to Spa- tial Variables The Gaussian beams u with y as the principal variable considered in this section are in the following form, u(y; x, z, t) = A(x, z, t)eiΦ(y−y0)e 2 Im(τyy)(y−y∗ − 1 0)2 , (4.67) where (cid:0)x0, y0, z0 (cid:1) is the central point of the beam ansatz. All functions below are defined at this point if not specified, and Φ(y−y0) = τy(y−y0)+ Re(τyy)(y − y0)2 2 +Re(τxy)(y−y0)(x−x0)+Re(τzy)(y−y0)(z−z0). (4.68) τ is the phase function. To define the scalar y∗ 0, we have the following equation where all second order derivative terms are imaginary part only, (cid:18) x − x0, y − y0, z − z0 (cid:19) (cid:18) y − y0 + = τyy  τxx τxy τxz τxy τyy τyz τxz τzy τzz    x − x0 y − y0 z − z0 (cid:19)2 = τyy(y − y0)2 + 2(τxy(x − x0) + τxz(z − z0))(y − y0) + B(x, z) τxy(x − x0) + τxz(z − z0) τyy + ˜B(x, z). (4.69) The complete square term in equation (4.69) is defined as (y − y∗ 0)2 and ˜B(x, z) is a constant with respect to y. 123 To make the expression clear, we denote R = Re(τyy) and I = Im(τyy) throughout this section. Applying decomposition (4.66) to u(y; x, z, t), we have u(y; x, z, t) = A(x, z, t) √ 2eiωp(y−y (cid:48) − ω 2 |y−y )e (cid:48)|2 (cid:48) (cid:48) , )dpdy ψ(p, y (4.70) (cid:16) ω (cid:17)3/2(cid:90) 2π R2 and (cid:90) (cid:48) ψ(p, y ) = ei ˜B(x,z) e−iωp(˜y−y (cid:48) − ω 2 |˜y−y )e (cid:48)|2 eiΦ(˜y−y0)e 2|˜y−y∗ − I 0|2 d˜y. (4.71) We will apply the stationary phase approximation described in the following lemma to ). calculate ψ(p, y Lemma 4.6.1. We consider the behavior of I(ω) =(cid:82) b a f (t)eiωg(t)dt, where f and g are smooth enough to admit Taylor approximation near some appropriate points in [a, b], and g is real-valued. Suppose there is some point c0 ∈ [a, b] and g (t) (cid:54)= 0 everywhere (cid:48) else in the closed interval [a, b]. Moreover, g (cid:48)(cid:48) (c0) (cid:54)= 0. When ω >> 1, we have (cid:48) (cid:90) b a (cid:115) (cid:18) 1 (cid:19) ω I(ω) = f (t)eiωg(t)dt = f (c0)eiωg(c0)eiπδ/4 2π (cid:48)(cid:48) (c0)| + O ω|g , (4.72) where δ is the sign of g (cid:48)(cid:48) (c0). To use Lemma 4.6.1, we first substitute y (cid:48) = y0 + ma into u(y; x, z, t), and the value of a and ω will be defined later. By equation (4.71), we have (cid:90) (cid:18) Φ(˜y−y0) ω iω e (cid:19) −p(˜y−y0−ma) − ω 2 |˜y−y0−ma|2 e 2|˜y−y∗ − I 0|2 e d˜y. e−i ˜B(x,z)ψ(p, y0 + ma) = (4.73) 124 To compute the critical point ˜y0 of the phase term Φ(˜y−y0) ω − p(˜y − y0 − ma), we have (˜y − y0) ω (cid:48) Φ R(˜y − y0) + Re(τxy)(x − x0) + Re(τyz)(z − z0) − p, ω 0 = 0 = + τy ω − p. (4.74) ˜y0 = y0 + ωp − τy − Re(τxy)(x − x0) − Re(τyz)(z − z0) R . (4.75) that is We denote −Re(τxy)(x − x0) − Re(τyz)(z − z0) E(x, z) = ˜y0 − y0 = ωp − τy R R + E. , (4.76) Notice that E(x, z) is independent of the variable y, together with Lemma 4.6.1, the term ψ(p, y0 + ma) is equal to e−i ˜B(x,z)ψ(p, y0 + ma) = eiω(cid:0) Φ( ωp−τy R +E) ω −p( ωp−τy R +E−ma)(cid:1) (cid:115) 2π |R|. 2| ωp−τy − I R +E−y∗ 0+y0|2 × e i π 4 e R|R| 2 | ωp−τy − ω e R +E−ma|2 Substitute the expression of the term Φ into the above equation, (cid:115) 2π (cid:16) ωp−τy (cid:17)−( R·E |R| e R|R| i π 4 ω ) ψ(p, y0 + ma) = ei ˜Be (cid:16) ωp−τy (cid:16) τy ω R +E × e iω 2| ωp−τy R −y∗ − I 0+y0+E|2 (cid:16) ωp−τy (cid:17)−p (cid:17) R +E−ma R +E 2 | ωp−τy − ω e R +E−ma|2 ωp−τy R +E)2(cid:17) + R 2ω ( . (4.77) 125 We insert equation (4.77) into equation (4.70), u(y; x, z, t) = aA(x, z, t) (cid:17)3(cid:114) π (cid:114)(cid:16) ω i π 4 |R| e 2π 2 | ωp−τy − ω R +E−ma|2 (cid:17)2 (cid:16) ωp−τy iω e R +E ≈ 2Aa × e × e i R 2 (cid:114)(cid:16) ω (cid:17)3(cid:90) √ (cid:90) 2π R|R| ei ˜B (cid:16) ωp−τy (cid:16) τy ω R +E − ω 2 |y−y0−ma|2 2eiωp(y−y0−ma)e 2| ωp−τy R −y∗ − I 2 |y−y0−ma|2 − ω eiωp(y−y0−ma)e (cid:16) ωp−τy (cid:16) ωp−τy (cid:17)− R·E (cid:17)(cid:17) e R +E−ma (cid:17)−p R +E ψ(p, y0 + ma)dpdm 0+y0+E|2 ω dpdm. (4.78) By applying the stationary phase approximation to the variable ˜y, we reduce the triple integral (4.66) to the double integral (4.78). Notice that in Lemma 4.6.1, the assumption R (cid:54)= 0 is required. In general, it is not true. However, if R = 0, then the integral about the function ψ(p, y (cid:48) ) is nothing but the Fourier transform of a Gaussian function about y, so we omit it here. 4.6.2 Stationary Phase Approximation with Respect to Mo- mentum Variables Starting from the double integral (4.78), we would like to apply the stationary phase approximation again. However, this time the variable we apply over is the momentum 126 variable p. (cid:115) (cid:114)(cid:16) ω (cid:17)3 2| ωp−τy R −y∗ − I (cid:16) ωp−τy (cid:17)− R·E 2π i π 4 (cid:90) R|R| a − ω 2 |y−y0−ma|2 4π |R|e e 2 | ωp−τy − ω 0+y0+E|2 (cid:17)−p (cid:16) ωp−τy (cid:17) R +E−ma R +E−ma|2 (cid:16) ωp−τy + R 2ω R +E e ω dm R +E u(y; x, z, t) = A(x, z, t)ei ˜B (cid:90) × iω × e eiωp(y−y0−ma)e (cid:18) τy (cid:16) ωp−τy R +E ω (cid:17)2(cid:19) dp. (4.79) For the inner integral of equation (4.79), the phase function g(p) is (cid:19) + E (cid:18) ωp − τy (cid:18) (cid:19) R − p E − ma + τy ω ωp − τy R g(p) = p(y − y0 − ma) + (cid:18) − R · E ω E + (cid:19) (cid:18) ωp − τy R + R 2ω ωp − τy R (cid:19)2 + E , and the smooth function f (p) is, 2 | ωp−τy − ω R +E−ma|2 e f (p) = e 2| ωp−τy − I R −y∗ 0+y0+E|2 . (4.80) To compute the critical point, (cid:48) g (p) = y− y0− ma + − E− τy R (cid:18) E − ma + ωp − τy R (cid:19) (cid:18) ωp − τy R − pω R + (cid:19) + E , (4.81) 127 then the critical point p0 is ωp R = y − y0 − E + (cid:0)y − y0 − E(cid:1) + τy R R ω τy ω = y − y0 − E. ⇒ p0 = ⇒ ωp0 − τy R The second order derivative of the phase function g at the critical point p0 is (cid:48)(cid:48) g (p0) = − ω R . (4.82) (4.83) So the approximation of the inner integral of equation (4.79) is (cid:90) iω × e =(cid:114) πR eiωp(y−y0−ma)e (cid:18) τy (cid:16) ωp−τy ω R +E 2| ωp−τy R −y∗ − I (cid:17)− R·E (cid:16) ωp−τy ω R +E 0+y0+E|2 2 | ωp−τy − ω (cid:17)−p (cid:16) ωp−τy e R +E−ma R +E−ma|2 (cid:17) + R 2ω (cid:16) ωp−τy R +E (cid:17)2(cid:19) dp (cid:18) ω2 e −i πR 4|R| eiωp0(y−y0−ma)e (cid:18) τy−R·E (y−y0) (cid:19)(cid:19) ω 2|y−y∗ − I 0|2 (cid:16)−p0(y−y0−ma)+ R e − ω 2 |y−y0−ma|2 2ω (y−y0)2(cid:17) iω e . iω × e We summarize the stationary analysis conducted in this section in the following lemma, Lemma 4.6.2. Suppose the principal variable selected for the function u(y; x, z, t) is y u(y; x, z, t) = A(x, z, t)eiΦ(y−y0)e 2 Im(τyy)(y−y∗ − 1 0)2 , (4.84) 128 and R = Re(τyy) is nonzero, then the following decomposition can be obtained, (cid:114) ω 2|y−y∗ − I 0|2 e 2π ei˜τy(y−y0)ei R 2 (y−y0)2 dm + O (cid:18) 1 (cid:19) ω (4.85) u(y; x, z, t) = aA(x, z, t)ei ˜B(x,z) (cid:90) e−ω|y−y0−ma|2 where ˜τy = τy − R · E and R · E = −Re(τxy)(x − x0) − Re(τyz)(z − z0). (4.86) The fixed parameter a is defined as a = 1√ I . The value of ω will be specified later, but its order is O(I). 4.7 Sharpening Beams by Reinitialization In this section, we would like to propose a new reinitialization strategy based on Lemma 4.6.2 from Section 4.6. Again, we base our proof on the assumption that the variable y is the principal variable. We first illustrate the reason why proposing a new reini- tialization strategy is necessary. 4.7.1 The First Motivation for Developing a New Reinitial- ization Strategy It is necessary to add a reinitialization process into the propagation since the width of a beam will increase exponentially in some generic medium [48]. We use the linear velocity and 1-D problem for explanation. Suppose c(x) = α + βx where α and β are 129 constants, then the Riccati equation about the Hessian M will be dM dt + 2M β = 0, M (0) = i. (4.87) Solving this simple linear ODE, we have M (t) = ie−2βt. (4.88) If the slope β > 0, then the width of the beam solution will be exponentially increasing. As we can see, the beam solution will lose its accuracy in the simple linear velocity, and each smooth velocity can be approximated by a linear function locally, therefore, the same phenomenon can be expected in other situations. 4.7.2 The Second Motivation for Developing a New Reinitial- ization Strategy The second motivation is to resolve the problem caused by reflection beams. The idea in Section 4.5 we have used to derive the reflection formula is theoretically correct, however, it will cause some problems when implementing it numerically, especially in the S-wave reflection case. The difference is that the SP-wave is more likely to be a grazing beam. To see this, we employ the 2D model for the illustration and let’s suppose the ray hitting the line {x = (x, y) : x = 0}. Then, according to the analysis in Section 4.5, 130 the y-component of the ray direction τy does not change. For the PS reflection, (cSτ P S x )2 = (cP|τ P x |)2 + (λ + 2µ)|τ P (τ P S x )2 = while the SP reflection follows, (cP )2 − (cS)2(cid:17) (cid:16) x |2 + (λ + µ)τ 2 (cS)2 y τ 2 y , (4.89) (τ SP x )2 = µ|τ S x |2 − (λ + µ)τ 2 y (cP )2 . (4.90) We then compute the angle θP S and θSP between the ray direction and the reflecting boundary {x = (x, y) : x = 0}, tan(θP S) = = = τ P S x τ P S y (cid:118)(cid:117)(cid:117)(cid:116)(λ + 2µ)|τ P (cid:115) λ + µ + µ y x |2 + (λ + µ)τ 2 (cS)2τ 2 y (λ + 2µ)|τ P x |2 . µτ 2 y Similarly, (cid:115) tan(θSP ) = µ|τ S x |2 (λ + 2µ)τ 2 y − λ + µ λ + 2µ . (4.91) (4.92) We claim that for the PS reflection, the angle between the ray direction and the boundary will be increasing after reflection, while for the SP reflection, this value will 131 be decreasing. To see this, tan(θP S) tan(θP ) (cid:33)2 = = λ+µ µτ 2 y µ + (λ+2µ)|τ P x |2 (cid:19)2 (cid:18) (cid:18) τy λ + µ τ P x τy + λ + 2µ µ µ τ P x (cid:19)2 . (cid:32) (cid:32) Apply the same idea to the SP reflection, (cid:33)2 tan(θSP ) tan(θS) = µ λ + 2µ − λ + µ λ + 2µ (cid:18) τy (cid:19)2 τ S x . (4.93) (4.94) As we can see from equation (4.93) and equation (4.94), as τy is increasing or the incidence angle is decreasing, the ratio for the PS reflection is increasing, which means that the angle after reflection is larger than the incidence angle, while the angle for the SP reflection is decreasing as a quadratic function of τy. It means that the angle for the SP wave θSP will be closer to zero even when the incoming S-wave’s incidence angle θ is away from zero. The grazing beam with larger width will interact with the boundary. Therefore, it is needed to be sharpened to guarantee the accuracy. To remark, there’s no problem with the reflection in the acoustic wave as the velocity is the same. We have conducted the experiment to justify the analysis in 3-D space and we show the result in the next section. 132 4.7.3 Sharpened Wavepackets and Convergence Analysis By taking advantage of Lemma 4.6.2, we have (cid:114) ω 2π u(y) = aAei ˜B (4.95) 2|y−y∗ − I 0|2 e ei˜τy(y−y0)ei R 2 (y−y0)2(cid:90) e−ω|y−y0−ma|2 dm + O (cid:18) 1 (cid:19) ω , The term ˜τy is the modified y-direction of the central ray, that is ˜τy = τy − R · E, (4.96) To sharpen Gaussian beams, we have the following lemma, Lemma 4.7.1. u(y) ≈ Aei ˜B (cid:114) ω 2π lka q(cid:88) k=0 ei˜τy(y−y0)e −ωk|y−y∗ 0|2 2|y−y∗ − I 0|2 e 2 (y−y0)2 ei R . (4.97) where a = 1√ I . Parameters ωk and q will be given in the proof. As we can see the extra term e −ωk|y−y∗ 0|2 reduces the size of beams. And positive ωk is obtained by choosing parameter ω appropriately. 133 Proof. To obtain equation (4.97), we first (cid:90) e−ω|y−y0−ma|2 dm = = (cid:90) (cid:90) = e 0+y∗ −ω|y−y∗ e −ω|y−y∗ 0|2 e −ω|y−y∗ 0|2(cid:90) e 0−y0−ma|2 −ω|y∗ dm 0−y0−ma|2 −2ω(y−y∗ e 0−y0−ma|2 −2ω(y−y∗ 0)(y∗ 0)(y∗ e −ω|y∗ e 0−y0−ma) 0−y0−ma)dm. (4.98) The integral part above is e −ω|y∗ (cid:90) (cid:90) ˜k+ 1 (cid:88) (cid:90) 1 (cid:88) ˜k− 1 2 ˜k 2 = k∈Z 2 − 1 2 0−y0−ma|2 −2ω(y−y∗ e 0−y0−ma)dm = −ω|y∗ e 0−y0−ma|2 0)(y∗ 0−y0−ma)dm 0)(y∗ −2ω(y−y∗ e e−ω|(k+δ)a|2 e2ω(y−y∗ 0)((k+δ)a)dδ. (4.99) We require k to be integers, which means the value of ˜k satisfies k = ˜k + . Notice first that we can truncate the above summation to finite terms, |k| ≤ q, since e−ω|y−y0−ma|2 is a L1 function. By monotone convergence theorem, a 0 y0−y∗ e−ω|y−y0−ma|2 dm = 0 (4.100) lim N→∞ . Case 1: k = 0 (cid:90) ∞ N 134 2 − 1 2 (cid:90) 1 e−ω|(k+δ)a|2 (cid:90) 1 (cid:90) 1 e−ω|δa|2 2 − 1 2 ≈ e−ω|δa|2 ≈ 2 − 1 2 (cid:90) 1 (cid:33) 2 − 1 2 0)((k+δ)a)dδ = e2ω(y−y∗ (cid:32)(cid:88) (cid:32)(cid:88) n n 2nωn(y − y∗ n! (2ω)2n(y − y∗ (2n)! 0)n(δa)n dδ (cid:33) 0)2n(δa)2n e−ω|δa|2 e2ω(y−y∗ 0)(δa)dδ dδ (4.101) The odd power terms in δ vanish in the last step above since e−ω|δa|2 is an even function about δ and all odd power functions are odd functions. The integral of all odd functions in [− 1 2 , 1 2] will be zero. (cid:90) 1 2 − 1 2 e−ω|δa|2 e2ω(y−y∗ 0)(δa)dδ ≈ ≈ (cid:90) 1 (cid:90) 1 2 − 1 2 2 − 1 2 e−ω|δa|2(cid:16) 0)2(δa)2(cid:17) 1 + 2ω2(y − y∗ dδ e−ω|δa|2 dδ. (4.102) 0)2(δa)2 e2ω2(y−y∗ (cid:18) (2ω)4(y−y∗ 0)4(δa)4 4! (cid:19) . The leading order error from equation (4.102) is O Case 2: k (cid:54)= 0 135 When k < 0, (cid:90) 1 2 − 1 2 e−ω|(k+δ)a|2 e2ω(y−y∗ 0)((k+δ)a)dδ = = = 2 − 1 2 (cid:90) 1 (cid:90) 1 (cid:90) 1 2 − 1 2 2 − 1 2 e−ω|−|k|a+δa|2 e2ω(y−y∗ 0)(−|k|a+δa)dδ e−ω|−|k|a−˜δa|2 e2ω(y−y∗ 0)(−|k|a−˜δa)d˜δ e−ω||k|a+δa|2 e −2ω(y−y∗ 0)(|k|a+δa)dδ, (4.103) by setting ˜δ = −δ. When k > 0 (cid:90) 1 (cid:90) 1 e2ω(y−y∗ e−ω|(k+δ)a|2 0)((k+δ)a)dδ = 2 − 1 2 (4.104) Given k > 0, we add (4.104) for k > 0 and (4.103) for −k < 0, so that we have 2 − 1 2 e−ω|(|k|+δ)a|2 e2ω(y−y∗ 0)((|k|+δ)a)dδ Here we use the Taylor expansion of the exponential function. Furthermore, e2aω(y−y∗ 0)(k+δ) + e 1 + 2ω2(y − y∗ −2aω(y−y∗ 0)(k+δ)(cid:17) 0)2(k + δ)2a2(cid:17) dδ dδ (4.105) 0)2(k + δ)2a2(cid:17) dδ 1 + 2ω2(y − y∗ e2ω2(y−y∗ 0)2((k+δ)a)2 dδ. (4.106) (cid:90) 1 e−ω|ka+δa|2(cid:16) (cid:90) 1 2e−ω|ka+δa|2(cid:16) 2 − 1 2 = 2 − 1 2 (cid:90) 1 2e−ω|ka+δa|2(cid:16) (cid:90) 1 2 − 1 2 2e−ω(k+δ)2a2 ≈ 2 − 1 2 136 (cid:18) (2ω)4(y−y∗ 0)4(δa+ka)4 4! The error term is O (cid:19) (4.102) can be summarized as an universe form, i.e, O . The error of approximations (4.106) and (cid:18) (2ω)4(y−y∗ (cid:19) . 0)4(δa+ka)4 4! Now integral (4.99) becomes 0−y0−ma|2 −ω|y∗ e (cid:90) (cid:90) 1 e−ω||k|a+δa|2 (cid:90) 1 (cid:88) 2 − 1 2 e + 2 − 1 2 k>0 0)(y∗ 0−y0−ma)dm = −2ω(y−y∗ e −2ω(y−y∗ 0)(|k|a+δa)dδ 2e−ω(k+δ)2a2 e2ω2(y−y∗ 0)2((k+δ)a)2 dδ. (4.107) To get the value lk and ωk, we start with k > 0, 0|2(cid:90) 1 −ω|y−y∗ e 2e−ω(k+δ)2a2 e2ω2(y−y∗ 0)2((k+δ)a)2 dδ ≈ −ω|y−y∗ 0|2 2e 2 − 1 2 e2ω2(y−y∗ −ω|y−y∗ 0|2 = lke 0)2(ka)2 e−ω|ka|2 e2ω2(y−y∗ 0)2(ka)2 (4.108) First, the first step above is obtained by choosing δ = 0. Second, the value of lk is defined by Similarly, when k = 0, lk = 2e−ω|ka|2 , lk = 1. 137 (4.109) (4.110) The next step is to compute the value of ωk and define the parameter ω. We start with making the exponent in equation (4.108) negative, −ω + 2ω2k2a2 ≤ 0 ω ≤ 1 2a2k2 , k = 1, 2,··· , q (4.111) and a is previously defined as 1√ I equation (4.97) is in Lemma 4.6.2. If we choose ω = I 3q2 , then ωk in (cid:18)3q2 − 2k2 (cid:19) ωk = 9q4 I. (4.112) After defining the values lk and ωk, we will characterize the size of the error term, (2ω)4(y − y∗ 0)4(δa + ka)4 4! = max k≤q (2ω)4(y − y∗ 0)4(δa + qa)4 4! (cid:19)4(cid:18) 1√ (cid:19)4 I (cid:18) I ≤ 2 3 ≤ 2 243 3q2 1 q4 (qa)4 (4.113) Although this is only the leading order in the series we truncated, the rest of them will be easily controlled from the following two concerns. The first concern is the rest of the series will have high power about q, i.e. is that the coefficient 22n (2n)! will decay exponentially. 1 q2n , for n = 3, 4,··· . The second concern 138 4.8 Numerical Examples In this section, we present all numerical experiments conducted to justify the proposed method through some complex velocity and general initial conditions. All the refer- ence solution in this section are provided by the FDTD method with staggered grid. Reference solution’s correctness will be examined in Appendix B. 4.8.1 Beam Reinitialization 4.8.1.1 SP Reflection V.S. PS Reflection In the subsection, we will provide the numerical results of the analysis in Section 4.7.2 and Section 4.7.3. We firstly conduct the following experiment to illustrate the difference between the SP reflection and the PS reflection. The experiment’s setup for the S-wave is,  f S = (4.114) and − sin(α) sin(36πx + 36π sin(α)(y − 0.25))e sin(36πx + 36π sin(α)(y − 0.25))e 0 x2+(y−0.25)2+(z−0.25)2(cid:17) −36π2(cid:16) x2+(y−0.25)2+(z−0.25)2(cid:17) −36π2(cid:16)  , gS = 0. (4.115) 139 The experiment’s setup for P-wave is,  f P = and  , (4.116) sin(36πx + 36π sin(α)(y − 0.25))e sin(α) sin(36πx + 36π sin(α)(y − 0.25))e −36π2(cid:16) x2+(y−0.25)2+(z−0.25)2(cid:17) x2+(y−0.25)2+(z−0.25)2(cid:17) −36π2(cid:16) 0 gP = 0 The initial width of the beams are all set to be 36π2 and λ = µ = 1. The experiment’s result is displayed in Figure 4.4. The red star line in Figure 4.4 is the ratio sin(θP S ) , sin(α) Figure 4.4: PS Reflection V.S. SP Reflection: Different Ratio Behaviors. PS Reflection(’o’), SP Reflection (’-’) and the blue line is the ratio sin(θSP ) sin(α) . Same as the analysis in Section 4.7.2, the angle for the SP reflection will decrease to zero compared with the original hitting angle. Therefore, the SP reflection wave should be modified to be more focused. 140 00.10.20.30.40.50.600.20.40.60.811.21.4sin(α)The ratio of Sin(θ) Figure 4.5: Sharpened Beams V.S. the Original Beam on Fixed (y, z). Left: y=0.1, z=0.1 Right: y=0.08, z=0.13. Beam Solution after Reinitialization(’o’), Original Beam Solution (’-’). 4.8.1.2 Sharpened Beams V.S. Original Beams The second experiment in this subsection is conducted to show the fact that the pre- cision will not be impacted after adding the new reinitialization (4.97) process. We compare the wavefield after the reinitialization with the original one. Suppose the S-wave hits at the boundary x = 0 with the central point (0, 0.1, 0.1) and the Hessian equals to  36π + 36π2i 12π2i 12π2i 7π2 + 36π2i 0 0 0 0 4π2 + 36π2i  . Its amplitude norm is set as 100. We choose two different sets of the y-value and z-value to show our reinitialization method’s correctness, The wavefield in Figure 4.5 is plotted along the x-axis. As we can see, the reinitialization algorithm won’t affect the accuracy of the result, while the width of each new beam has been decreased. 141 00.10.20.30.4020406080100xu00.10.20.30.4020406080100xu 4.8.2 Periodic Boundary Condition In this subsection, we would like to display some numerical results to the periodic boundary problem. 4.8.2.1 Example 1: The Single Wavepacket in the Constant velocity We first test our algorithm via the constant elastic moduli λ = µ = 1, and the nonzero initial velocity g.  f = sin(36πx + 9πy)e−36π2(cid:0)(x−0.25)2+(z−0.25)2+(z−0.25)2(cid:1) sin(36πx + 9πy)e−36π2(cid:0)(x−0.25)2+(z−0.25)2+(z−0.25)2(cid:1)  , (4.117) 0 and the initial velocity g is g = 0.5f . (4.118) We exhibit our result at the plane z = 0.25. The comparison between two results shows that the propagation dynamics, including the eikonal equation and the transport equation are correct. The Multi-scale Gaussian Wavepacket Transform is also justified, although this is a single wavepacket initial value. We will test the transform further with some more general initial condition. 4.8.2.2 Example 2: General Initial Condition in the Constant velocity The next thing we would like to try is some general initial condition other than the single Gaussian wavepacket to verify our initial decomposition. The initial value is 142 Figure 4.6: Example 1: Single Wavepacket Propagation with the Periodic Boundary Condition. Left: FDTD Solution Right: Gaussian Beam Solution.  f = sin(72πx2)e−36π2(cid:0)(x−0.25)2+(z−0.25)2+(z−0.25)2(cid:1) sin(72πx2)e−36π2(cid:0)(x−0.25)2+(z−0.25)2+(z−0.25)2(cid:1) 0  (4.119) defined as and g = 0; (4.120) Let’s compare the result along the x-axis by setting z = 0.25 and y = 6 32. Now let’s compare 2-D wavefields at z = 0.25. For other more complex velocity, we will exhibit those results in the Dirichlet boundary condition, since the periodic boundary problem essentially tests the correctness of propagation dynamic and the decomposition process of the initial condition as proposed 143 xy 00.20.400.10.20.30.40.5−0.4−0.200.20.4xy 00.20.400.10.20.30.40.5−0.3−0.2−0.100.10.20.3 Figure 4.7: Example 2: General Initial Value Propagation with Periodic Boundary Condition along x-axis FDTD Solution(’o’), Gaussian Beam Solution (’-’) Figure 4.8: Example 2: General Initial Value Propagation with Periodic Boundary Condition. Left: FDTD Solution Right: Gaussian Beam Solution. in Section 4.4. This can be examined in the Dirichlet boundary problem as well. 144 00.10.20.30.40.5−0.2−0.15−0.1−0.0500.050.10.150.2xUxy 00.20.400.10.20.30.40.5−0.3−0.2−0.100.10.20.3xy 00.10.20.30.40.500.10.20.30.40.5−0.3−0.2−0.100.10.20.3 4.8.3 Reflection: P-wave We have so far used the periodic boundary condition to justify our decomposition process and the dynamic system, now we would like to test our reflection scheme on the pure P-wave initial condition. In this subsection, we fix our initial condition as the following and test our algorithms over the different velocity. −36π2(cid:16) −36π2(cid:16) (x−0.15)2+(y−0.25)2+(z−0.25)2(cid:17) (x−0.15)2+(y−0.25)2+(z−0.25)2(cid:17) sin(32πx + 8πy)e f = 1 4 sin(32πx + 8πy)e   (4.121) (4.122)   and the initial velocity g is −36π2(cid:16) −36π2(cid:16) (x−0.15)2+(y−0.25)2+(z−0.25)2(cid:17) (x−0.15)2+(y−0.25)2+(z−0.25)2(cid:17) 2 sin(32πx + 8πy)e g = 1 2 sin(32πx + 8πy)e 0 0 4.8.3.1 Example 3: P-wave Reflection in the Constant velocity Our first setting is still under the constant elastic moduli, λ = 1 and µ = 2. We compare the wavefield at z = 0.25 and T = 0.14, after the primary reflection happens. The P-wave reflection dynamics can be justified after this numerical experiment. Here the mesh size of the FDTD method employed is 1 640 and this scale will be used for all the FDTD results in the rest of the paper. 145 Figure 4.9: Example 3: P-wave Reflection in Constant velocity with Dirichlet Boundary Condition. Left: FDTD Solution Right: Gaussian Beam Solution. 4.8.3.2 Example 4: P-wave Reflection in the Linear velocity The second velocity in the P-wave reflection section we use is µ = 2 and λ = 1 + 0.2x. We first compare the result at the fixed (y, z) = (0.125, 0.25) along x-axis. Then we set z fixed as 0.25, The P-wave reflection dynamics in some general velocity has been justified further. As we can see in Figure 4.10, the Gaussian beam solution performs quite well in the major region, and the tolerable error shows up near the boundary. 4.8.3.3 Example 5: P-wave Reflection in the Sinusoidal velocity Now let’s try the sinusoidal elastic moduli λ = 1 + sin(4πx), in which there will be some caustics points. Again, we compare the result along the x-axis first by fixing y = z = 0.25. 146 xy 00.20.400.10.20.30.40.5−0.2−0.100.10.2xy 00.20.400.10.20.30.40.5−0.2−0.100.10.2 Figure 4.10: Example 4: P-wave Reflection in the Linear velocity with Dirichlet Boundary Condition. FDTD Solution (’o’), Beam Solution(’-’) Then we fix z = 0.25, To remark, the reinitialization scheme is not involved in all the P-wave reflection results shown above. The small error implies that the P-wave reflec- tion does not require the reinitialization, while it is necessary in the S-wave reflection. 4.8.4 Reflection S-wave In this subsection, we will test the reinitialization process, but at first we will justify the necessity of adding reinitialization process. 4.8.4.1 Example 6: S-wave Reflection with Orthogonal Hitting Angle Firstly, we see the S-wave reflection with orthogonal hitting angle, that is sin(α) = 0 in Figure 4.4. It is displayed by Figure 4.14 showing that the original method without the reinitialization is good enough. 147 00.10.20.30.40.5−0.15−0.1−0.0500.050.10.150.2xy Figure 4.11: Example 4: P-wave Reflection in Linear velocity with Dirichlet Boundary Condition. Left: FDTD Solution Right: Gaussian Beam Solution. Figure 4.12: Example 5: P-wave Reflection in Sinusoidal velocity along x-axis. FDTD Solution(’o’), Beam Solution (’-’) The setup of our experiment is  f = sin(36πy)e−36π2((x−0.25)2+(y−0.15)2+(z−0.25)2) 0 0 148  . (4.123) xy 00.20.400.10.20.30.40.5−0.2−0.100.10.2xy 00.20.400.10.20.30.40.5−0.15−0.1−0.0500.050.10.1500.10.20.30.40.5−0.2−0.15−0.1−0.0500.050.10.150.2xu Figure 4.13: Example 5:P-wave Reflection in Sinusoidal velocity with Dirichlet Boundary Condition. Left: FDTD Solution, Right: Gaussian beam Solution and there is no initial velocity, i.e. g = 0. Two velocity parameters λ and µ here are both constants, λ = 1; µ = 2; Figure 4.14 compares the wave field generated by our method to the one from FDTD method at z = 0.25, x = 0.25 and T = 0.2. Now we fix z = 0.25 and y = 5 32, As we can see from Figure 4.4, when sin(α) = 0, the regular reflection method is expected to be well enough and the experiment result above justifies our conclusion. 4.8.4.2 Example 7: S-wave Reflection with Non-Orthogonal Hitting Angle Now if we change the ray direction to increase the width of the SP-wave, we will see that the regular reflection dynamics fails in this case. To make that happen, we specify 149 xy 00.10.20.30.40.500.10.20.30.40.5−0.2−0.15−0.1−0.0500.050.10.150.2xy 00.10.20.30.40.500.10.20.30.40.5−0.15−0.1−0.0500.050.10.150.2 Figure 4.14: Example 6: S-wave Reflection with Orthogonal hitting Angle along y-axis FDTD Solution (’o’), GB Solution (’-’). Figure 4.15: Example 6: S-wave Reflection with Orthogonal hitting Angle along x-axis. FDTD Solution (’o’), GB Solution (’-’).  −36π2(cid:16) −36π2(cid:16) (x−0.25)2+(y−0.15)2+(z−0.25)2(cid:17) (x−0.25)2+(y−0.15)2+(z−0.25)2(cid:17)  . (4.124) our initial value as, 2 sin(36πy + 18πx)e − sin(36πy + 18πx)e f = 0 150 00.10.20.30.40.5−0.25−0.2−0.15−0.1−0.0500.050.10.150.20.25yu00.10.20.30.40.5−0.0200.020.040.060.080.1xu Figure 4.16: Example 7: S-wave Reflection with Non-orthogonal Hitting Angle. Left: Gaussian Beam Solution without reinitialization, Right: Gaussian Beam Solution with reinitialization. FDTD Solution (’o’), GB Solution (’-’) We first see that the result generated without the extra reinitialization. The above left Figure 4.16 is plotted along the x-axis with z = 0.25, y = 0.125. As we mentioned before, although the main pattern is captured with good accuracy, the tail region of the Gaussian beam wavefield is not clean enough due to the fact that the SP-wave is involved. After adding the reinitialization, in the above right Figure 4.16, the beam solution with the reinitialization shows the better result in the tail region without hurting the accuracy of other parts. The experiment shown in Figure 4.16 illustrates that the analysis in Section 4.7 is correct. The SP-wave reflection will lose the accuracy to some degree such that adding the reinitialization scheme is necessary. 151 00.10.20.30.40.5−0.4−0.200.20.4xU00.10.20.30.40.5−0.4−0.200.20.4xU 4.8.4.3 Example 8: S-wave Reflection: Linear velocity Let’s see the comparison under some more complicated elastic moduli µ = 2 + 0.2y and λ = 1. We compare the wavefield without the reinitialization to the FDTD result first at z = 0.25. Figure 4.17: Example 8: S-wave Reflection in Linear velocity without Reinitialization. Left: FDTD Solution Right: Gaussian Beam Solution without Reinitialization. As we can see in the upper left corner of Figure 4.17, there is some significant per- turbations in the FDTD result, while the beam method without the reinitialization, which is shown in Figure 4.17, is not able to cover that part. Now we will see the result after using the reinitialization process. As we can see from the right Figure 4.20, the missing part in the upper left region is covered by the reinitialization algorithm. 152 xy 00.20.400.10.20.30.40.5−0.4−0.200.20.4xy 00.20.400.10.20.30.40.5−0.4−0.200.20.4 Figure 4.18: Example 8: S-wave Reflection in Linear velocity with Reinitialization. Left: FDTD Solution Right: Gaussian Beam Solution with Reinitialization. 4.8.4.4 Example 9: S-wave Reflection: Sinusoidal velocity In this example, we set the elastic moduli µ = 2 + 0.2 sin(x) to make the Hessian of the velocity nonzero. And all the other components remain the same as the last one, including the initial value, λ and the terminal time T . Compare two wavefields on the plane z = 0.25, In this more complex velocity, the advantage of the Gaussian beam ansatz has shown up as the caustics problem is resolved automatically. There will be caustics in this sinusoidal velocity as the eikonal equation will be multivalued in some region. As Figure 4.20 suggests, the beam solution will perform well even when the caustics shows up. 153 xy 00.20.400.10.20.30.40.5−0.4−0.200.20.4xy 00.20.400.10.20.30.40.5−0.4−0.200.20.4 Figure 4.19: Example 9: S-wave Reflection in Sinusoidal velocity. GB Solution(’o’), FDTD Solution(’-’) Figure 4.20: Example 9: S-wave Reflection in Sinusoidal velocity with Reinitialization. Left: FDTD Solution Right: Gaussian Beam Solution with Reinitialization. 4.8.5 General Initial Condition Now after showing the effect of the reinitialization, especially after comparing it to the result without the reinitialization, it is confident to say that the proposed reinitializa- 154 00.10.20.30.40.5−0.15−0.1−0.0500.050.10.150.2xyxy 00.20.400.10.20.30.40.5−0.15−0.1−0.0500.050.10.15xy 00.20.400.10.20.30.40.5−0.15−0.1−0.0500.050.10.15 tion algorithm is correct and necessary in this problem. We will end our numerical tests with some more general initial conditions. 4.8.5.1 Example 10: The General Initial Condition in The Sinusoidal ve- locity We test Multiscale Gaussian Wavepacket method via the general initial condition and sinusoidal elastic moduli λ = 1 + sin(4πx), µ = 2. The initial condition is −36π2(cid:16) (x−0.15)2+(y−0.25)2+(z−0.25)2(cid:17) −36π2(cid:16) (x−0.15)2+(y−0.25)2+(z−0.25)2(cid:17) (x−0.15)2+(y−0.25)2+(z−0.25)2(cid:17) −36π2(cid:16)  . (4.125)  sin(32πx + 32πy2)e f = sin(32πx + 10πy)e sin(16πz)e Here we let our wavefield propagate until T = 0.2. We first display the first component in the resulting wavefield. Let’s see the second component of the resulting wavefield, and see the comparison between two methods along the x-axis by fixing y = z = 0.25. With more general initial value, we include P and S-wave at the same time, meanwhile, different types of the reflection happen simultaneously, so as the different reflection modes. Remark 4.8.1. Compared with the FDTD algorithm with parallel computing scheme, our asymptotic algorithm has the larger time complexity. However, as the wavenumber of the initial value is increasing, the FDTD scheme requires finer grid size, leading to the requirement of some larger storage. This is impossible in the current GPU units, while our method’s storage complexity is independent of the wavenumber as well as the time complexity. 155 Figure 4.21: Example 10: General Initial Condition Propagation with Reinitialization (First Component). Left: FDTD Solution Right: Gaussian Beam Solution with Reinitial- ization. Figure 4.22: Example 10: General Initial Condition Propagation with Reinitialization (Second Component). Left: FDTD Solution Right: Gaussian Beam Solution with Reinitialization. 156 xy 00.20.400.10.20.30.40.5−0.2−0.100.10.2xy 00.20.400.10.20.30.40.5−0.2−0.100.10.2xy 00.20.400.10.20.30.40.5−0.0500.05xy 00.20.400.10.20.30.40.5−0.0500.05 Figure 4.23: Example 10: General Initial Condition Propagation with Reinitialization (Second Component). FDTD Solution (’o’), GB Solution(’-’). 4.8.5.2 Convergence Rate Analysis In the end, we propose the convergence rate analysis. The initial condition is set as −42π2(cid:16) −42π2(cid:16) (x−0.25)2+(y−0.15)2+(z−0.25)2(cid:17) (x−0.25)2+(y−0.15)2+(z−0.25)2(cid:17)  (4.126)  sin(η(16πx + 8πy))e sin(η(16πx + 8πy))e f = 0 The velocity here are all constants, λ = 1; µ = 2; The amplifying factor η is a geometric series, 1, 1.5, 2.25,··· , 1.55. The blue star line is the logarithm of the L2 norm of the error at different η, while the red line is the linear function with the slope as 1 2 log(1.5). It is well known that the convergence rate of the Gaussian beam is 1√ ω , and as proved in the paper [5], the 157 00.10.20.30.40.5−0.15−0.1−0.0500.050.10.150.2xu Figure 4.24: Log-log plot: Convergence Rate of the New Gaussian Beam Method GB Method Error Curve(’-*’), Line with the slope = 1 2 log(1.5)(’-’) Multiscale Gaussian wavepacket transform also follows O( 1√ ω ). This pattern can be seen in Figure 4.24. 158 3.544.555.56−5.5−5−4.5−4log(wavenumber)log(error) Chapter 5 Conclusion We propose two methods based on multi-scale Gaussian beam method in this thesis. The first one is solving the elastic wave propagation in the bounded domain and the second one is for the prestack inversion process, which is an inverse problem in geo- physical applications. These two methods are both based on the multiscale Gaussian beam method described in Chapter 2. Therefore, both methods are capable of re- solving caustics problem automatically and they both take advantage of the parabolic scaling principle for efficiency. In the first part (Chapter 3), we present a new prestack inversion process, which con- nects the boundary data to the wavefront set of the perturbation. We first modify the multiscale Gaussian wavepacket transform [48] appropriately to suit to the imag- ing operator. Secondly, the multi-valued traveltime information is preserved due to the Gaussian beam function. This improves the quality of resulting migration image. Another big advantage of our multiscale Gaussian beam inversion method is its ro- bustness to the polluted data. Since we recover the reflector by its wavefront set, the noise in the boundary data, which is far from the target frequency, won’t affect the imaging result. Lastly, our imaging condition is performed in the time domain to avoid the extra Fourier transform on the data set. This feature makes our algorithm more 159 applicable considering the large size of the trace dataset in the real world. In the second part (Chapter 4), we present a novel Multiscale Gaussian beam method to solve the elastic wave equation in the bounded domain. Firstly, a new vector-valued wavepacket transform is developed to adapt to the highly oscillated vector-valued initial condition following the parabolic scaling principle. Secondly, a novel reinitialization strategy is added in the process to improve the efficiency and accuracy. There are several advantages about this new reinitialization method. The first one is that the new reinitialization is applied to the single wavepacket instead of the whole wavefield. This will improve the efficiency greatly. The second one is that the center of each new wavepacket after the reinitialization is same to the center before, which guarantees all computation happening inside the domain without extra assumption outside the domain. Although the typical FDTD (Finite Difference Time Domain) method is faster by implementing in parallel, the requirement of large storage will still make FDTD method unfeasible in the high frequency regime. 160 APPENDICES 161 Appendix A Proof in Inverse Process Hessian Matrix in Corollary 3.3.1 In Corollary 3.3.1, the quadratic term ||r − ˆr0||2 in the phase function is not necessary to be a Gaussian profile, since ˆM is only the Schur complement of a semi ˆM(ˆtc) positive definite matrix. Lemma A.0.1. For any boundary points r = {x = (x1,··· , xd) : xd = 0}, we have (cid:32) Im( ˆM )(ˆtc) − Im(ˆτtx)(ˆtc)Im(ˆτtx)T (ˆtc) (cid:33) r ≥ arT Im( ˆM )(ˆtc)r, (A.1) rT Im(ˆτtt)(ˆtc) for some constant a, if there is no grazing ray, |ˆpd(ˆtc)| ≥ b||ˆp(ˆtc)||, (A.2) where b is an universal lower bound. By equation (3.60) and (3.61), we have Im(ˆτtx)(ˆtc; y, p) = ∓v(ˆy(ˆtc))Im( ˆM )(ˆtc) ˆp(ˆtc) ||ˆp(ˆtc)|| , (A.3) 162 and Im(ˆτtt)(ˆtc; y, p) = v2(ˆy(ˆtc)) ˆpT (ˆtc) ||ˆp(ˆtc)|| Im( ˆM )(ˆtc) ˆp(ˆtc) ||ˆp(ˆtc)||. (A.4) From now on, we let all functions be defined at t = ˆtc without writing it out explicitly. Then, rT Im(ˆτtx)Im(ˆτtx)T Im(ˆτtt) r = rT Im( ˆM )ˆpˆpT Im( ˆM ) ˆpT Im( ˆM )ˆp r. We introduce the following notation, (cid:48) p = ˆp(cid:113) ˆpT Im( ˆM )ˆp . (A.5) (A.6) Then equation (A.1) can be translated to the following optimization problem. We can instead prove that the optimal value of the following optimization problem with fixed (cid:48) p is a, min r subject to 1 −(cid:16) (cid:48)(cid:17)2 , rT Im( ˆM )p rT Im( ˆM )r = 1, rT ed = 0, (A.7) (A.8) where ed = (0,··· , 0, 1)T . It is equivalent to, min r subject to 1 − rT Im( ˆM )p (cid:48) , rT Im( ˆM )r = 1, rT ed = 0, since the optimization problem (A.7) is the square term of problem (A.8). Using 163 Lagrange multipliers, L(r; λ1, λ2) = 1 − rT Im( ˆM )p (cid:48) + λ1rT Im( ˆM )r + λ2rT ed. (A.9) Differentiate L with respect to r, −Im( ˆM )p (cid:48) + 2λ1Im( ˆM )r + λ2ed = 0, (A.10) then 2λ1r = −λ2Im( ˆM )−1ed + p (cid:48) . By the second restriction rT ed = 0, (cid:48) eT d p d Im( ˆM )−1eT eT d λ2 = d Im( ˆM )−1eT d . Therefore, (cid:113) We denote A = ˆpT Im( ˆM )ˆp and D = eT and the optimizer r(cid:63) satisfies, (cid:48) eT d p AD . λ2 = r(cid:63) (cid:107) r0, where r0 is defined as, r0 ≡ p (cid:48) (cid:48) − eT d p AD Im( ˆM )−1ed. 164 (A.11) (A.12) (A.13) (A.14) (A.15) By restriction rT Im( ˆM )r = 1, we have, (cid:113) Equivalently, 2λ1 = (cid:113) r(cid:63) = r0 0 Im( ˆM )r0 rT , (A.16) 0 Im( ˆM )r0. Equation (A.7) then becomes, rT 1 −(cid:16) min r (cid:48)(cid:17)2 = 1 − rT Im( ˆM )p  0(cid:113) rT 0 Im( ˆM )r0 rT (cid:48) Im( ˆM )p 2 . (A.17) The first term needed to be evaluated above is, rT 0 Im( ˆM )p (cid:48) = (p = (p (cid:48) − eT d p AD (cid:48) )T Im( ˆM )p (cid:48) Im( ˆM )−1ed)T Im( ˆM )p (cid:48) (cid:48) (cid:48) − (eT )2 d p A2D . (cid:48) (cid:48) − eT d p AD Im( ˆM )−1ed) d Im( ˆM )−1ed − 2eT d p AD eT d p (cid:48) (cid:48) The second term is rT 0 Im( ˆM )r0 = (p = (p = (p = (p = (p (cid:48) (cid:48) − eT d p AD (cid:48) )T Im( ˆM )p (cid:48) (cid:48) (cid:48) )T Im( ˆM )p )T Im( ˆM )p )T Im( ˆM )p Im( ˆM )−1ed)T Im( ˆM )(p (cid:48) (cid:48) + + (cid:48) (cid:48) (cid:48) )2 )2 (eT d p (AD)2 eT (AD)2 D − 2eT (eT d p d p AD (cid:48) − 2(eT d p A2D (cid:48) )2 (cid:48) + (eT )2 d p A2D (cid:48) (cid:48) − (eT )2 d p A2D (cid:48) eT d p A . (A.18) 165  1 − 0(cid:113) rT 0 Im( ˆM )r0 rT 2 (cid:48) Im( ˆM )p = 1 − = 1 −  rT (cid:113) (cid:32) (cid:48) 2 (cid:48) 0 Im( ˆM )p 0 Im( ˆM )r0 rT (cid:33)2 (p )T Im( ˆM )p (cid:48) (p )T Im( ˆM )p (cid:48) (cid:48) − (eT )2 d p A2D (cid:48) (cid:48) − (eT )2 p d A2D (cid:48) = (eT )2 d p A2D . (A.19) To summarize, (cid:32) rT and Im( ˆM ) − (r − ˆr0)T Im(ˆτtx)Im(ˆτtx)T Im(ˆτtt) (cid:33) r ≥ arT Im( ˆM )(ˆtc)r (A.20) a = b2 A2D = b2 d (Im( ˆM ))−1ed) (ˆpT Im( ˆM )ˆp)(eT . (A.21) Proof of Proposition 3.3.5 Proposition A.0.1. Consider two scattering beams (ˆy(t), ˆp(t), ˆM (t), ˆA(t)) and (ˆx(t), ˆξ(t), ˆN (t), ˆC(t)), and assume that there exists significant interaction effects be- tween these two beams. There exists two constants C∗ 2 related to the background 1 and C∗ velocity, such that d|| ˆM (t) − ˆN (t)|| dt ≤ C∗ 1 (cid:112)||p|| + C∗ 2|| ˆM (t) − ˆN (t)||. (A.22) where || ˆM (t) − ˆN (t)|| is defined as the matrix norm induced by the vector 2-norm. 166 Proof. Both ˆM (t) and ˆN (t) satisfy d ˆM dt d ˆN dt = −Gxx(ˆy(t), ˆp(t)) − ˆM Gxp(ˆy(t), ˆp(t)) − GT = −Gxx(ˆx(t), ˆξ(t)) − ˆN Gxp(ˆx(t), ˆξ(t)) − GT xp(ˆy(t), ˆp(t)) ˆM − ˆM Gpp(ˆy(t), ˆp(t)) ˆM , xp(ˆx(t), ˆξ(t)) ˆN − ˆN Gpp(ˆx(t), ˆξ(t)) ˆN . (A.23) Take the difference between above equations, we have − d( ˆM (t) − ˆN (t)) (cid:16) Gxx(ˆy(t), ˆp(t)) − Gxx(ˆx(t), ˆξ(t)) dt = (cid:16) (cid:17) (cid:16) (cid:17) ˆN (t) + 2 ˆN (t) Gxp(ˆy(t), ˆp(t)) − Gxp(ˆx(t), ˆξ(t)) (cid:17) Gpp(ˆy(t), ˆp(t)) − Gpp(ˆx(t), ˆξ(t)) + ˆN (t) + 2( ˆM (t) − ˆN (t))Gxp(ˆy(t), ˆp(t)) + ( ˆM (t) − ˆN (t))Gpp(ˆy(t), ˆp(t))( ˆM (t) + ˆN (t)). (A.24) The first term in equation (A.24), ||Gxx(ˆy(t), ˆp(t))−Gxx(ˆx(t), ˆξ(t))|| = (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)∇∇v(ˆy(t))||ˆp(t)|| − ∇∇v(ˆx(t))|| ˆξ(t)||(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ C3(||ˆy(t) − ˆx(t)||)||ˆp(t)|| + C2(||ˆp(t) − ˆξ(t)||), (A.25) where C3 is the maximum value of the third order derivative of the velocity v and C2 maximum value of the second order derivative of the velocity v. According to Lemma 3.3.2, we have ||Gxx(ˆy(t), ˆp(t)) − Gxx(ˆx(t), ˆξ(t))|| ∼ O((cid:112)||ˆp(t)||). (A.26) 167 The second term in equation (A.24), 2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ˆN (t) (cid:16) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ˆN (t) = 2 ≤ 2|| ˆN (t)|| Gxp(ˆy(t), ˆp(t)) − Gxp(ˆx(t), ˆξ(t)) (cid:18) ˆp(t) ∇v(ˆy(t)) (cid:32) C2||ˆy(t) − ˆx(t)|| + C1 ||ˆp(t)|| (cid:17)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:32) ˆξ(t) (cid:19)T − ∇v(ˆx(t)) (cid:33)(cid:33) (cid:32)||ˆΞ(t)|| || ˆξ(t)|| ||ˆp(t)|| , (cid:33)T(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (A.27) (A.28) (cid:33)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ˆN (t) ˆN (t) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:33)T ˆN (t) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12). (A.29) where C1 is the largest value of ||∇v||. Therefore, (cid:16) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ˆN (t) 2 Gxp(ˆy(t), ˆp(t)) − Gxp(ˆx(t), ˆξ(t)) (cid:17)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∼ O((cid:112)||ˆp(t)||). The third term in equation (A.24), (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ˆN (t) (cid:16) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ˆN (t) + Gpp(ˆy(t), ˆp(t)) − Gpp(ˆx(t), ˆξ(t)) v(ˆy(t)) ||ˆp(t)|| (cid:18) ˆp(t) ||ˆp(t)|| (cid:19)(cid:18) ˆp(t) ||ˆp(t)|| (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:32) (cid:17) ˆN (t) ||ˆp(t)|| − v(ˆx(t)) (cid:33)(cid:32) ˆξ(t) (cid:32) ˆξ(t) (cid:19)T − v(ˆx(t)) || ˆξ(t)|| v(ˆy(t)) || ˆξ(t)|| || ˆξ(t)|| || ˆξ(t)|| First, we have (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)v(ˆy(t)) (cid:32) ||ˆp(t)|| − v(ˆx(t)) || ˆξ(t)|| ||ˆy(t) − ˆx(t)|| C1 ||ˆp(t)|| (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ˆN (t) ˆN (t) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ + v(ˆy(t)) ||ˆp(t) − ˆξ(t)|| ||ˆp(t)||2 (cid:33) || ˆN (t) ˆN (t)|| ∼ O((cid:112)||ˆp(t)||). (A.30) 168 Second, (cid:18) ˆp(t) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) v(ˆy(t)) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)||ˆy(t) − ˆx(t)|| ||ˆp(t)|| ||ˆp(t)|| ||ˆp(t)|| C1 (cid:19)(cid:18) ˆp(t) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) + C1 (cid:32) ˆξ(t) (cid:19)T − v(ˆx(t)) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) v(ˆy(t))(1 − ˆκ(t))||ˆp(t)||2 ||ˆp(t)|| || ˆξ(t)|| || ˆξ(t)|| ||ˆp(t)||3 (cid:33)(cid:32) ˆξ(t) (cid:33)T(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ O(||ˆp(t)||−3/2). || ˆξ(t)|| To summarize, (cid:16) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ˆN (t) Gpp(ˆy(t), ˆp(t)) − Gpp(ˆx(t), ˆξ(t)) (cid:17) ˆN (t) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∼ O((cid:112)||ˆp(t)||). (A.31) (A.32) Insert these asymptotic analysis equations (A.26), (A.28), (A.32) into equation (A.24), we have = O((cid:112)||ˆp(t)||) + 2( ˆM (t) − ˆN (t))Gxp(ˆy(t), ˆp(t)) dt − d( ˆM (t) − ˆN (t)) + ( ˆM (t) − ˆN (t))Gpp(ˆy(t), ˆp(t))( ˆM (t) + ˆN (t)) = O((cid:112)||ˆp(t)||) + ( ˆM (t) − ˆN (t)) (cid:16) (cid:17) 2Gxp(ˆy(t), ˆp(t)) + Gpp(ˆy(t), ˆp(t))( ˆM (t) + ˆN (t)) . The coefficient matrix in front of ˆM (t) − ˆN (t) satisfies, ||2Gxp(ˆy(t), ˆp(t)) + Gpp(ˆy(t), ˆp(t))( ˆM (t) + ˆN (t))|| ≤ 2||∇v(ˆy(t))|| + 2 ≤ O(1). (A.33) || ˆM (t) + ˆN (t)|| ||ˆp(t)|| (A.34) Then, ˆM (t) − ˆN (t) satisfies, d|| ˆM (t) − ˆN (t)|| dt ≤ C∗ 1 (cid:112)||p|| + C∗ 2|| ˆM (t) − ˆN (t)||, || ˆM (0) − ˆN (0)|| = 0. (A.35) 169 Proof of Lemma 3.3.5 Lemma A.0.2. Consider two scattering beams (ˆy(t), ˆp(t), ˆM (t), ˆA(t)) and (ˆx(t), ˆξ(t), ˆN (t), ˆC(t)), and there exists significant interaction effects between these two beams. Suppose the function g(t) is then we have and g(t) = ˆp(t) · (ˆy(t) − ˆx(t)), (A.36) g(t) = g(0) + O(1), (A.37) (cid:32) (cid:33) . ||ˆΞ(t)||2 ˆκ(t)2||ˆp(t)|| 1 2 (cid:48) g (t) = v(ˆx(t)) (A.38) Proof. We start with calculating the derivative dg(t) dt be positive, i.e. G(x, p) = v(x)||p||. Negative branch will be the same. and assume the Hamiltonian to (cid:18) dˆy(t) dt (cid:19) − dˆx(t) dt (cid:32) v(ˆy(t)) ˆp(t) ||ˆp(t)|| − v(ˆx(t)) ˆξ(t) || ˆξ(t)|| (A.39) (cid:33) . dg dt = dˆp(t) dt · (ˆy(t) − ˆx(t)) + ˆp(t) · = −∇v(ˆy(t)) · (ˆy(t) − ˆx(t))||ˆp(t)|| + ˆp(t) · 170 Use the Taylor expansion of the velocity v at the point ˆy(t), = ||ˆp(t)|| (v(ˆx(t)) − v(ˆy(t))) + ||ˆp(t)||v(ˆy(t)) − v(ˆx(t)) dg dt ˆp(t) · ˆξ(t) || ˆξ(t)|| + O(||ˆx(t) − ˆy(t)||2) (cid:32) = ||ˆp(t)||v(ˆx(t)) − v(ˆx(t)) = v(ˆx(t)) ||ˆp(t)|| − ˆp(t) · ˆξ(t) || ˆξ(t)|| ˆp(t) · ˆξ(t) (cid:33) || ˆξ(t)|| + O(||ˆx(t) − ˆy(t)||2) + O(||ˆx(t) − ˆy(t)||2). Substitute decomposition (3.56) into the fraction term ˆp(t)· ˆξ(t) || ˆξ(t)|| . Since ˆΞ(t) is orthogonal to ˆp(t), ˆp(t) · ˆξ(t) || ˆξ(t)|| = (cid:113) ˆκ(t)||ˆp(t)||2 ˆκ(t)2||ˆp(t)||2 + ||ˆΞ(t)||2 ˆκ(t)||ˆp(t)||2 ˆκ(t)||ˆp(t)|| = = ||ˆp(t)|| − 1 2 1 ||ˆΞ(t)||2 ||ˆκ(t)ˆp(t)||2 1 + 1 2 ||ˆΞ(t)||2 |ˆκ(t)|2||ˆp(t)||. Here we use the Taylor expansion of the square root function and Geometric series to approximate. Then dg dt = v(ˆx(t)) = O(1). (cid:18) ||ˆp(t)|| − ˆκ(t)||ˆp(t)||2 ˆκ(t)||ˆp(t)|| (cid:19) + v(ˆx(t)) 2 ||ˆΞ(t)||2 |ˆκ(t)|2||ˆp(t)|| (A.40) Naturally, after finite time, g(t) = g(0) + O(1). (A.41) 171 Proof of Lemma 3.3.6 Lemma A.0.3. Consider two scattering beams (ˆy(t), ˆp(t), ˆM (t), ˆA(t)) and (ˆx(t), ˆξ(t), ˆN (t), ˆC(t)), and there exists significant interaction effects between these two beams. Suppose the pure imaginary matrix ˆM (0) has a symmetric positive definite imaginary part and is the initial condition of the Hessian for the beam, then (y − x)T ˆM (0)(y − x) = (ˆy(t) − ˆx(t))T ˆM (t)(ˆy(t) − ˆx(t)) + O(1). (A.42) Proof. We denote the function g(t) as g(t) = ||ˆy(t) − ˆx(t)||2 ˆM (t) Throughout this proof, we use the Hamiltonian G(x, p) = v(x)||p|| and the negative Hamiltonian will be the same. The derivative of g is, dg(t) = 2(Gp(ˆy(t), ˆp(t)) − Gp(ˆx(t), ˆξ(t)))T ˆM (t)(ˆy(t) − ˆx(t)) Gpx(ˆy(t), ˆp(t))(ˆy(t) − ˆx(t)) + Gpp(ˆy(t), ˆp(t))(ˆp(t) − ˆξ(t)) dt + (ˆy(t) − ˆx(t))T d ˆM (t) (cid:16) dt = 2 + (ˆy(t) − ˆx(t))T d ˆM (t) dt (ˆy(t) − ˆx(t)) (ˆy(t) − ˆx(t)). (cid:17)T ˆM (t)(ˆy(t) − ˆx(t)) (A.43) From Riccati equation (2.11), we have dM dt = −Gxx − M Gxp − GT xpM − M GppM. 172 Insert the above equation into equation (A.43) and abbreviate G = G(ˆy(t), ˆp(t)), if there’s no variable specified, (cid:16) ˆM (t)(ˆy(t) − ˆx(t)) (cid:17) Gpp dg(t) = −(ˆy(t) − ˆx(t))T Gxx(ˆy(t) − ˆx(t)) − 2(ˆy(t) − ˆx(t))T ˆM (t)Gxp(ˆy(t) − ˆx(t)) dt + 2(ˆy(t) − ˆx(t))T ˆM (t)Gpx(ˆy(t) − ˆx(t)) + 2(ˆp(t) − ˆξ(t))T Gpp ˆM (t)(ˆy(t) − ˆx(t)) −(cid:16) ˆM (t)(ˆy(t) − ˆx(t)) (cid:17)T −(cid:16) ˆM (t)(ˆy(t) − ˆx(t)) (cid:17)T (cid:16) ˆM (t)(ˆy(t) − ˆx(t)) (cid:17) = −(ˆy(t) − ˆx(t))T Gxx(ˆy(t) − ˆx(t)) + 2(ˆp(t) − ˆξ(t))T Gpp ˆM (t)(ˆy(t) − ˆx(t)) Gpp = −(ˆy(t) − ˆx(t))T Gxx(ˆy(t) − ˆx(t)) − || ˆM (t)(ˆy(t) − ˆx(t)) − (ˆp(t) − ˆξ(t))||2 + ||ˆp(t) − ˆξ(t)||2 Gpp Gpp. (A.44) The first term in equation (A.44) satisfies, (ˆy(t) − ˆx(t))T (Gxx(ˆy(t), ˆp(t)))(ˆy(t) − ˆx(t)) ∼ O(1). (A.45) First, Gxx(ˆy(t), ˆp(t)) ∼ O(||ˆp(t)||), since v is smooth. On the other hand, ˆy(t) − ˆx(t)’s order is O(1/(cid:112)||ˆp(t)||) by Lemma 3.3.2. The second term in equation (A.44) satisfies, || ˆM (t)(ˆy(t) − ˆx(t)) − (ˆp(t) − ˆξ(t))||2 Gpp(ˆy(t),ˆp(t)) ∼ O(||ˆp(t)||1/2||ˆp(t)||−1||ˆp(t)||1/2). (A.46) The term Gpp(ˆy(t), ˆp(t)) is of the order O(1/||ˆp(t)||) implied by Gpp’s expression. Again, Lemma 3.3.2 demonstrates that ˆM (t)(ˆy(t)− ˆx(t))−(ˆp(t)− ˆξ(t)) ∼ O((cid:112)||ˆp(t)||). 173 The third term in equation (A.44) satisfies, ||ˆp(t) − ˆξ(t)||2 Gpp(ˆy(t),ˆp(t)) ∼ O(||ˆp(t)||1/2||ˆp(t)||−1||ˆp(t)||1/2) ∼ O(1). (A.47) It can be justified by combining Gpp’s expression and the fact that ||ˆp(t) − ˆξ(t)|| ∼ O((cid:112)||ˆp(t)||). Proof of Proposition 3.3.3 Proposition A.0.2. iˆγ(ˆt0(x, ξ); x, ξ)|ω − ˆτt(ˆt0(x, ξ); x, ξ) − ˆϑ(ˆt0(x, ξ); x, ξ)T (r − ˆx(ˆt0(x, ξ)))|2 ≈ iˆγ(ˆt0(y, p); x, ξ)|ω − ˆτt(ˆt0(y, p); x, ξ) −(cid:16) ˆϑ(ˆt0(y, p); x, ξ)) (cid:17)T (cid:18) 1 (cid:19) (r − ˆx(ˆt0(y, p)))|2 (A.48) + O ||p|| . Proof. The coefficient ˆγ satisfies, ˆγ(ˆt0(x, ξ); x, ξ) = − 1 2ˆτtt(ˆt0(x, ξ); x, ξ) . By Assumption 3.2.3, it is safe to say that the background velocity v around ˆx(ˆt0(y, p)) is a constant function. Consequently, we have Hamiltonian satisfying Gx = 0, and ˆGp(ˆx(ˆt0(x, ξ); x, ξ), ˆξ(ˆt0(x, ξ); x, ξ)) = ±v(ˆx(ˆt0(y, p))) ˆξ(ˆt0(y, p))) || ˆξ(ˆt0(y, p))|| = ˆGp(ˆx(ˆt0(y, p); x, ξ), ˆξ(ˆt0(y, p); x, ξ)), (A.49) 174 since ray direction does not change in the constant slowness. Then ˆτtt(ˆt0(x, ξ); x, ξ) = ˆGT (cid:33) p (ˆx(ˆt0(x, ξ)), ˆξ(ˆt0(x, ξ))) ˆM (ˆt0(x, ξ); x, ξ) ˆGp(ˆx(ˆt0(x, ξ)), ˆξ(ˆt0(x, ξ))) (ˆt0(x, ξ) − ˆt0(y, p)) p (ˆx(ˆt0(y, p)), ˆξ(ˆt0(y, p))) ˆM (ˆt0(y, p); x, ξ) + (cid:32) = ˆGT d ˆM dt ˆGp(ˆx(ˆt0(y, p)), ˆξ(ˆt0(y, p))) = ˆτtt(ˆt0(y, p); x, ξ) + O((cid:112)||p||), (A.50) dt ∼ O(||p||) by Lemma 3.3.6. On the other hand, |ˆt0(y, p)− ˆt0(x, ξ)| ∼ since we have d ˆM O(||p||−1/2). We then have ˆγ(ˆt0(x, ξ); x, ξ) = = ≈ 1 2(ˆτtt(ˆt0(y, p); x, ξ) + O((cid:112)||p||)) 1 + ((cid:112)||p||)−1 2(ˆτtt(ˆt0(y, p); x, ξ) (cid:32) 1 1 (cid:33) 1 2(ˆτtt(ˆt0(y, p); x, ξ) = ˆγ(ˆt0(y, p); x, ξ). (A.51) Inside the quadratic term, we first have an invariant, ω − ˆτt(ˆt0(x, ξ); x, ξ) = ω − ˆτt(ˆt0(y, p); x, ξ), (A.52) since ˆτt will be a constant along the ray. 175 The next term is ˆϑ(ˆt0(x, ξ); x, ξ). Re(ˆτtx(ˆt0(x, ξ); x, ξ)) = −Re( ˆM )(ˆt0(x, ξ); x, ξ)Gp(ˆx(ˆt0(x, ξ); x, ξ), ˆξ(ˆt0(x, ξ); x, ξ)) = −Re( ˆM )(ˆt0(y, p); x, ξ)Gp(ˆx(ˆt0(y, p); x, ξ), ˆξ(ˆt0(y, p); x, ξ)) + dt (A.53) (A.54) (A.55) (A.56) (A.57) dRe( ˆM )(ˆt0(y, p); x, ξ) = Re(ˆτtx(ˆt0(y, p); x, ξ)) + O((cid:112)||p||). (ˆt0(x, ξ) − ˆt0(y, p))Gp(ˆx(ˆt0(y, p); x, ξ), ˆξ(ˆt0(y, p); x, ξ)) We have the similar conclusion for imaginary part, Im(ˆτtx(ˆt0(x, ξ); x, ξ)) = Im(ˆτtx(ˆt0(y, p); x, ξ)) + O((cid:112)||p||). Consequently, by equation (3.75) ˆϑ(ˆt0(x, ξ); x, ξ)) = ˆϑ(ˆt0(y, p); x, ξ)) + O((cid:112)||p||). Next, r − ˆx(ˆt0(x, ξ)) = r − ˆx(ˆt0(y, p)) + ˆx(ˆt0(y, p)) − ˆx(ˆt0(x, ξ)) = r − ˆx(ˆt0(y, p)) + G± = r − ˆx(ˆt0(y, p)) ± v(ˆx(ˆt0(y, p)))(ˆt0(y, p) − ˆt0(x, ξ)) p (ˆx(ˆt0(y, p)), ˆξ(ˆt0(y, p)))(ˆt0(y, p) − ˆt0(x, ξ)) ˆξ(ˆt0(y, p)) || ˆξ(ˆt0(y, p))|| = r − ˆx(ˆt0(y, p)) + O(1/(cid:112)||p||). 176 G± represents the different Hamiltonian by its sign. Therefore, (cid:33) iˆγ(ˆt0(x, ξ); x, ξ)|ω − ˆτt(ˆt0(x, ξ); x, ξ) − ˆϑ(ˆt0(x, ξ); x, ξ)T (r − ˆx(ˆt0(x, ξ)))|2 = iˆγ(ˆt0(y, p); x, ξ) |ω − ˆτt(ˆt0(y, p); x, ξ) −(cid:16) ˆϑ(ˆt0(y, p); x, ξ) + O((cid:112)||p||) (cid:17)T (r − ˆx(ˆt0(x, ξ))|2 + O ≈ iˆγ(ˆt0(y, p); x, ξ)|ω − ˆτt(ˆt0(y, p); x, ξ) −(cid:16) ˆϑ(ˆt0(y, p); x, ξ)) (cid:17)T (r − ˆx(ˆt0(y, p)))|2. 1(cid:112)||p|| (cid:32) (A.58) Proof of Proposition 3.3.6 Proposition A.0.3. Some real-valued phase terms, ˆ(r, ˆtc; y, p) − ˆθ(r, ˆtc; y, p) and ˆ(r, ˆtc; x, ξ) − ˆθ(r, ˆtc; x, ξ), can be ignored since they are constant order terms with respect to the large wavenumber ||ξl,i|| = ||p + q||. Proof. The first term: ( ˆF − ω) ˆQ (ˆτt(ˆtc; y, p) − ω + Re(ˆτtx)(ˆtc; y, p)) (cid:32) (cid:33) Im(ˆτtx(ˆtc; y, p))T (r − ˆy(ˆtc)) (cid:32) Im(ˆτtt(ˆtc; y, p)) Im(ˆτtx(ˆtc; t, ξ))T (r − ˆx(ˆtc)) (cid:33) − (ˆτt(ˆtc; x, ξ) − ω + Re(ˆτtx)(ˆtc; x, ξ)) Im(ˆτtt(ˆtc; x, ξ)) ≈ (ˆτt(ˆtc; x, ξ) − ω)(ˆy(ˆtc) − ˆx(ˆtc)) + (ˆτt(ˆtc; y, p) − ˆτt(ˆtc; x, ξ))(r − ˆy(ˆtc)) + O(1). By Corollary 3.3.1, we notice the scale of |ˆτt(ˆtc; x, ξ) is controlled by Im(ˆγ) ∼ O( 1||p||). 177 Therefore, by Lemma 3.3.2, |ˆτt(ˆtc; y, p) − ˆτt(ˆtc; x, ξ)| = (cid:12)(cid:12)(cid:12)ˆτt(ˆtc; x, ξ) − ω (cid:12)(cid:12)(cid:12)||ˆy(ˆtc) − ˆx(ˆtc)|| ∼ O(1). (cid:12)(cid:12)(cid:12)v(ˆy(ˆtc))||ˆp(ˆtc)|| − v(ˆx(ˆtc))|| ˆξ(ˆtc)||(cid:12)(cid:12)(cid:12) ≤ |∇v(ˆy(ˆy(ˆtc))|||(ˆy(ˆtc) − ˆx(ˆtc)||||ˆp(ˆy(ˆtc))|| + v(ˆy(ˆtc))(||ˆp(ˆtc)|| − || ˆξ(ˆtc)||). By Lemma 3.3.2, |ˆτt(ˆtc; y, p) − ˆτt(ˆtc; x, ξ)| ∼ O((cid:112)||p||), and ||r − ˆy(ˆtc)|| is controlled by Hessian ˆM (ˆtc). So (ˆτt(ˆtc; y, p) − ˆτt(ˆtc; x, ξ))(r − ˆy(ˆtc)) ∼ O(1). The second term: 1 2 Re(ˆτtt) ˆQ2 1 2 Re(ˆτtt) ˆQ2 ≈ 1 2 Re(ˆτtt)(r − ˆy(ˆtc))2 ∼ O(1), (A.59) by Lemma 3.3.2. And the same analysis can be applied to the term associated with the beam (x, ξ). Proof of Proposition 3.3.7 Proposition A.0.4. For the first two terms in equation (3.107), their exponents sat- isfy, − Im( ˆβ)|ω − ˆτt(ˆtc; y, p) − ˆζT (r − ˆy(ˆtc))|2 − Im(ˆγ)|ω − ˆτt(ˆtc; x, ξ) − ˆϑT (r − ˆx(ˆtc))|2 = −||r − ˆx(ˆtc) + ˆy(ˆtc) 2Im( ˆβ)ˆζ ˆζT − Im( ˆβ) ||2 ||ˆy(ˆtc) − ˆx(ˆtc)||2 ˆζ ˆζT 2 2 (cid:32) (cid:33) 1(cid:112)||p|| , (A.60) − Im( ˆβ)|ω − ˆτt(ˆtc; y, p)|2 − Im(ˆγ)|ω − ˆτt(ˆtc; x, ξ)|2 + O 178 where ˆβ, ˆγ, ˆζ and ˆϑ are all defined at ˆtc. Similarly, ˆg in equation (3.108) is an O(1) term. Proof. The quadratic term about r in equation (A.60) is, −rT(cid:16) Im( ˆβ)ˆζ ˆζT + Im(ˆγ) ˆϑ( ˆϑ)T(cid:17) r. (A.61) The matrix in the parentheses is positive semi-definite matrix as it is the sum of two positive semi-definite matrices. We apply the eigenvalue decomposition Im( ˆβ)ˆζ ˆζT + Im(ˆγ) ˆϑ( ˆϑ)T = QT ΛQ, (A.62) where Λ is a diagonal matrix with non-negative entries and QT Q = I. The cross term about r in equation (A.60) is 2Im( ˆβ)(ω − ˆτt(ˆtc; y, p))ˆζT r + 2Im( ˆβ)ˆy(ˆtc)T (ˆζ ˆζT )r +2Im(ˆγ)(ω − ˆτt(ˆtc; x, ξ)) ˆϑT r + 2Im(ˆγ)ˆx(ˆtc)T ( ˆϑ ˆϑT )r = 2 ˆJ T 1 (ˆtc, ω; y, p, x, ξ)r + 2 ˆJ T 2 (ˆtc, ω; y, p, x, ξ)r, (A.63) where ˆJ1(ˆtc, ω; y, p, x, ξ) = Im( ˆβ)(ω − ˆτt(ˆtc; y, p))ˆζ + Im(ˆγ)(ω − ˆτt(ˆtc; x, ξ)) ˆϑ; (A.64) ˆJ2(ˆtc, ω; y, p, x, ξ) = Im( ˆβ)ˆζ ˆζT ˆy(ˆtc) + Im(ˆγ) ˆϑ ˆϑT ˆx(ˆtc). (A.65) 179 Combine the cross term (A.63) and second order term (A.61), −(Qr)T Λ(Qr) + 2 ˆJ T 1 QT Qr + 2 ˆJ T 2 QT Qr. To make a complete quadratic term, its central point rc is, (cid:32) rc = QT Λ−1Q( ˆJ1 + ˆJ2) = QT Λ−1Q ˆJ2 + O (cid:33) , 1(cid:112)||p|| (A.66) (A.67) since and || ˆJ1|| ≤ ||Im( ˆβ)(ω − ˆτt(ˆtc; y, p))ˆζ|| + ||Im(ˆγ)(ω − ˆτt(ˆtc; x, ξ)) ˆϑ|| (cid:18) 1 (cid:19) ||p|| O (cid:16)(cid:112)||p||(cid:17) ≤ O (cid:16)(cid:112)||p||(cid:17) , O (||p||) ≤ O || ˆJ2|| ≤ ||Im( ˆβ)ˆζ ˆζT ˆy(ˆtc)|| + ||Im(ˆγ) ˆϑ ˆϑT ˆx(ˆtc)|| O(1) ≤ O(||p||). (cid:16)||p||2(cid:17) ≤ O (cid:18) 1 (cid:19) O ||p|| Here the inverse matrix is defined as the pseudo-inverse matrix for a rank-deficient matrix, that is Λ−1 (cid:18) we set its center rc’s ith coordinate to be the same as ˆy(ˆtc)+ˆx(ˆtc) If the diagonal term in Λ−1 ii = 0, if Λii = 0. ’s, since we have is zero, then (cid:19) ii 2 in the expression of ˆB. Therefore, we carry out exp − 1 4||2r − ˆx(ˆtc) − ˆy(ˆtc)||2 ˆM(ˆtc) 180 calculation by assuming Λ invertible. QT Λ−1Q = (Im( ˆβ)ˆζ ˆζT + Im(ˆγ) ˆϑ( ˆϑ)T )−1 = (2Im( ˆβ)ˆζ ˆζT + Im(ˆγ) ˆϑ( ˆϑ)T − Im( ˆβ)ˆζ ˆζT )−1. To evaluate the order of Im(ˆγ) ˆϑ( ˆϑ)T − Im( ˆβ)ˆζ ˆζT , we first have | ˆβ − ˆγ| = 1 2 1 (cid:12)(cid:12)(cid:12) ≤(cid:12)(cid:12)(cid:12) ˆτtt(ˆtc; y, p) − ˆτtt(ˆtc; x, ξ) ˆτtt(ˆtc; y, p)ˆτtt(ˆtc; x, ξ) − ˆτtt(ˆtc; y, p) ˆτtt(ˆtc; x, ξ) 1 (cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12). By Corollary 3.3.3, we have |ˆτtt(ˆtc; y, p) − ˆτtt(ˆtc; x, ξ)| ∼ O((cid:112)||p||). Consequently, (A.68) | ˆβ − ˆγ| ≤ O(||p||− 3 2 ). On the other hand, by using Corollary 3.3.3, ||ˆζ − ˆϑ|| ≤ O((cid:112)||p||). (A.69) (A.70) Then ||Im(ˆγ) ˆϑ( ˆϑ)T−Im( ˆβ)ˆζ ˆζT|| ≤ ||Im( ˆβ) ≤ O((cid:112)||p||). (cid:16)ˆζ ˆζT − ˆϑ ˆϑT(cid:17)|| + ||Im( ˆβ − ˆγ) ˆϑ ˆϑT|| (A.71) 181 Therefore, QT Λ−1Q = (2Im( ˆβ)ˆζ ˆζT + O((cid:112)||p||))−1 ≈ (2Im( ˆβ)ˆζ ˆζT )−1. (A.72) Similarly, ˆJ2 = Im( ˆβ)ˆζ ˆζT (ˆy(ˆtc) + ˆx(ˆtc)) − Im( ˆβ)ˆζ ˆζT ˆx(ˆtc) + Im(ˆγ) ˆϑ( ˆϑ)T ˆx(ˆtc) = Im( ˆβ)ˆζ ˆζT (ˆy(ˆtc) + ˆx(ˆtc)) + O((cid:112)||p||). ≈ Im( ˆβ)ˆζ ˆζT (ˆy(ˆtc) + ˆx(ˆtc)). (A.73) The central point rc becomes, rc ≈ QT Λ−1Q ˆJ2 ≈ (2Im( ˆβ)ˆζ ˆζT )−1Im( ˆβ)ˆζ ˆζT (ˆy(ˆtc) + ˆx(ˆtc)) = ˆx(ˆtc) + ˆy(ˆtc) 2 . (A.74) Equation (A.60) now becomes, − Im( ˆβ)|ω − ˆτt(ˆtc; y, p) − ˆζT (r − ˆy(ˆtc))|2 − Im(ˆγ)|ω − ˆτt(ˆtc; x, ξ) − ( ˆϑ)T (r − ˆx(ˆtc))|2 = −Im( ˆβ)|ω − ˆτt(ˆtc; y, p)|2 − Im(ˆγ)|ω − ˆτt(ˆtc; x, ξ)|2 − ||r − ˆx(ˆtc) + ˆy(ˆtc) ||2 2Im( ˆβ)ˆζ ˆζT 2 + ˆJ T 2 (2Im( ˆβ)ˆζ ˆζT )−1 ˆJ2 − Im( ˆβ)|ˆζT ˆy(ˆtc)|2 − Im(ˆγ)|( ˆϑ)T ˆx(ˆtc)|2. 182 Using equation (A.71), 2 (2Im( ˆβ)ˆζ ˆζT )−1 ˆJ2 − Im( ˆβ)|ˆζT ˆy(ˆtc)|2 − Im(ˆγ)|( ˆϑ)T ˆx(ˆtc)|2 ˆJ T ≈ ||ˆy(ˆtc) + ˆx(ˆtc)||2 Im( ˆβ)ˆζ ˆζT − ||ˆx(ˆtc)||2 − ||ˆy(ˆtc)||2 Im( ˆβ)ˆζ ˆζT Im( ˆβ)ˆζ ˆζT 2 = −||ˆy(ˆtc) − ˆx(ˆtc)||2 Im( ˆβ)ˆζ ˆζT . 2 Equation (A.60) now becomes, (A.75) − Im( ˆβ)|ω − ˆτt(ˆtc; y, p) − ˆζT (r − ˆy(ˆtc))|2 − Im(ˆγ)|ω − ˆτt(ˆtc; x, ξ) − ( ˆϑ)T (r − ˆx(ˆtc))|2 ≈ −Im( ˆβ)|ω − ˆτt(ˆtc; y, p)|2 − Im(ˆγ)|ω − ˆτt(ˆtc; x, ξ)|2 − ||r − ˆx(ˆtc) + ˆy(ˆtc) ||ˆy(ˆtc) − ˆx(ˆtc)||2 2Im( ˆβ)ˆζ ˆζT − Im( ˆβ) ||2 ˆζ ˆζT . 2 2 By applying the similar computation, the exponent of the last term in equation (3.107) denoted as ˆg contains two O(1) terms. Proof of Proposition 3.3.8 Proposition A.0.5. The function ˆφ1 satisfies ˆφ1(t; x, ξ, y, p) = ||ˆp(t) − ˆξ(t)||2 Gpp + O 1 2 (cid:32) (cid:33) . 1(cid:112)||p|| (A.76) Proof. Using Lemma 3.3.5 for ˆφ1(t), ˆφ1(t; x, ξ, y, p) = v(ˆx(t)) 183 (cid:32) (cid:33) . ||ˆΞ(t)||2 ˆκ(t)2||ˆp(t)|| 1 2 We then would like to explore the term ||ˆp(t) − ˆξ(t)||2 Gpp(ˆy(t),ˆp(t)) by taking advantage of Gpp’s expression. ||ˆp(t) − ˆξ(t)||2 Gpp(ˆy(t),ˆp(t)) = (ˆp(t) − ˆξ(t))T(cid:0)Gpp(ˆy(t), ˆp(t))(cid:1) (ˆp(t) − ˆξ(t)) ||ˆp(t)||3 (ˆp(t) − ˆξ(t))T(cid:16)||ˆp(t)||2I − ˆp(t)ˆp(t)T(cid:17) (cid:17) ||ˆp(t)|| (−ˆΞ(t))T(cid:16) ||ˆp(t)|| (− ˆξ(t) + ˆκ(t)ˆp(t))T (ˆp(t) − ˆξ(t)) (1 − ˆκ(t))ˆp(t) − ˆΞ(t) v(ˆy(t)) v(ˆy(t)) v(ˆy(t)) = = = = v(ˆy(t)) ||ˆΞ(t)||2 ||ˆp(t)|| . Then, 1 2 ||ˆp(t) − ˆξ(t)||2 ˆGpp = v(ˆy(t)) 2 ||ˆΞ(t)||2 ||ˆp(t)|| . (ˆp(t) − ˆξ(t)) (A.77) (A.78) Compare ˆφ1(t) with equation (A.77), (cid:32) v(ˆx(t)) (cid:33) ||ˆΞ(t)||2 ˆκ(t)2||ˆp(t)|| 1 2 = O(||ˆy(t) − ˆx(t)||). − 1 2 v(ˆy(t)) ||ˆΞ(t)||2 ||ˆp(t)|| ≈ − 1 2 (cid:18) v(ˆy(t)) − v(ˆx(t)) ˆκ(t)2 (cid:19) ||ˆΞ(t)||2 ||ˆp(t)|| (A.79) This is because ˆκ(t) ∼ 1 + O(||ˆp(t)||− 1 ˆξ(t)|| ∼ O(||p||1/2) in Lemma 3.3.2. 2 ) and ˆΞ(t) ∼ O((cid:112)||ˆp(t)||) followed by ||ˆp(t) − Then the derivative ˆφ1 becomes, ˆφ1(t; x, ξ, y, p) ≈ 1 2 ||ˆp(t) − ˆξ(t)||2 Gpp 184 Proof of Proposition 3.3.10 Proposition A.0.6. There exists a linear map ˆJ (ˆtc; y, p), such that ˆy(ˆtc) − ˆx(ˆtc) ˆp(ˆtc) − ˆξ(ˆtc)  ≈ ˆJ (ˆtc; y, p) y − x p − ξ  . (A.80) Proof. We consider (ˆy(t)−ˆx(t), ˆp(t)− ˆξ(t)) as a function about t and its initial condition (y − x, p − ξ). We denote this initial condition as (∆x, ∆ξ) and ˆJ (t; y, p) as the dynamical variational system, ˆy(ˆtc) − ˆx(ˆtc; ∆x, ∆ξ) ˆp(ˆtc) − ˆξ(ˆtc; ∆x, ∆ξ)  ≈ ˆJ (ˆtc; y, p)  . ∆x ∆ξ It is well known that ˆJ (t; y, p) satisfies,  Gxp(ˆy(t), ˆp(t)) d ˆJ (t) dt = Gpp(ˆy(t), ˆp(t)) −Gxx(ˆy(t), ˆp(t)) −GT xp(ˆy(t), ˆp(t))  ˆJ (t), and its initial condition is an identity matrix, ˆJ (0) = I2d. 185 (A.81) (A.82) (A.83) By solving ˆJ (t), we can get ˆy(t) − ˆx(t; y − x, p − ξ) ˆp(t) − ˆξ(t; y − x, p − ξ)  = ˆJ (t; y, p) y − x p − ξ  . (A.84) The transform ˆJ defined above is invertible due to the uniqueness of ODE system’s solution. Proof of Lemma 3.3.9 (cid:90) Lemma A.0.4. The result after taking the integral about ω can be approximated, − ei(ω−ˆτt(ˆtc;y,p))∆ˆt0(x,ξ;y,p)ei(ω−˜τt(˜tc;y,q))∆˜t0(x,η;y,q)ω2 ˆB(x, ξ, ω; y, p) ˜B(x, η, ω; y, q)dω = eiO(1)K(p, q, y) ˆB(x, ξ; y, p, q) ˜B(x, η; y, p, q)e − i 2||˜y(˜tc)−˜x(˜tc)||2 e Re( ˜M (˜tc)). − i 2||ˆy(ˆtc)−ˆx(ˆtc)||2 Re( ˆM (ˆtc)) Proof. All terms containing ω in ˆB(x, ξ, ω; y, p) (3.110) are as the following, (cid:16)−Im( ˆβ)|ω − ˆτt(ˆtc; y, p)|2 − Im(ˆγ)|ω − ˆτt(ˆtc; x, ξ)|2(cid:17) ≈ (cid:32) (cid:32) (cid:19)2 (cid:18) ω − ˆτt(ˆtc; y, p) + ˆτt(ˆtc; x, ξ) 2 + (ˆτt(ˆtc; y, p) − ˆτt(ˆtc; x, ξ))2 (cid:33)(cid:33) exp exp ≈ e −Im( ˆβ) (cid:18) −2Im( ˆβ) ω− ˆτt(ˆtc;y,p)+ˆτt(ˆtc;x,ξ) 2 e−Im( ˆβ)|ˆτt(ˆtc;y,p)−ˆτt(ˆtc;x,ξ)|2 . (A.85) 2 (cid:19)2 186 Similarly, for ˜B, (cid:16)−Im( ˜β)|ω − ˜τt(˜tc; y, q)|2 − Im(˜γ)|ω − ˜τt(˜tc; x, η)|2(cid:17) exp (cid:18) −2Im( ˜β) ≈ e ω− ˜τt(˜tc;y,q)+˜τt(˜tc;x,η) 2 e−Im( ˜β)|˜τt(˜tc;y,q)−˜τt(˜tc;x,η)|2 . (A.86) Combine the terms containing ω in equations (A.85) and (A.86) with −ω2, and denote (cid:19)2 (cid:18) (cid:90) dωe (cid:19)2 − ℵ(ˆtc, ˜tc; x, ξ, η, y, p, q) = (cid:18) −2Im( ˆβ) e ω− ˆτt(ˆtc;y,p)+ˆτt(ˆtc;x,ξ) 2 −2Im( ˜β) ω− ˜τt(˜tc;y,q)+˜τt(˜tc;x,η) 2 (cid:19)2 ω2 ei(ω−ˆτt(ˆtc;y,p))∆ˆt0(x,ξ;y,p)ei(ω−˜τt(˜tc;y,q))∆˜t0(x,η;y,q). (A.87) Then the target integral becomes, (cid:90) − = ℵe−Im( ˜β)|˜τt(˜tc;y,q)−˜τt(˜tc;x,η)|2 ei(ω−ˆτt(ˆt0;y,p))∆ˆt0(x,ξ;y,p)ei(ω−˜τt(˜tc;y,q))∆˜t0(x,η;y,q)ω2 ˆB(x, ξ, ω; y, p) ˜B(x, η, ω; y, q)dω e−Im( ˆβ)|ˆτt(ˆtc;y,p)−ˆτt(ˆtc;x,ξ)|2 , (A.88) without considering the constant terms in ˆB and ˜B. If we can approximate ℵ by the product of a constant K(p, q, y), functions on the receiver side and functions on the source side, then the proposition is proved. 187 Compute the expression (A.87), −ℵ(ˆtc, ˜tc; x, ξ, η, y, p, q) = e−iˆτt(ˆtc;y,p)∆ˆt0(x,ξ;y,p)−i˜τt(˜tc;y,p)∆˜t0(x,η;y,q)e −2 Im( ˆβ)Im( ˜β)S2 2 Im( ˆβ)+Im( ˜β) ω2eiω(∆ˆt0(x,ξ;y,p)+∆˜t0(x,η;y,q))e−2(Im( ˆβ)+Im( ˜β))(ω−S1)2 dω, (cid:90) where S1(ˆtc, ˜tc; x, ξ, η, y, p, q) = (cid:18) Im( ˆβ) ˆτt(ˆtc;y,p)+ˆτt(ˆtc;x,ξ) 2 (cid:19) + Im( ˜β) Im( ˆβ) + Im( ˜β) (A.89) (cid:16) ˜τt(˜tc;y,q)+˜τt(˜tc;x,η) 2 (cid:17) , (A.90) S2(ˆtc, ˜tc; x, ξ, η, y, p, q) = ˆτt(ˆtc; y, p) + ˆτt(ˆtc; x, ξ) 2 − ˜τt(˜tc; y, q) + ˜τt(˜tc; x, η) 2 . (A.91) We then have, ℵ = e−iˆτt(ˆtc;y,p)∆ˆt0(x,ξ;y,p)−i˜τt(˜tc;y,p)∆˜t0(x,η;y,q)e −2 Im( ˆβ)Im( ˜β)S2 2 Im( ˆβ)+Im( ˜β) eiS1te d2  − t2 2Im( ˆβ+ ˜β) dt2 (cid:12)(cid:12)(cid:12)t=∆ˆt0(x,ξ;y,p)+∆˜t0(x,η;y,q) . (A.92) Proposition A.0.7. Both S1 and S2 can be approximated as constants only related to 188 the fixed parameter (y, p) and (y, q), that is S1(ˆtc, ˜tc; x, ξ, η, y, p, q) ≈ Im( ˆβ)(cid:0)ˆτt(ˆtc; y, p)(cid:1) + Im( ˜β)(cid:0)˜τt(˜tc; y, q)(cid:1) , Im( ˆβ) + Im( ˜β) S2(ˆtc, ˜tc; x, ξ, η, y, p, q) ≈ ˆτt(ˆtc; y, p) − ˜τt(˜tc; y, q). Proof. For S1, S1(ˆtc, ˜tc; x, ξ, η, y, p, q) = Im( ˆβ)(cid:0)ˆτt(ˆtc; y, p)(cid:1) + Im( ˜β)(cid:0)˜τt(˜tc; y, q)(cid:1) Im( ˆβ) + Im( ˜β) Im( ˆβ)(ˆτt(ˆtc; x, ξ) − ˆτt(ˆtc; y, p)) + Im( ˜β)(˜τt(˜tc; x, η) − ˜τt(˜tc; y, q)) + 2Im( ˆβ) + 2Im( ˜β) Im( ˆβ)(cid:0)ˆτt(ˆtc; y, p)(cid:1) + Im( ˜β)(cid:0)˜τt(˜tc; y, q)(cid:1) ≈ Im( ˆβ)(cid:0)ˆτt(ˆtc; y, p)(cid:1) + Im( ˜β)(cid:0)˜τt(˜tc; y, q)(cid:1) Im( ˆβ) + Im( ˜β) = Im( ˆβ) + Im( ˜β) + O((cid:112)||ˆpt||) + O((cid:112)||˜qt||) , (A.93) since the term in the last step above is about O(||p|| + ||q||), and 2Im( ˆβ) + 2Im( ˜β) Im( ˆβ)(ˆτt(ˆtc; x, ξ) − ˆτt(ˆtc; y, p)) + Im( ˜β)(˜τt(˜tc; x, η) − ˜τt(˜tc; y, q)) ≤ |ˆτt(ˆtc; x, ξ) − ˆτt(ˆtc; y, p)| + |(˜τt(˜tc; x, η) − ˜τt(˜tc; y, q)| ≤ O((cid:112)||ˆpt|| + ||˜qt||). S1(ˆtc, ˜tc; x, ξ, η, y, p, q) ≈ Im( ˆβ)(cid:0)ˆτt(ˆtc; y, p)(cid:1) + Im( ˜β)(cid:0)˜τt(˜tc; y, q)(cid:1) . Im( ˆβ) + Im( ˜β) For S2, S2(ˆtc, ˜tc; x, ξ, η, y, p, q) = ˆτt(ˆtc; y, p) − ˜τt(˜tc; y, q) + O((cid:112)||p||), (A.94) (A.95) (A.96) due to the fact that |ˆτt(ˆtc; y, p) − ˆτt(ˆtc; x, ξ)| = v(y)||p|| − v(x)||ξ|| and ||p − ξ|| ∼ 189 O((cid:112)||p||). Then, −2 e Im( ˆβ)Im( ˜β)S2 2 Im( ˆβ)+Im( ˜β) ≈ e −2 Im( ˆβ)Im( ˜β)(ˆτt(ˆtc;y,p)−˜τt(˜tc;y,q))2 Im( ˆβ)+Im( ˜β) eO(1), (A.97) since Im( ˆβ) ∼ O( 1||p||) and Im( ˜β) ∼ O( 1||q||). We then approximate S2 by S2(ˆtc, ˜tc; x, ξ, η, y, p, q) ≈ ˆτt(ˆtc; y, p) − ˜τt(˜tc; y, q). We also obtain |ˆτt(ˆtc; y, p) − ˜τt(˜tc; y, q)| is around O(max((cid:112)||p||,(cid:112)||q||)). (A.98) Proposition A.0.8. −|∆ˆt0(x,ξ;y,p)|2 4Im( ˆβ+ ˜β) −|∆˜t0(x,η;y,q)|2 4Im( ˆβ+ ˜β) e −2 e Im( ˆβ)Im( ˜β)(S2)2 Im( ˆβ)+Im( ˜β) ℵ ≈ −S2 1e (A.99) Proof. The second order time derivative in equation (A.92), (cid:32) iS1 − ∆ˆt0(x, ξ; y, p) + ∆˜t0(x, η; y, q) − 1 (cid:33)2  Im( ˆβ + ˜β) Im( ˆβ + ˜β) −|∆ˆt0(x,ξ;y,p)+∆˜t0(x,η;y,q)|2 2(Im( ˆβ+ ˜β)) . eiS1(∆ˆt0(x,ξ;y,p)+∆˜t0(x,η;y,q))e We now conduct some asymptotic analysis about the terms in the above equation, S1 ∼ O(||p|| + ||q||), 1 Im( ˆβ + ˜β) ∼ O(||p|| + ||q||), (A.100) 190 since ˆβ = 1 2ˆτtt(ˆtc;y,p) and ˜β = 1 2˜τtt(˜tc;y,q) ∆ˆt0(x, ξ; y, p) + ∆˜t0(x, η; y, q) Im( ˆβ + ˜β) ∼ O (cid:32) O(||p|| + ||q||). (A.101) . On the other hand, (cid:33) 1(cid:112)||p|| + 1(cid:112)||q|| Moreover, we have −|∆ˆt0(x, ξ; y, p) − ∆˜t0(x, η; y, q)|2 2(Im( ˆβ + ˜β)) − |∆ˆt0(x, ξ; y, p)|2 4Im( ˆβ + ˜β) = −|∆˜t0(x, η; y, q)|2 4Im( ˆβ + ˜β) |∆ˆt0(x, ξ; y, p) − ∆˜t0(x, η; y, q)|2 = −|∆˜t0(x, η; y, q)|2 4Im( ˆβ + ˜β) 4Im( ˆβ + ˜β) − |∆ˆt0(x, ξ; y, p)|2 4Im( ˆβ + ˜β) + Consequently, ℵ ≈ −S2 1ei(S1−ˆτt)(∆ˆt0(x,ξ;y,p))ei(S1−˜τt)(∆˜t0(x,η;y,q))e −|∆ˆt0(x,ξ;y,p)|2 4Im( ˆβ+ ˜β) −|∆˜t0(x,η;y,q)|2 4Im( ˆβ+ ˜β) e −2 e Im( ˆβ)Im( ˜β)S2 2 Im( ˆβ)+Im( ˜β) . + O(1). (A.102) With respect to two real-valued phase terms in equation (A.102), ei(S1−ˆτt)(∆ˆt0(x,ξ;y,p)) = e i Im( ˜β)(ˆτt(ˆtc;y,p)−˜τt(˜tc;y,q)) Im( ˆβ+ ˜β) ∆ˆt0(x,ξ;y,p) ∼ eiO(1), (A.103) and ei(S1−˜τt)(∆˜t0(x,η;y,q)) ∼ eiO(1). (A.104) 191 To summarize, −|∆ˆt0(x,ξ;y,p)|2 ℵ ≈ −S2 4Im( ˆβ+ ˜β) 1e −|∆ˆt0(x,ξ;y,p)|2 4Im( ˆβ+ ˜β) −|∆˜t0(x,η;y,q)|2 4Im( ˆβ+ ˜β) −|∆˜t0(x,η;y,q)|2 4Im( ˆβ+ ˜β) −2 e ≈ K(p, q, y)e e Im( ˆβ)Im( ˜β)(ˆτt(ˆtc;y,p)−˜τt(˜tc;y,q))2 Im( ˆβ)+Im( ˜β) e . (A.105) This is exactly the goal (A.88) we want to achieve. 192 Appendix B Proof in Elastic Wave Proof of Positive Definite Hessian Matrix Theorem B.0.1. The imaginary part of the Hessian for every beam preserves the S.P.D property after reflection, if it is neither grazing ray nor evanescent wave after reflection. Proof. In this appendix, we consider the reflection happens on the surface {x = (x, y, z) : x = 0} without the loss of generality. We denote the phase function of both P and S-wave as τ . The same rule can be applied to the velocity c. To simplify the presentation, all Hessian matrices mentioned below are about the imaginary part only, if not specified. We assume the reflection point is x0 = (x0, y0, z0), and all terms below are defined at this point, if not specified. The last simplification is that we follow the positive Hamiltonian throughout this proof, and the negative Hamiltonian will be treated similarly. 193 Transformation between Hessian matrices We first define a new matrix ˜M at the reflection point  ˜M = τtt τty τtz τty τyy τyz τtz τyz τzz  . (B.1) Therefore, we can define a transform between the Hessian matrix M and ˜M following the certain eikonal equation. ℵ( ˜M ) =  τxx τxy τxz τxy τyy τyz τxz τyz τzz  = M. (B.2) The terms involved with the variable x are defined by following the certain eikonal equation and in the way shown in Section 4.5. Moreover, ℵP means that the transform follows the P-wave eikonal equation and ℵS follows the S-wave eikonal equation. If we can prove the transform ℵ and its inverse transform ℵ−1 preserve the S.P.D. property, then the theorem is proved since M new = ℵP ( ˜M ) = ℵP (ℵ−1 ℵS( ˜M ) = ℵS(ℵ−1 P (M )). To prove this, instead of considering two types of matrix in S (M )) or M new = equation (B.2) directly, we would like to base our proof first on the complete matrix Mc. Mc =  τtt ∇τ T ∇τt M t  , where ∇τt is the gradient of the phase function’s time derivative τt and M is the 194 original Hessian. Both ˜M and M are submatrix of Mc. Proof by Contradiction: ∇c = 0 We start with the simpler case, i.e. ∇c = 0 at the point x0. Lemma B.0.5. If ∇c vanishes at x0, then Mc is a positive semi-definite matrix and with rank-three. Proof. We write the complete matrix Mc first,  τtt ∇τ T ∇τt M t  = c2∇τ T |∇τ| M ∇τ|∇τ| c M∇τ |∇τ|  . c∇τ T M|∇τ| M To show that the matrix Mc is positive semi-definite, we use v = (α, p)T , vT Mcv = α2c2∇τ T |∇τ| M ∇τ |∇τ| + p (cid:18) αc = ∇τ |∇τ| + 2αc (cid:19)T (cid:18) M αc ∇τ T M p |∇τ| + pT M p, ∇τ |∇τ| + p (cid:19) . (B.3) (B.4) (B.5) Equation (B.5) shows that the null space of Mc is an one-dimensional space and its basis is ˜v, ˜v = (B.6)  1(cid:112) − c(cid:112) 1+c2 1+c2 ∇τ|∇τ|  . The assumption that there are no grazing rays guarantees that τx (cid:54)= 0 for both beams before and after reflection. 195 We now start to prove the transform ℵ and its inverse transform preserve the S.P.D. property when ∇c = 0. In other words, if M is S.P.D, then ˜M = ℵ−1(M ) is S.P.D. On the other hand, if we have ˜M is S.P.D, then ℵ( ˜M ) is also S.P.D. for both P-wave and S-wave eikonal equations. Case I: ˜M = ℵ−1(M ) There are several steps involved to prove ˜M is S.P.D. We first consider ˜M is a submatrix of the corresponding complete matrix Mc. Then we use Lemma B.0.5 to show this submatrix is S.P.D. For any vector u = (u1, u2, u3) ∈ R3, (cid:18) (cid:19) u1 0 u2 u3 Mc uT ˜M u =  , (B.7) u1 0  u2 u3 if the reflection happens on the surface {x = (x, y, z) : x = 0}. Since there’s only a single basis ˜v in the null space of the matrix Mc, all vectors in the form (u1, 0, u2, u3)T is not parallel to ˜v. Moreover, the complete matrix Mc is a positive semi-definite matrix. Then for any vector {u = (u1,··· , u4) : u2 = 0}, equation (B.7) will be positive. Consequently, uT ˜M u > 0 for any u and ˜M is S.P.D. Case 2: M = ℵ( ˜M ) Similar idea will be applied here. The Hessian M firstly is treated as a submatrix of the complete matrix Mc and then use the fact that Mc is a positive semi-definite 196 matrix. In terms of the transform M = ℵ( ˜M ), for any vector u ∈ R3, (cid:18) 0 uT (cid:19) Mc uT M u = 0  . u (B.8) Obviously, all the vectors concerned above (0, u)t is not in the null space of the complete matrix Mc. In other words, for any vector (0, u), (0, u)T (cid:54)= β ˜v, ∀β ∈ R(cid:54)=0 (B.9) Consequently, uT M u > 0 for any u and M is positive definite. Proof by Contradiction: ∇c (cid:54)= 0 We will follow the similar path as the constant velocity case. First, we prove the imaginary part of the complete matrix Mc is a positive semi-definite matrix. Lemma B.0.6. The imaginary part of the complete matrix Mc is a positive semi- definite matrix and with rank-three. Moreover, the single basis ˜v in its null space is ˜v = (B.10)  1 −c ∇τ|∇τ|  197 Proof. We first prove ˜v is in the null space Im(Mc) = |∇τ| c2∇τ T Im(M )∇τ |∇τ|2 c Im(M )∇τ  ∇τ T Im(M ) |∇τ| c Im(M ) (B.11) Here, although the gradient of the velocity ∇c (cid:54)= 0, this will only affect the real part. Therefore, we can apply the same argument in Lemma B.0.5 to prove. To prove that the transform ℵ and its inverse transform will preserve S.P.D property, we can use the same idea in the case ∇c = 0. The reason is that we only care about the imaginary part of the matrix and their imaginary parts are exactly the same thing as the ones in constant velocity case. FDTD As we mentioned previously, the reference solution is generated by the FDTD solution with the staggered grid. Its correctness will be checked here. To test this, we compare the FDTD solution with the exact solution in the general boundary value problem. If this more general problem is solved correctly, then our reference solution is justified. The parameters used here are, λ = 2; µ = 1; ρ = 1; 198 while the initial condition f here is zero vector −→ 0 , and the initial velocity g is 8π cos(8πt) sin(8πx) 0 0   With the appropriate boundary condition, we know its exact solution is sin(8πt) sin(8πx). We compare the result at T = 0.8, In Figure B.1, the blue line is the correct result, Figure B.1: FDTD solution justify. while the red star curve is the FDTD result with mesh size h = 0.01. Furthermore, we display its convergence rate in Figure B.2. We start the mesh size from 1 50 to 1 400 and each time the grid size is reduced by half. The blue line in Figure B.2 shows the logarithm of L2-error on each mesh size, while the red star line is a linear function with the slope log(1/2) for comparison. 199 00.10.20.30.40.5−1.5−1−0.500.511.5xy Figure B.2: Convergence Rate of FDTD algorithm 200 11.522.533.54−5.5−5−4.5−4−3.5−3−2.5number of grid pointslog(error) BIBLIOGRAPHY 201 BIBLIOGRAPHY [1] K. Aki and P. Richards. Quantitative seismology. Freeman and Co., 1980. [2] U. Albertin, D. Yingst, P. Kitchenside, and V. Tcheverda. True-amplitude beam migration. In 74th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, pages 949–952. Soc. Expl. Geophys., 2004. [3] V. M. Babich and V. S. Buldyrev. Asymptotic methods in short wave diffraction problems (in Russian). Nauka, Moscow, 1972. [4] G. Bao, J. Lai, and J. Qian. Fast multiscale Gaussian beam methods for wave equations in bounded domains. J. Comput. Phys., 261:36–64, 2014. [5] G. Bao, J. Qian, L. Ying, and H. Zhang. A convergent multiscale Gaussian-beam parametrix for the wave equation. Commun. PDEs, 38:92–134, 2013. [6] J. D. Benamou. An introduction to Eulerian geometrical optics (1992 - 2002). J. Sci. Comp., 19:63–93, 2003. [7] G. Beylkin and R. Burridge. Linearized inverse scattering problem of acoustics and elasticity. Wave Motion, 12:15–22, 1990. [8] N. Bleistein. On imaging of reflectors in the earth. Geophysics, 52:931–942, 1987. [9] S. Bougacha, J. Akian, and R. Alexandre. Gaussian beams summation for the wave equation in a convex domain. Commun. Math. Sci., 7:973–1008, 2009. [10] R. Burridge. Asymptotic evaluation of integrals related to time-dependent fields near caustics. SIAM J. Appl. Math., 55:390–409, 1995. [11] E. Cand`es and L. Demanet. The curvelet representation of wave propagators is optimally sparse. Commun. Pure Appl. Math., 58:1472–1528, 2005. [12] E. Candes, L. Demanet, and L. Ying. Fast computation of Fourier integral operators. SIAM J. Sci. Comput., 29:2464–2493, 2007. [13] E. Cand`es and D. Donoho. New tight frames of curvelets and optimal represen- tations of objects with piecewise C2 singularities. Commun. Pure Appl. Math., 57:219–266, 2004. [14] V. Cerveny. Seismic Ray Theory. Cambridge University Press, Cambridge, 2001. 202 [15] V. Cerveny, M. Popov, and I. Psencik. Computation of wave fields in inhomo- geneous media-Gaussian beam approach. Geophys. J. R. Astr. Soc., 70:109–128, 1982. [16] J. Claerbout. Towards a unified theory of reflection mapping. Geophysics, 36:467–481, 1971. [17] J. F. Claerbout. Imaging the earth’s interior. Blackwell Scientific Publications, Cambridge, MA, 1985. [18] M. G. Crandall and P. L. Lions. Viscosity solutions of Hamilton-Jacobi equations. Trans. Amer. Math. Soc., 277:1–42, 1983. [19] L. Demanet and L. Ying. Wave atoms and sparsity of oscillatory patterns. Appl. Comput. Harmon. Anal., 23(3):368–387, 2007. [20] L. Demanet and L. Ying. Wave atoms and time upscaling of wave equations. Numerische Mathematik, 113(1):1–71, 2009. [21] Feng Deng and George A. McMechan. True-amplitude prestack depth migration. GEOPHYSICS, 72(3):S155–S166, 2007. [22] B. Engquist and O. Runborg. Computational high frequency wave propagation. Acta Numerica, 12:181–266, 2003. [23] D. Geba and D. Tataru. A phase space transform adapted to the wave equation. Commun. Partial Diff. Eq., 32:1065–1101, 2007. [24] Yu Geng, Ru-Shan Wu, and Jinghuai Gao. Gabor-frame-based gaussian packet migration. Geophysical Prospecting, 62(6):1432–1452, 2014. [25] S. Geoltrain and J. Brac. Can we image complex structures with first-arrival traveltime. Geophysics, 58:564–575, 1993. [26] S. Gray. Gaussian beam migration of common shot records. Geophysics, 70:S71–S77, 2005. [27] S. Gray and W. May. Kirchhoff migration using eikonal equation traveltimes. Geophysics, 59:810–817, 1994. [28] N. Hill. Gaussian beam migration. Geophysics, 55:1416–1428, 1990. [29] N. Hill. Prestack Gaussian-beam depth migration. Geophysics, 66:1240–1250, 2001. 203 [30] L. H¨ormander. On the existence and the regularity of solutions of linear pseudo- differential equations. L’Enseignement Mathematique, 17:99–163, 1971. [31] S. Leung and J. Qian. Eulerian Gaussian beam methods for Schr¨odinger equations in the semi-classical regime. J. Comput. Phys., 228:2951–2977, 2009. [32] S. Leung and J. Qian. The backward phase flow and FBI-transform-based J. Comput. Phys., Eulerian Gaussian beams for the Schr¨odinger equation. 229:8888–8917, 2010. [33] S. Leung, J. Qian, and R. Burridge. Eulerian Gaussian beams for high frequency wave propagation. Geophysics, 72:SM61–SM76, 2007. [34] S. Leung, J. Qian, and S. Serna. Fast Huygens Sweeping methods for Schrodinger equations in the semiclassical regime. Methods Appl. Analy., 21:031–066, 2014. [35] D. Ludwig. Uniform asymptotic expansions at a caustic. Commun. Pure Appl. Math., 19:215–250, 1966. [36] S. Luo and J. Qian. Factored singularities and high-order Lax-Friedrichs sweep- ing schemes for point-source traveltimes and amplitudes. J. Comput. Phys., 230:4742–4755, 2011. [37] S. Luo, J. Qian, and R. Burridge. Fast Huygens sweeping methods for Helmholtz equations in inhomogeneous media. submitted:xxx, 2012. [38] S. Luo, J. Qian, and R. Burridge. Fast Huygens sweeping methods for Helmholtz equations in inhomogeneous media in the high frequency regime. J. Comput. Phys., 270:378–401, 2014. [39] S. Luo, J. Qian, and H.-K. Zhao. Higher-order schemes for 3-D traveltimes and amplitudes. Geophysics, 77:T47–T56, 2012. [40] V. P. Maslov. The Complex WKB Method for Nonlinear Equations I: Linear theory. Birkhauser Verlag, Basel, 1994. [41] V. P. Maslov and M. V. Fedoriuk. Semi-classical approximation in quantum mechanics. D. Reidel Publishing Company, 1981. [42] M. Motamed and O. Runborg. Taylor expansion and discretization errors in Gaussian beam superposition. Wave Motion, 47:421–439, 2010. [43] D. Nichols. Imaging complex structures using band limited Green’s functions. PhD thesis, Stanford University, Stanford, CA94305, 1994. 204 [44] C. Nolan and W. W. Symes. Global solution of a linearized inverse problem for the wave equation. Commun. Partial Diff. Eq., 22:919–952, 1997. [45] M. M. Popov. A new method of computation of wave fields using Gaussian beams. Wave Motion, 4:85–97, 1982. [46] M.I. Protasov and Vladimir Cheverda. True amplitude imaging by inverse gen- eralized radon transform based on gaussian beam decomposition of the acoustic green’s function. Geophysical Prospecting, 59:197 – 209, 03 2011. [47] J. Qian and L. Ying. Fast Gaussian wavepacket transforms and Gaussian beams for the Schr¨odinger equation. J. Comput. Phys., 229:7848–7873, 2010. [48] J. Qian and L. Ying. Fast multiscale Gaussian wavepacket transforms and mul- tiscale Gaussian beams for the wave equation. SIAM J. Multiscale Modeling and Simulation, 8:1803–1837, 2010. [49] J. Ralston. Gaussian beams and the propagation of singularities. Studies in partial differential equations. MAA Studies in Mathematics, 23. Edited by W. Littman. pp.206-248., 1983. [50] H. Smith. A parametrix construction for wave equations with C1,1 coefficients. Ann. Inst. Fourier Grenoble, 48:797–835, 1998. [51] N. Tanushev, B. Engquist, and R. Tsai. Gaussian beam decomposition of high frequency wave fields. J. Comput. Phys., 228:8856–8871, 2009. [52] N. Tanushev, J. Qian, and J. Ralston. Mountain waves and Gaussian beams. SIAM J. Multiscale Modeling and Simulation, 6:688–709, 2007. [53] A. Toflove and S. C. Huganess. Computational Electrodynamics: The Finite Difference Time Domain Method, Second Editions. Artech House, Norwood, MA, 2000. [54] B. S. White. The stochastic caustic. SIAM J. Appl. Math., 44:127–149, 1984. [55] B. S. White, A. Norris, A. Bayliss, and R. Burridge. Some remarks on the Gaussian beam summation method. Geophys. J. R. Astr. Soc., 89:579–636, 1987. [56] Yu Zhang, Guanquan Zhang, and Norman Bleistein. True amplitude wave equa- Inverse tion migration arising from true amplitude one-way wave equations. Problems, 19(5):1113, 2003. [57] Yu Zhang, Guanquan Zhang, and Norman Bleistein. Theory of true-amplitude one-way wave equations and true-amplitude common-shot migration. 70, 07 2005. 205