NEAR-FIELD IMAGING VIA INVERSE SCATTERING By Junshan Lin A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Applied Mathematics 2011 ABSTRACT NEAR-FIELD IMAGING VIA INVERSE SCATTERING By Junshan Lin Near-field optics is an emerging research topic in the past few decades, mostly motivated by applications in the near-field microscopy in an effort to break the diffraction limit. In the far-field imaging, only the propagating wave components with spatial frequency below the wavenumber are available, and it is well-known that the resolution of the image is approximately λ/2 (diffraction limit) [17, 51, 53]. In the near field, however, the bandwidth of the spatial frequency may be expanded by taking account of the evanescent (exponentially decayed) waves. Nowadays there exist various configurations for the near-field microscopies, see for example [51]. However, it is recognized that the images obtained from the near-field microscopies are problematic by visualizing the object in an analogical way [18, 47]. Therefore, the inverse scattering theory is applied to understand how the structure of the scattering object is encoded in the measured scattered field. When single scattering (or Born approximation) is assumed, the studies are complete for the near-field scanning optical microscopy (NSOM) and the total internal reflection microscopy (TIRT) within the framework of the inverse scattering theory [19, 20, 57]. In this thesis, we focus on one specific problem where the imaging target is a ground plane with some local disturbance. The data is collected in the near-field regime with a distance above the surface displacement that is smaller than the wavelength. In the recent paper [29], a linearized model has been introduced for the nonlinear inverse scattering problem by the single scattering assumption. The authors also proposed a broadband imaging strategy for denoising and improving the resolution of the image. In the thesis, we investigate the more general case by considering the full scattering model, for which the linearized model in [29] is no longer valid. By the analysis of the scattered field, it is confirmed that the evanescent wave modes which are not accessible in the far-field regime become significant in the near field. Evanescent wave modes make it possible to break the diffraction limit. It is shown that such exponentially decayed modes of the scattered wave contain the high spatial frequency information (fine features) of the profile. We formulate explicitly the connection between the evanescent wave modes and the high frequency components of the surface displacement, and present a new numerical scheme to reconstruct the surface displacement from the boundary measurements. By extracting the information carried by the evanescent modes effectively, it is shown that the resolution of the reconstructed image is significantly improved in the near field. Numerical examples show that images with a resolution of λ/10 are obtained. To overcome the ill-posedness and the presence of local minima associated with this nonlinear imaging problem, we propose to use multiple frequency data to image the profile of the surface displacement in the second part of the thesis. The main idea is to march from the lowest wavenumber to the highest wavenumber. At the fixed wavenumber, by an analysis of the domain derivative for the forward scattering map, a vector field is chosen such that the defined cost functional decreases. The reconstructed profile evolves with the chosen vector field at the fixed wavenumber and the evolution process continues until it reaches the highest wavenumber. The proposed reconstructed scheme is able to capture the main feature of the profile at low frequency and recover the fine details at higher frequency. In particular, for a multiple scale profile, it resolves the fine scale with sufficiently high frequency. ACKNOWLEDGMENT I am deeply grateful to my thesis advisor Professor Gang Bao, whose continuous support and encouragements were invaluable for my four-year PhD studies at Michigan Sate University. He introduced to me the project on near-field imaging, and generously shared his wide expertise with me. His view on applied mathematics also helps me to shape my own understanding in the field, and will continue to influence me in the long run. I am also much obliged to him for helping me to improve my English writing skill, and for providing opportunities of attending conferences where I met the leading figures in the field and did presentations. I would like to thank Professors Zhengfang Zhou, Andrew Christlieb, Chichia Chiu, Di Liu and Jianliang Qian for serving in my thesis committee. I am so indebted to them and Professor Habib Ammari for providing letters of reference. I am also grateful to Professor Baisheng Yan and Barbara Miller for taking care of everything, in particular for their help during my first year study as a fresh PhD student. Many thanks also go to the research group here at Michgan State University. We have various interesting discussions and a lot of fun together. Thanks all the group members for sharing your expertise and happiness with me: Justin Droba, Guanghui Hu, Jun Lai, Peijun Li, Songting Luo, Yuanchang Sun, Yuliang Wang, Zhengfu Xu, KiHyun Yun, and Hai Zhang. Thanks in particular Yuanchang, Yuliang and Maureen for taking care of me when I first came to US in 2007. Thanks all my friends here, in particular Lianzhang Bao, Langhua Hu, Jie Niu, Xun Wang, Tianshuang Wu and Qiong Zheng. Finally, I would like to thank my parents and parents in law for their strong support. I owe my deepest gratitude to my wife Yu Chen for all the sacrifice she made. I can always feel her strong support behind me! iv TABLE OF CONTENTS List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 Introduction 1.1 Near-field 1.2 Near-field 1.3 Near-field . . . . . . . . . . . . . . Optics . . . . . . . . . . . . . Imaging Microscopy . . . . . . Imaging via Inverse Scattering . . . . . . . . . . . . . . . . . . . 2 Mathematical Theory of Inverse Problem . . . 2.1 Concepts of Ill-Posed Problem . . . . . . . . . . . 2.2 Singular Value Decomposition for Linear Compact 2.3 Regularization Methods for Linear Problems . . 2.4 Regularization Methods for Nonlinear Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Operators . . . . . . . . . . . . vii . . . . . . . . . . . . . . . . . . . 1 1 3 4 . . . . . . . . . . . . . . . . . . . . 7 7 8 10 13 . . . . 3 Near-field Imaging of the Surface Displacement on an Infinite Ground Plane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 3.1 Forward and Inverse Scattering Problems . . . . . . . . . . . . . . . . 15 3.2 Analysis of the Scattered Wave . . . . . . . . . . . . . . . . . . . . . 18 3.2.1 Layer Potential and Boundary Integral Equations . . . . . . . 18 3.2.2 Scattered Wave in Near-field Regime . . . . . . . . . . . . . . 23 3.3 Near-field Imaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 3.3.1 Inversion Method . . . . . . . . . . . . . . . . . . . . . . . . . 27 3.3.2 Fr´chet Differentiability of the Nonlinear Operator . . . . . . . 32 e 3.4 Numerical Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . 39 3.5 Discussions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44 4 Imaging with Multiple Frequency Data . . . . . . . 4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . 4.1.1 Problem Formulations . . . . . . . . . . . . . . . 4.1.2 Why Multiple Frequency Data . . . . . . . . . . . 4.2 Analysis of the Forward Scattering Problem . . . . . . . 4.2.1 Notations . . . . . . . . . . . . . . . . . . . . . . 4.2.2 Dirichlet-to-Neumann Map . . . . . . . . . . . . . 4.2.3 Well-posedness of the Forward Scattering Problem Domain . . . . . . . . . . . . . . . . . . . . . . . 4.3 Domain Derivative of the Forward Scattering Map . . . . 4.4 Imaging with Multiple Frequency Data . . . . . . . . . . 4.4.1 Descent Direction for the Cost Functional . . . . 4.4.2 Reconstruction Scheme . . . . . . . . . . . . . . . 4.5 Numerical Examples . . . . . . . . . . . . . . . . . . . . v . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . in Bounded . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45 45 45 47 47 47 49 51 53 59 59 61 64 4.5.1 4.5.2 Numerical approximation of the forward scattering problem . Imaging of the local surface displacement . . . . . . . . . . . . 64 65 5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . vi 78 LIST OF FIGURES Images in this dissertation are presented in color 1.1 Evanescent wave. For interpretation of the references to color in this and all other figures, the reader is referred to the electronic version of this thesis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 3.1 Setup of the problem. . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 3.2 eik2 (κ)d for d ∈ (0, 1.5λ) when the spatial frequency κ = 1.2k and 1.5k respectively . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 3.3 kc /k versus the distance varying from λ/10 to λ. . . . . . . . . . . . . 29 3.4 Real part (left) and the imaginary part (right) of the scattered field. . 40 3.5 Near-field (top) and far-field (bottom) images. The solid line represents the real image, and the dotted line is the reconstruction. . . . . . . . 41 3.6 Relative error with respect to Newton iterations. . . . . . . . . . . . . 41 3.7 Near-field (left) and far-field (right) images. The solid line represents the real image, and the dotted line is the reconstruction. . . . . . . . 42 3.8 Real profile, the distance between neighboring bumps is λ/10. . . . . 42 3.9 Comparison of the reconstruction (dotted line) with the real profile (solid line) for the near-field case. . . . . . . . . . . . . . . . . . . . . 43 Problem geometry. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46 4.1 vii 4.2 Top: real profile. Bottom: reconstruction (dash line) at k = 16 compared with the real curve (solid line). . . . . . . . . . . . . . . . . . 67 4.3 Relative error with respect to the wavenumber. . . . . . . . . . . . . 68 4.4 Evolution of the reconstruction at k = 1, 5, 8, 10, 12, 16 (dash line, left to right, top to bottom). . . . . . . . . . . . . . . . . . . . . . . . . . 69 Top: relative residual with respect to the iteration number at k = kM when only single frequency data is used. Bottom: reconstruction (dash line) with single frequency only. . . . . . . . . . . . . . . . . . . . . 70 Reconstruction (dash line) at k = 30 compared with the real curve (solid line). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 4.7 Relative error with respect to the wavenumber. . . . . . . . . . . . . 72 4.8 Reconstruction (dash line) at k = 16 compared with the real curve (solid line). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72 Reconstruction at k = 10, 20, 33, 40 (dash line, top to bottom) compared with the real curve (solid line). . . . . . . . . . . . . . . . . . . 73 4.5 4.6 4.9 viii Chapter 1 Introduction 1.1 Near-field Optics Near-field optics is the study of the evanescent wave fields and their interactions with matter on a sub-wavelength scale. Usually the evanescent fields are localized to the optical source region or the surface of the scattering object, and the study of the nearfield optics is mostly concerned with the localized region within one wavelength. The modern interests on the topic was mainly motivated by its applications in near-field imaging microscopies. We refer the reader to [51, 62] and the references therein for detailed discussions. The study of the near-field optics has its origin in an effort to break the diffraction limit imposed by the far-field imaging. In far-field optics, the cut-off of the spatial spectrum is very strict: only the propagating wave components with spatial frequency below the wavenumber can be used. The loss of higher spatial frequencies leads to the diffraction limit, which is also known as the Rayleigh resolution limit. At the end of the nineteenth century, Abbe and Rayleigh [1, 53] derived a criterion for this limit. The minimum distance ∆x between two point sources at which they can still 1 Figure 1.1: Evanescent wave. For interpretation of the references to color in this and all other figures, the reader is referred to the electronic version of this thesis. be unambiguously distinguished as two separate sources reads as ∆x = 0.61λ/N A, where N A is the numerical aperture, and the best possible N A for optical glasses is roughly 1.5. Therefore, the spatial resolution for a far-field optical microscopy is approximately λ . In near-field optics, however, the bandwidth of spatial frequency 2 is expanded by taking account of the evanescent waves. The evanescent waves can be described by plane waves of the form ei(k·x−ωt) , where at least one component of the wave vector k describing the direction of propagation is imaginary. They play a central role in near-field optics. In the spatial direction defined by the imaginary component of k, the wave no longer propagates, but decays exponentially. Figure 1.1 is the plot of an evanescent wave that propagates in the xy plane, but decays exponentially in the z direction. In fact, the total internal reflection at the surface of a dielectric medium generates 2 such an evanescent wave. Let us consider a plane wave impinging on a flat surface between two media characterized by the refraction index n1 and n2 respectively and n1 > n2 . By the boundary conditions on the interface, the wave vectors k1 and k2 in the two media take the form of k1 = (kx , ky , kz1 ), k2 = (kx , ky , kz2 ), n ω |k1 | = 1 , c n2 ω . |k2 | = c Here c is the speed the light in the vacuum. If the incident angle θi is the larger than n the critical angle θc = arcsin n2 , then kz2 becomes imaginary, and the exponential 1 n1 2 2 decay constant is |k2 | ( ) sin θi − 1. n2 The evanescent waves are always restricted to the surfaces of imaging objects, thus the study of the optical interactions on a subwavelength scale becomes very significant. This usually complicates the analysis as well as the computation — the price we have to pay for the inclusion of evanescent waves to obtain images with better spatial resolution. 1.2 Near-field Imaging Microscopy The central idea of near-field microscopies is to retain the spatial frequencies associated with evanescent waves, thereby increasing the bandwidth of spatial frequencies. In principle, arbitrary resolution can be achieved provided the bandwidth is infinite. However, this is at the expense of strong coupling between the source and the imaging object. Here we briefly summarize several existing near-field imaging modalities, and refer to the monograph [51] and [30, 25, 26, 54] for references. The list is of course by no means to be complete. Near-field scanning optical microscopy (NSOM) [51, 25, 30] applies the excitation 3 beam emanating from a tiny aperture (e.g., pointed optical fiber) or a tiny metal tip. The electric field distribution changes significantly as the aperture size becomes smaller. In particular, when the aperture size is below the wavelength, the z component of the wave vector becomes imaginary. The strong localization of such evanescent waves yields the subwavelength resolution that can be achieved in near-field scanning optical microscopy. The scanning tunneling optical microscope (STOM) [26], also called the phonon scanning tunneling microscope (PSTM) [54] uses far-field illumination and near-field detection. To illuminate the sample, a laser beam undergoes total internal reflection at the surface of the sample support. A bare tapered glass fiber is dipped into this evanescent field to locally couple some of the light into the probe where it is converted into propagating modes that are guided towards a detector. There are other imaging modalities such as the near-field illumination and nearfield detection configuration which makes use of near-field interactions for both excitation and detection, and the energy-transfer microscopy where the near-field interaction between probe and sample is achieved through dipole-dipole coupling. There are new methods being developed continuously. Apparently large diversity of methods are usually categorized according to their specific illumination and detection conditions, and in practice, it is usually desirable to have different specialized modalities combined together to give more efficient imaging. 1.3 Near-field Imaging via Inverse Scattering In an optical microscopy, the object is usually visualized in an analogical way with little or no numerical treatment. Therefore, the images of the scattering object is usually problematic [18, 47]. There is an alternative type of near-field imaging technique which relies entirely on a numerical inversion procedure to reconstruct the sample 4 from the data of the scattered field, i.e., to solve the inverse scattering problem. In [15], M. Bertero, P. Boccacci and M. Piana discussed the concept of resolution for the inverse diffraction problems that arise in the acoustic holography. They showed that the super-resolution is possible due to the information conveyed by evanescent waves in near-field. P. Carney and J. Schotland also studied the inverse scattering problem for the near-field microscopy and three-dimensional total internal reflection microscopy (TIRM). They also showed subwavelength spatial resolution [19, 20]. Their framework on the near-field imaging is all based on the weak-scattering approximation (or Born approximation). More recently, A. Sentenac et al carried out the whole nonlinear inversion process for the total internal reflection microscopy [5, 6]. Unfortunately, in optics the highest refractive index available for the prism function in TIRM is close to two, which limits the spatial frequency of illuminating field. In order to increase the spatial frequencies of the illuminating field beyond that reachable with a prism, they proposed to deposit the sample on an optimized grating [21]. This is the so called grating-assisted optical diffraction tomography. The starting point of this thesis is also based on the idea of imaging by applying the inverse scattering theory. We study the inverse diffraction problem for an unbounded obstacle which is a ground plane with some local disturbance. New numerical schemes are developed to reconstruct the surface displacement from the boundary measurements. In Chapter 2, we briefly summarize the mathematical theories of inverse problems, which serves as the building blocks for the study of the inverse scattering problem. Chapter 3 is devoted to the analysis the scattered field in near-field regime, and the design a numerical method that makes use of the evanescent modes effectively to improve the resolution of the near-field image. Most of Chapter 3 comes from [10]. 5 In Chapter 4, we use multiple frequency data to image the profile of the surface displacement. A stable and accurate reconstruction method is presented and investigated with numerical simulations. In particular, for a multiple scale profile, the proposed method resolves the fine scales with sufficiently high frequency information. Chapter 4 is mostly extracted from [11]. 6 Chapter 2 Mathematical Theory of Inverse Problem In this chapter, we briefly summarize the mathematical theories for the inverse problems, with focus on the linear compact operators in Hilbert spaces. It is intended as an introduction to the basic ideas on the ill-posed problems and regularization methods for their solutions. We refer to the monographs [31, 43, 24, 37] and references therein for detailed discussions. 2.1 Concepts of Ill-Posed Problem In his lecture published in [38], Hadamard claims that a mathematical model for a physical problem has to be properly posed or well-posed in the sense that it has following properties: • There exists a solution of the problem (existence). • There exists at most one solution of the problem (uniqueness). • The solution depends continuously on the data (stability). 7 If any of the above three criteria is violated, the problem is called ill-posed. In practice, the instability is one of the primary interest in the study of ill-posed problems. The violations of stability always creates serious numerical problems: an infinitesimal noise in the measurement will give rise to large errors in the solution. No mathematical trick can make an inherently unstable problem stable. A remedy is to use the regularization methods. But all that a regularization method can do is to recover partial information of the solution as stable as possible. Consider the operator equation Kx = y, (2.1) where K is an operator between Hilbert spaces X and Y . The typical example of an ill-posed problem is when K is a compact operator, since the inverse K −1 is unbounded. We will base our exposition on (2.1) in the following sections. 2.2 Singular Value Decomposition for Linear Compact Operators A better way to understand the structure of linear compact operator K (X → Y ) is from the spectral theory. For an self-adjoint operator K, all of its eigenvalues are real. Moreover, K has at least one but at most a countable number of eigenvalues with 0 as the only possible accumulation point. Assume that the sequences {λn }∞ of the nonzero eigenvalues is ordered such n=1 that |λ1 | ≥ |λ2 | ≥ |λ3 | ≥ · · · 8 Let xn be a sequence of corresponding orthonormal eigenvectors. Then for x ∈ X, ∞ λn (x, xn )xn . Kx = n=1 For a proof of this spectral decomposition for self adjoint operators, see for example [45]. The spectral theorem for compact self-adjoint operators has an extension to nonselfadjoint operators. If K : X → Y is a linear compact operator, then its adjoint operator K ∗ : Y → X is also compact. We call the nonnegative square roots of the eigenvalues for the self-adjoint compact operator K ∗ K the singular values. Theorem 2.2.1. (Singular Value Decomposition) Let K : X → Y be a linear compact operator, K ∗ : Y → X be its adjoint operator, and µ1 ≥ µ2 ≥ µ3 · · · > 0 be ordered sequences of the positive singular values of K. Then there exist orthonormal systems {xn }∞ ⊂ X and {yn }∞ ⊂ Y with the following properties: n=1 n=1 Kxn = µn yn and K ∗ yn = µn xn for all n ∈ N. The system (µn , xn , yn ) is called a singular system for K. For each x ∈ X, there exists a singular value decompostion ∞ x = x0 + (x, xn )xn n=1 for some x0 ∈ Ker(K) and ∞ Kx = µn (x, xn )yn . n=1 The following theorem expresses the solution to the equation Kx = y in terms of the singular system. 9 Theorem 2.2.2. (Picard) Let K : X → Y be a linear compact operator with the singular system (µn , xn , yn ). The solution to Kx = y is solvable if and only if f ∈ (Ker(K ∗ ))⊥ and ∞ 1 |(y, yn )|2 < ∞. 2 n=1 µn In this case ∞ x= n=1 1 (y, yn )xn . µn Picard’s theorem demonstrate the ill-posed nature of the equation Kx = y. If we perturb the right hand side by δy, the perturbation of the solution x can be made arbitrarily large due to the fact that the singular values tend to zero. 2.3 Regularization Methods for Linear Problems Since the inverse of the operator K is not bounded, one remedy is to use the regularization methods. A regularization strategy is a family of linear and bounded operators Rα : Y → X, α>0 such that lim Rα Kx = x for all x ∈ X. α→0 Let y δ be the measured data (with error) with y − y δ ≤ δ, then xα,δ := Rα y δ 10 is an approximation of the solution of (2.1). The error x − xδ can be split into two parts: δ Rα and Rα Kx − x , where the first term described the error in the data multiplied by Rα and the second term denotes the approximation error (Rα − K −1 )y . The art of choose the regularization parameter α will always be to find the right compromise between accuracy and stability. A convenient way to construct admissible regularization strategies is given by filtering singular value systems. The idea of using filters has a long history and is very convenient for theoretical purposes [36, 58]. Let (µn , xn , yn ) be a singular system for K. The regularization strategies is constructed by damping the factors 1/µn . Theorem 2.3.1. (Regularization by damping [43]) Let K : X → Y be a linear compact operator with singular system (µn , xn , yn ). q : (0, ∞) × (0, K ) → R is a function with the following properties: • |q(α, µ)| ≤ 1 for all α > 0 and 0 < µ < K . • For each α > 0 there exists c(α) such that |q(α, µ)| ≤ c(α)µ and 0<µ≤ K . • limα→0 q(α, µ) = 1 for each 0 < µ < K . Then the operator Rα : Y → X, defined by ∞ x = Rα y = n=1 q(α, µn ) (y, yn )xn µn 11 is a regularization strategy with Rα ≤ c(α). The function q is called a regularizing filter for K. There are various choices of the function q that satisfies the above properties. Typically, we list three filter functions as follows: (1) q is defined by    1, µ2 ≥ α, q(α, µ) =  0, µ2 < α.  This choice of q is also known as the spectral cutoff. (2) q(α, µ) = µ2 /(α + µ2 ). This choice of the filter function q is equivalent to the Tikhonov regularization strategy [59, 60]. The Tikhonov regularization tries to minimize the functional Jα (x) := Kx − y + α x , and the minimum xα of the Tikhonov functional is the unique solution of the normal equation αxα + K ∗ Kxα = K ∗ y. It is clear that the solution xα can be written in the form xα = Rα y with Rα := (αI + K ∗ K)−1 K ∗ = ∞ µn 2 n=1 α + µn (·, yn )xn . (3) q(α, µ) = 1 − (1 − aµ2 )1/α . This choice of the filter function q leads to the so-called Landweber iteration method [16, 35, 48]. The iteration scheme is the steepest descent algorithm for the quadratic functional x → Kx − y 2 , where 12 xm is computed recursively by xm = xm−1 − aK ∗ (Kxm−1 − y), m = 1, 2, · · · The resulting regularized operator has the form ∞ 1 − (1 − aµ2 )m ∗ K)j K ∗ = Rm := a (I − aK (·, yn )xn . µn n=1 j=0 There are many other regularization strategies for the solution of the linear illposed problems, for example, the conjugate gradient method [27], the regularization by discretization [14, 45], etc. One important issue for the regularization strategy is the choice of the regularization parameter α such that the accuracy and stability are balanced. In practice, usually the posteriori parameter choice rules are applied. One of the most widely used strategies is the so-called discrepancy principle due to Morozov [49]. We also refer to [31, 43] for the discussions of other posteriori rules to choose the regularization parameter. 2.4 Regularization Methods for Nonlinear Problems Let F be a nonlinear operator between Hilbert spaces X and Y , and we want to solve F (x) = y. (2.2) The regularization theory for the nonlinear ill-posed problem (2.2) is by far not so well developed as in the linear case. 13 As in the linear case, the Tikhonov regularization minimizes the functional F (x) − y + α x . (2.3) The minimization problem (2.3) admits a solution, though the solution may not be unique in general. Moreover, the Tikhonov regularization strategy is stable in the sense of continuous dependence of the solution on the data y [31, 55]. If y in (2.3) is replaced by the perturbed data y δ , then the corresponding solution xα,δ of (2.3) converges to the minimum norm solution of (2.2) for general nonlinear problems when δ → 0[55]. The rates of convergence are presented in [32, 50]. The nonlinear Landweber iteration to the nonlinear ill-posed problem updates the solution along the steepest descent direction of the functional F (x) − y : xm = xm−1 − aF (xm−1 )∗ (F (xm−1 ) − y), m = 1, 2, · · · For nonlinear problems, the iteration will not have a global convergence property. However, under suitable condition on the nonlinear operator F with an appropriate stopping rule, the convergence is obtained in [39]. There are also Newton’s type methods for solving (2.2). The main idea is to repeatedly linearize the nonlinear operator equation by solving F (xm−1 )(x − xm−1 ) = y − F (xm−1 ). Usually F (xm−1 ) is compact, and regularization strategy is applied when solving the linearized problem. We refer to [31] for detailed discussions. 14 Chapter 3 Near-field Imaging of the Surface Displacement on an Infinite Ground Plane 3.1 Forward and Inverse Scattering Problems Consider the scattering of the time harmonic electromagnetic wave that impinges on an unbounded obstacle. The boundary of the obstacle is assumed to be a perturbation Figure 3.1: Setup of the problem. 15 of the x1 x3 plane. It is also assumed that the surface displacement is invariant along the x3 direction, and is local with respect to x1 . We further restrict the study to the TM polarization or E-parallel case, i.e., the electric field E = (0, 0, u(x1 , x2 )). Consequently, the Maxwell equations are reduced to the two dimensional Helmholtz equation. Before introducing the forward scattering model, we describe the geometry of the obstacle shown in Figure 3.1. Let γ ⊂ R be bounded, open, and ∂γ be the boundary of γ. Denote the closure of γ as γ. The ground plane Γ0 := R\γ, and the local surface displacement is represented by Γ := {x = (x1 , x2 ) | x1 ∈ γ, x2 = f (x1 )}, where the function f is defined on γ: f (x1 ) > 0 for x1 ∈ γ, f (x1 ) = 0 for x1 ∈ ∂γ. By requiring f > 0 on γ, the surface displacement is directed upward. Clearly, ∂D := Γ∪Γ0 is the boundary of the whole unbounded obstacle on which the electromagnetic wave impinges. The domain above ∂D is denoted as D (⊂ R2 ). The incident wave field ui = eikq·x is a plane wave that propagates along the direction q = (sin θ, − cos θ)T , where θ is the incident angle and k = ω is the c wavenumber. Here ω is the angular frequency, and c is the speed of the wave propagating in the vacuum. Let λ = 2π denote the wavelength. If the obstacle is a flat k perfect conductor, then the reflected field ur = −eikq ·x produced by the flat surface is a plane wave propagating along the direction q = (sin θ, cos θ)T . In general, the total field ut from the scattering by Γ ∪ Γ0 consists of three parts: the incident wave ui , the reflected wave ur , and the scattered field. The scattered field u satisfies the Helmholtz equation ∆u + k 2 u = 0 in D. (3.1) Assuming that the obstacle is a perfect conductor, the total field vanishes on the 16 boundary. Hence u = −(ui + ur ) on ∂D. (3.2) It is easily seen that u = 0 on Γ0 . Moreover, the scattered field satisfies the Sommerfield radiation condition: lim r→∞ √ r ∂u − iku ∂r = 0, r = |x| . (3.3) The forward scattering problem (4.1)-(4.3) admits a unique solution u ∈ C 2 (D) ∩ ¯ C(D) if ∂D is C 2 and boundary data u|∂D is continuous [61]. Clearly here the plane wave ui , ur ∈ C ∞ (R2 ). It is worth mentioning that there are many results on related scattering problems in the literature. The well-posedness of the scattering problem for an obstacle with locally downward surface displacement (f (x1 ) < 0 for x1 ∈ γ) was studied in [3, 4]. There are also general studies on the scattering by a non-local perturbed half plane. See for example [64] and the references therein. In our framework for the inverse scattering problem, data is collected on the line x2 = d above the surface displacement with a distance that is smaller than the wavelength λ (near-field regime). To be more precise, it is required that 0 < d − maxx ∈¯ f (x1 ) < λ. The inverse problem is to reconstruct f from the scattered 1 γ field u(·, d) collected on the line x2 = d. Our work is originally motivated by the recent paper [29], in which a linearized model has been introduced for the nonlinear inverse scattering problem by the single scattering assumption. The authors also proposed a broadband imaging strategy for denoising and improving the resolution f 1 and the modulus of the image. However, this linearized model is valid only if λ of its derivative f 1 simultaneously. Here we investigate the more general case by considering the full scattering model, for which the linearized model in [29] is no longer valid. For the reconstruction of the star-like local disturbance from the far field pattern of the scattered field, we refer to [46]. 17 This imaging problem shares many of the well-known difficulties with other inverse boundary value problems, particularly nonlinearity and ill-posedness. However, by collecting data in the near-field regime, the evanescent wave modes which are not accessible in the far-field regime (d − max f λ) become significant. This crucial fact may be confirmed by the analysis of the scattered field in Section 3.2. Evanescent wave modes make it possible to break the diffraction limit. It is shown that such exponentially decayed modes of the scattered wave contain exactly the high spatial frequency information (fine features) of the profile f . Our study is to analyze the scattered field carefully, and design a numerical method that makes use of the evanescent modes effectively, thus to improve the resolution of the image. Numerical examples confirm that a resolution of λ/10 is obtained in the near field. 3.2 Analysis of the Scattered Wave 3.2.1 Layer Potential and Boundary Integral Equations Introduce Green’s function G(x, y) := Φ(x, y) − Φ(xr , y), i (1) H (k|x − y|) is the fundamental solution for the Helmholtz 4 0 equation in R2 and xr is the reflection of x by the x1 axis, i.e., xr = (x1 , −x2 ). where Φ(x, y) = Denote Γ = Γ ∪ ∂γ, Γr := { (x1 , −x2 ) | x ∈ Γ }, Γ ∪ Γr = Γ ∪ Γr ∪ ∂γ, and Dr := { (x1 , −x2 ) | x ∈ D }. For a function ψ ∈ C(Γ), we define the single layer potential: u(x) = Γ G(x, y)ψ(y)dsy , x ∈ R2 \Γ ∪ Γr . The following lemma is concerned with the limit of the normal derivative of the single 18 layer potential, when it is extended from above and below the boundary Γ. The limit for the case when Γ is the smooth boundary of a bounded obstacle is well known [23, 24]. Here Γ is not a closed curve. For completeness, the proof of the lemma is provided. Lemma 3.2.1. Assume that Γ is C 2 . For the single layer potential with continuous density ψ, the following holds: ∂G(x, y) ∂u (x) = ψ(y)dsy ∂ν ± Γ ∂νx where ν is the unit normal directed into D, hν(x)). 1 ψ(x), 2 x ∈ Γ, ∂u (x) := limh→0+ ν(x) · ∂ν ± u(x ± Proof. For x ∈ Γ, > 0, denote Γx, := { y ∈ Γ, |y − x| < }. We first show that lim lim →0 h→0+ Γx, ∂Φ(x + hνx , y) 1 dsy = − . ∂νx 2 Let ∂Ω (the boundary of some bounded connected domain Ω) be a C 2 closed curve such that Γ ⊂ ∂Ω. Moreover, for x ∈ Γ, the unit outward normal of x on ∂Ω coincides with νx . Let ψ ≡ 1 on ∂Ω , from the classical results in [23, 24], ∂u ∂Φ(x, y) 1 (x) = dsy − , ∂ν + 2 ∂Ω ∂νx x ∈ Γ. (3.4) On the other hand, for any small fixed number > 0, lim h→0+ ∂Ω ∂Φ(x + hνx , y) dsy = lim ∂νx h→0+ = + ∂Φ(x + hνx , y) dsy ∂νx Γx, ∂Ω\Γx, ∂Φ(x, y) ∂Φ(x, y) dsy + lim dsy . ∂νx ∂νx ∂Ω\Γx, h→0+ Γx, 19 Now letting → 0, we get lim h→0+ ∂Ω ∂Φ(x, y) ∂Φ(x + hνx , y) ∂Φ(x, y) dsy = dsy + lim lim dsy . ∂νx ∂νx →0 h→0+ Γx, ∂Ω ∂νx (3.5) From (3.4), (3.5), we have lim lim →0 h→0+ Γx, ∂Φ(x + hνx , y) 1 dsy = − . ∂νx 2 (3.6) . For the integral on Γ, for any fixed small number > 0, lim h→0+ Γ ∂Φ(x + hνx , y) ψ(y)dsy ∂νx ∂Φ(x + hνx , y) ∂Φ(x + hνx , y) ψ(y)dsy + lim dsy ψ(x) ∂νx ∂νx h→0+ Γx, ∂Φ(x + hνx , y) (ψ(y) − ψ(x))dsy + lim ∂νx h→0+ Γx, ∂Φ(x, y) ∂Φ(x + hνx , y) = ψ(y)dsy + lim dsy ψ(x) ∂νx ∂νx Γ\Γx, h→0+ Γx, = lim h→0+ Γ\Γx, ∂Φ(x + hνx , y) (ψ(y) − ψ(x))dsy . ∂νx + lim h→0+ Γx, On the other hand, there exists some constant M that ∂Φ(x + hνx , y) dsy ≤ M ∂νx Γx, and |ψ(y) − ψ(x)| → 0, as → 0 for y ∈ Γx, . Therefore, by letting → 0 and noting (3.6), we obtain lim h→0+ Γ ∂Φ(x + hνx , y) ∂Φ(x, y) 1 ψ(y)dsy = ψ(y) − ψ(x). ∂νx 2 Γ ∂νx 20 For x ∈ Γ, lim h→0+ Γ ∂Φ(xr + hνx , y) ∂Φ(xr , y) ψ(y)dsy = ψ(y). ∂νx ∂νx Γ By combining the above, we have ∂u ∂G(x, y) 1 (x) = ψ(y)dsy − ψ(x), ∂ν + 2 Γ ∂νx The proof for x ∈ Γ. ∂u can be carried out in the same fashion. ∂ν − The following representation result may serve as a starting point for the analysis of the scattered field. Lemma 3.2.2. If ∂D is C 3 , then there exists ψ ∈ C(Γ) such that the solution to (4.1)-(4.3) can be expressed as the single layer potential u(x) = Γ G(x, y)ψ(y)dsy x ∈ D. (3.7) Proof. Let u be the solution of (4.1)-(4.3). Note u = 0 on Γ0 , by Green’s theorem and the radiation condition, the scattered field takes the following form u(x) = ∂G(x, y) ∂u(y) u(y) − G(x, y) dsy , ∂νy Γ ∂νy x ∈ D. (3.8) ˜ Denote the domain bounded by Γ ∪ Γr as D. Let u0 := ui + ur . Then u0 ∈ C ∞ (R2 ) ˜ satisfies the Helmholtz equation in D. By Green’s theorem, we have Γ + Γr ∂u (y) ∂Φ(x, y) u0 (y) − Φ(x, y) 0 ∂νy ∂νy 21 dsy = 0, x ∈ D, (3.9) where νy is the unit normal directed to D for y ∈ Γ, νy is the unit normal directed to Dr for y ∈ Γr . By noting that u0 (y) = −u0 (yr ), and ∂u (yr ) ∂u0 (y) =− 0 for y ∈ Γ, we get ∂νy ∂νyr ∂Φ(x, y) u0 (y)dsy = Γr ∂νy ∂u (y) Φ(x, y) 0 dsy = ∂νy Γr ∂Φ(xr , y) (−u0 (y))dsy , ∂νy Γ ∂u (y) Φ(xr , y)(− 0 )dsy . ∂νy Γ (3.10) Substituting (3.10) into (3.9) yields ∂u (y) ∂G(x, y) u0 (y) − G(x, y) 0 dsy = 0, ∂νy Γ ∂νy x ∈ D. (3.11) A combination of (3.8) and (3.11) leads to ∂u ∂u ∂G(x, y) (u + u0 )(y) − G(x, y) + 0 ∂νy ∂νy Γ ∂νy ∂u ∂u = − G(x, y) + 0 (y)dsy . x ∈ D. ∂νy ∂νy Γ u(x) = (y)dsy , ∂u ∂u + 0 . ψ ∈ C(Γ) follows by the standard regularization theory ∂νy ∂νy for second-order elliptic equations and the Sobolev imbedding theorems [34]. The Let ψ = − proof is now complete. Introduce the integral operator K : C(Γ) → C(Γ), (Kψ)(x) := Γ G(x, y)ψ(y)dsy , x ∈ Γ. (3.12) Denote g = −(ui + ur )| . Then the existence of the solution to the integral equation Γ Kψ = g follows from Lemma 3.2.2. Note that since the single layer potential can be continuously extended to Γ, the density function ψ defined in Lemma 3.2.2 is a 22 solution of the integral equation Kψ = g. Regarding the uniqueness, we have the following result. Proposition 3.2.3. If ∂D is C 3 , then Kψ = g is uniquely solvable if k 2 is not the ˜ ˜ eigenvalue of −∆ in D for the Dirichlet problem. Here D is the domain bounded by Γ ∪ Γr . Proof. If Kψ = 0 for some ψ ∈ C(Γ), then the single layer potential u(x) = Γ G(x, y)ψ(y)dsy solves the exterior problem   ∆u + k 2 u = 0   1 1   u =0  1   √ ∂u1  lim  − iku1 r→∞ r ∂r in D, on ∂D, (3.13) =0 and the interior problem    ∆u + k 2 u = 0 ˜ in D, 2 2  u =0  2 on Γ ∪ Γr . (3.14) ˜ If k 2 is not the eigenvalue of −∆ in D for the Dirichlet problem, then (3.13) and ∂u1 ∂u (3.14) attains a unique solution respectively. Hence − 2 = 0. By the jump ∂v ∂v ∂u1 ∂u2 condition in Lemma 3.2.1 we have −ψ = − = 0 on Γ. Note that ψ ∈ C(Γ), ∂v ∂v thus ψ ≡ 0 on Γ. 3.2.2 Scattered Wave in Near-field Regime Now we are ready to examine the scattered field in the near-field regime. We point out that the exponentially decayed (evanescent) wave modes which are localized to 23 the surface the local disturbance are significant in the near field, and formulate explicitly the connection between the evanescent wave modes and the high frequency components for the profile of the local disturbance. For convenience, the Green’s function G(x, y) and the fundamental solution Φ(x, y) are written with an explicit dependence on the first variable: ˜ ˜ ˜ G(x, y) = G(x1 − y1 ; x2 , y2 ) := Φ(x1 − y1 , x2 − y2 ) − Φ(x1 − y1 , x2 + y2 ) , i (1) ˜ where Φ(x1 , x2 ) = H0 (k x2 + x2 ). 1 2 4 ˜ For the free space fundamental solution Φ(x1 − y1 , x2 − y2 ), we have the following plane wave decomposition: 1 i(x −y )·κ ik (κ)|x −y | i ˜ 2 2 dκ, e 1 1 e 2 Φ(x1 − y1 , x2 − y2 ) = 4π R k2 (κ) where    k2 (κ) =  i  k 2 − |κ|2 |κ| < k (propagating modes), |κ|2 − k 2 |κ| > k (evanescent modes). (3.15) (3.16) The decomposition may be viewed as the sum of the plane waves that consist of propagating and evanescent modes. It is clear that all the wave modes propagate along the x1 direction. Along the x2 direction, when the magnitude of the spatial frequency |κ| is below k, the wave mode also propagates; otherwise, it decays exponentially along the x2 direction and is denoted as the evanescent mode. ˜ A similar plane wave decomposition holds for G(x1 − y1 ; x2 , y2 ) evaluated at x2 = d: 1 i(x −y )·κ ik (κ)|d−y | i ˜ 2 − eik2 (κ)|d+y2 | dκ. e 1 1 e 2 G(x1 − y1 ; d, y2 ) = 4π R k2 (κ) (3.17) Thus for the scattered field at x = (x1 , d), by noting the single layer potential (3.7), 24 some simple calculations yield i 1 i(x −y )·κ ik (κ)|d−f (y )| 1 − eik2 (κ)|d+f (y1 )| e 1 1 e 2 4π R γ k2 (κ) ψ(y1 , f (y1 )) J dy1 dκ, u(x1 , d) = where J = 1 + f 2 . This implies that the measured scattered field u on the line x2 = d can also be viewed as the superposition of the propagating and evanescent wave modes. It is important to note that the evanescent modes with spatial frequency beyond the wavenumber k decay exponentially along the x2 direction, and are localized to the surface of the obstacle within one wavelength. Therefore, in the far-field regime, such evanescent modes carried by the scattered field are lost. However, in the near-field regime, the evanescent modes are significant, and the measured scattered field carries more information for the profile of the local disturbance to be reconstructed. ˜ On the other hand, following from the Taylor expansion of G(x1 − y1 ; d, y2 ) at y2 = 0, the scattered field can also be expanded as u(x1 , d) = ˜ ˜ ∂ 2 G(x1 − y1 ; d, 0) ∂ G(x1 − y1 ; d, 0) f (y1 ) + (f (y1 ))2 2 ∂y2 ∂y2 γ +O(f 3 )] ψ(y1 , f (y1 )) J dy1 . ˜ [G(x1 − y1 ; d, 0) + where ˜ ∂ 2 G(x1 − y1 ; d, 0) ˜ ˜ = 0 by the symmetry property of G; (a) G(x1 − y1 ; d, 0) = 2 ∂y2 (b) for the high order term, by a direct calculation, asymptotically f O(f 3 ) = O ( )2 λ We assume that f 2 λ ∂G(x1 − y1 ; d, 0) f. ∂y2 1, and denote ϕ = ψ 25 1 + f 2 . Then the scattered field is simplified as u(x1 , d) ≈ = ˜ ∂ G(x1 − y1 ; d, 0) f (y1 )ϕ(y1 )dy1 ∂y2 γ ˜ ∂ G(x1 − y1 ; d, 0) f (y1 )ϕ(y1 )dy1 . ∂y2 R (3.18) The equality holds since the surface displacement is supported on γ. Here we implicitly extend the definition of f and ϕ to R by setting f as 0 for x1 ∈ R \ γ. On the other hand, from (3.17) a simple calculation yields ˜ ∂ G(x1 − y1 ; d, 0) 1 ei(x1 −y1 )·κ eik2 (κ)d dκ. = ∂y2 2π R Therefore, by taking the Fourier transform of (3.18), we arrive at u(κ, d) ≈ eik2 (κ)d (f ϕ)(κ), ˆ κ∈R, (3.19) where · denotes the Fourier transform. Remark 2.1. The expression (3.19) formulates explicitly the connection between the evanescent wave modes and the high frequency components of the profile f . It indicates that a high spatial frequency mode of the scattered field u(κ, d) carries high ˆ spatial frequency information (fine features) of f to be reconstructed. f 1 and the modulus of its derivative f λ reduced automatically to the linear model discussed in [29]. Remark 2.2. If 1, then (3.19) is Now we distinguish the far-field and the near-field cases based on (3.19). When the spatial frequency |κ| > k, eik2 (κ)d decreases exponentially with respect to the distance d and its value vanishes when d exceeds one wavelength (see Figure 2). Thus in the far-field regime (d λ), u(κ, d) ≈ 0 for |κ| > k, i.e., the high spatial frequency ˆ information of f is lost in the far-field measurement. In the context of imaging, this 26 1 κ=1.2 k κ=1.5 k 0.8 0.6 0.4 0.2 0 0.5 d/λ 1 1.5 Figure 3.2: eik2 (κ)d for d ∈ (0, 1.5λ) when the spatial frequency κ = 1.2k and 1.5k respectively . implies that it is impossible to recover f with very high resolution when any noise is present. However, in the near-field regime with d < λ, eik2 (κ)d is not close to 0 and the exponentially decayed modes are still significant in the scattered field. Therefore, the higher spatial frequency components of f can still be retrieved by inverting the evanescent modes of the scattered field. 3.3 3.3.1 Near-field Imaging Inversion Method Assume that the measurement u(·, d) is polluted with some additive noise n(x1 ), which takes the following form n(x1 ) = σ · rand(x1 ) · u(x1 , d). 27 σ is the noise level, and rand(x1 ) is a uniformly distributed random variable in [−1, 1] for each x1 ∈ R. Moreover, rand(x1 ) is mutually independent for different values of x1 . Based on the previous analysis, we present a reconstruction method. From (3.19), we introduce the pseudo-inverse operator Id as follows:   −ik2 (κ)d  e |κ| ≤ kc , Id (κ) =  0  |κ| > kc , (3.20) where Id (κ) is a cut-off regularized operator, and kc is a regularization parameter. In the far-field case, as we discuss in the previous section, only the propagating modes can be used for imaging if noise is present, thus the cutoff frequency kc = k. In the near-field regime, the bandwidth of the spatial frequency is expanded beyond the wavenumber k by taking account of the evanescent waves. Note that e−ik2 (κ)d is an exponentially increasing function with respect to |κ| when |κ| > k. Hence, the noise may be exponentially amplified for large |κ|. For fixed distance d, the cutoff frequency kc depends on the noise level (or signal-to-noise ratio). Here, following [29], we choose kc in such a way that 2 − kc −k 2 d eik2 (kc )d = e = σ. (3.21) That is, the spatial frequency with the transfer function eik2 (κ)d below the noise level σ is cut off. More explicitly,  k c = k 2 + 1 log σ d  2 1/2  . (3.22) In view of (3.21) or (3.22), the pseudo-inverse (3.20) offers a regularization strategy 28 5 kc/k 4 3 2 1 0.2 0.4 d/λ 0.6 0.8 1 Figure 3.3: kc /k versus the distance varying from λ/10 to λ. for the inverse problem. We plot the function kc /k for various distance d in Figure 3 at 5% noise level. It is easily seen that at the fixed noise level, the cutoff frequency kc k when d < λ, i.e., the bandwidth of the spatial frequency in the near field is much larger than in the far field. This guarantees better resolution for the final reconstruction in the near-field regime, since the higher spatial frequency components of f are recovered. Denote ˆ h(κ) = Id (κ)ˆ(κ, d). u (3.23) To compute h, the FFT may be applied to compute the inverse Fourier transform, where h is an approximation of f ϕ. To reconstruct f from h, we need to take into account of the boundary data on Γ, which turns out to be a (well-posed) nonlinear problem. We next introduce some notations for representing the surface displacement f . Denote by C 0,1 (γ) the set of Lipschitz continuous functions on γ. Introduce the 0,1 Banach space C0 (γ) := {f | f ∈ C 0,1 (γ), f = 0 on ∂γ} with the usual norm 29 f 0,1 = f ∞ + supx ,y ∈γ,x =y 1 1 1 1 |f (x1 ) − f (y1 )| . |x1 − y1 | For a fixed small number δ, we define γδ := {x1 ∈ γ | dist(x1 , ∂γ) < δ}. For x1 ∈ γδ , let xb ∈ ∂γ such that x1 − xb = dist(x1 , ∂γ). Assume that { x0 , x1 , x2 · · · xN } is a set of grid points defined on γ. We represent f by a piecewise 1 1 1 1 linear function, where j−1 j f is linear on [x1 , x1 ] for j = 1, 2, · · · , N , and is continuous on γ globally. Moreover, it is strictly greater than 0 for the interior grid points x1 , x2 · · · xN −1 , 1 1 1 and is 0 on the boundary x0 , xN . We denote the set of all such functions by P1 . It 1 1 ˜ 0,1 is clear that P1 is a subset of C0 (γ), which is defined as follows: 0,1 ˜ 0,1 C0 (γ) := {f ∈ C0 (γ) | f (x1 ) > 0 for x1 ∈ γ; ∃ , δ > 0, s.t. f (x1 ) ≥ x1 − xb for x1 ∈ γδ }. For fixed grid points { x0 , x1 , x2 · · · xN }, δ = min{ x1 − x0 , xN − xN −1 }, 1 1 1 1 1 1 1 1 N −1 ) 1) f (x1 f (x1 ˜ 0,1 = min{ , }. On the other hand, C0 (γ) is an open subset of 1 −x0 N −xN −1 x1 1 x1 1 0,1 C0 (γ). ˜ We rewrite the integral operator (3.12) on γ by introducing the operator Kf defined as ˜ (Kf ϕ)(x1 ) := γ G(x1 − y1 ; f (x1 ), f (y1 ))ϕ(y1 )dy1 , x1 ∈ γ, (3.24) ˜ 1 + f 2 . The kernel of the integral operator Kf has explicit depen˜ dence on f . Here we adopt the subscript for Kf to emphasize its dependence on f . Similarly, define gf (x1 ) := g(x1 , f (x1 )), where g = −(ui + ur )| is the boundary ˜ Γ data. where ϕ = ψ 30 A natural way of separating f from h computed in (3.23) is to minimize the functional min f ∈P1 ˜ Kf ϕ − gf ˜ L2 (γ) + fϕ − h L2 (γ) . (3.25) However, this minimization problem is difficult to solve in practice, since f and ϕ are both unknowns. One alternative is to solve min f ϕ − h 2 L (γ) f ∈P1 ˜ subject to Kf ϕ = gf . ˜ It should be pointed out that in general ϕ is not Fr´chet differentiable with respect e ˜ to f since the operator Kf is compact. Therefore, existing numerical methods such as Newton’s method can not be applied directly. We introduce a new function h0 such that h0 is Lipschitz continuous on γ , more¯ over    h(x ) x ∈ γ\γ , for sufficiently small δ; 1 1 δ h0 (x1 ) =  0  x1 ∈ ∂γ. (3.26) By choosing a small δ (usually the length of two neighboring grid points), h0 − h 2 L (γ) is small. In practice, h0 may be chosen as the following piecewise linear function:    h(xj ) j = 1, 2, · · · , N − 1; j 1 h0 (x1 ) =  0  j = 0, N. Now we match the data on the boundary by solving the minimization problem with a constraint: ˜ min Kf ϕ − gf 2 ˜ L (γ) f ∈P1 where ϕ ∈ C(γ) satisfies f ϕ = h0 . (3.27) Remark 3.1. The minimization problem (3.27) is a special case of (3.25). By letting f ϕ = h0 , (3.25) becomes 31 ˜ min Kf ϕ − gf 2 ˜ + h0 − h 2 , L (γ) L (γ) and h0 − h 2 is small by the definition of h0 . L (γ) ˜ 0,1 Remark 3.2. For f ∈ P1 ⊂ C0 (γ), the function ϕ ∈ C(γ) that satisfies f ϕ = h0 is well and uniquely defined. Moreover 1 ϕ ∞≤( + 1/ ) h0 0,1 . minx ∈γ\γ f (x1 ) 1 δ f The problem (3.27) can be solved by Newton’s method. Since ( )2 1, the λ iteration is expected to converge fast to the real solution, which is confirmed by our numerical examples. To linearize the problem, we require the mapping F (f ) := ˜ Kf ϕ − gf be Fr´chet differentiable with respect to f ∈ P1 . ˜ e 3.3.2 Fr´chet Differentiability of the Nonlinear Operator e ˜ Here and thereafter, M and M stand for some generic positive constants, whose values may vary from step to step but should be clear from the contexts. Let L(C(γ), C(γ)) be the set of all bounded linear operators that map the functional space C(γ) to itself. 0,1 ˜ 0,1 ˜ Lemma 3.3.1. If f ∈ C0 (γ) ⊂ C0 (γ), then the mapping f → Kf is Fr´chet e 0,1 differentiable from C0 (γ) to L(C(γ), C(γ)). Moreover, the Fr´chet derivative is the e 0,1 ˜ ˜ linear mapping δf → (Kf ) (δf ) for δf ∈ C0 (γ), where (Kf ) (δf ) ∈ L(C(γ), C(γ)) is defined as ˜ [(Kf ) (δf )]ϕ(x1 ) = ˜ ∂ Φ(x1 − y1 , f (x1 ) − f (y1 )) (δf (x1 ) − δf (y1 )) ∂x2 γ ˜ ∂ Φ(x1 − y1 , f (x1 ) + f (y1 )) (δf (x1 ) + δf (y1 )) ϕ(y1 )dy1 − ∂x2 for ϕ ∈ C(γ). 32 ˜ Proof. It is clear that the mapping δf → (Kf ) (δf ) is linear. ˜ Step1: the mapping δf → (Kf ) (δf ) is bounded. Denote r1 = (x1 − y1 )2 + (f (x1 ) − f (y1 ))2 , r2 = (x1 − y1 )2 + (f (x1 ) + f (y1 ))2 . We first estimate the first part: ˜ ∂ Φ(x1 − y1 , f (x1 ) − f (y1 )) ∂x2 ik 4 = ≤ k 4 (1) H0 (kr1 ) (1) H0 (kr1 ) f (x1 ) − f (y1 ) r1 . For a small fixed constant τ0 , if |y1 − x1 | ≥ τ0 (away from the singularity), then (1) H0 (kr1 ) ˜ ≤M . It follows that ˜ ∂ Φ(x1 − y1 , f (x1 ) − f (y1 )) (δf (x1 ) − δf (y1 ))ϕ(y1 )dy1 ∂x2 γ\{|y1 −x1 |<τ0 } ≤ M δf ϕ ∞. 0,1 On the other hand, if |y1 − x1 | < τ0 , (1) H0 (kr1 ) small. We have ≤ ∼ O( 1 ) for τ0 sufficiently r1 ˜ ∂ Φ(x1 − y1 , f (x1 ) − f (y1 )) (δf (x1 ) − δf (y1 ))ϕ(y1 )dy1 ∂x2 {|y1 −x1 |<τ0 } M ϕ ∞. ϕ ∞ ≤ M δf |x1 − y1 | dy1 δf 0,1 0,1 {|y1 −x1 |<τ0 } r1 33 For the second part, ˜ ∂ Φ(x1 − y1 , f (x1 ) + f (y1 )) (δf (x1 ) + δf (y1 ))ϕ(y1 )dy1 ∂x2 γ ˜ ∂ Φ(x1 − y1 , f (x1 ) + f (y1 )) = [δf (y1 ) − δf (x1 )]ϕ(y1 )dy1 ∂x2 γ ˜ ∂ Φ(x1 − y1 , f (x1 ) + f (y1 )) +2 ϕ(y1 )dy1 δf (x1 ) ∂x2 γ =: A1 + A2 From the estimate of the first term, the following inequality also holds: |A1 | ≤ M δf ϕ ∞. 0,1 For x1 ∈ γ\γ δ , 2 |A2 | ≤ Mδ δf ϕ ∞, 0,1 where Mδ = k min 2 x ∈γ\γ 1 δ 2 (1) H0 (kf (x1 )) . For x1 ∈ γ δ , 2 |A2 | ≤ k 2 γ\[x1 −δ/2,x1 +δ/2] + (1) H0 (kr2 ) [x1 −δ/2,x1 +δ/2] dy1 δf (x1 ) ϕ ∞ (1) kδ H0 ( 2 ˜ ≤ M ˜ +M  ˜ ≤ M x1 +δ/2 x1 −δ/2 (1) kδ H0 ( ) 2 ϕ ∞ δf 0,1 1 (y1 − x1 )2 + f (x1 )2 + ln 2 − ln(1 − 34 dy1 δf (x1 ) ϕ ∞  δ δ 2 + ( /2)2 x1 − xb 2 ) x1 − xb    ˜ ≤ M (1) kδ H0 ( ) 2 + ln 2 − ln(1 − δ δ 2 + ( /2)2 x1 − xb 2 ) x1 − xb  δf ϕ ∞. 0,1 The last inequality follows from the fact that there exists an such that f (x1 ) ≥ x1 − xb for x1 ∈ γδ , and δf (x1 ) ≤ δf x − xb . 0,1 1 Therefore |A2 | ≤ M ( , δ) δf ϕ ∞ , 0,1 where ˜ M ( , δ) = M ˜ −M (1) kδ (H0 ( ) 2  ˜ + M ln 2  δ ln(1 − inf x1 ∈γδ/2 δ 2 + ( /2)2 x1 − xb 2 ) x1 − xb  is a positive constant. Therefore ˜ [(Kf ) (δf )]ϕ ≤ M ( , δ, f ) δf ϕ ∞ ∞ 0,1 for any ϕ ∈ C(γ), i.e., ˜ (Kf ) (δf ) ≤ M ( , δ, f ) δf . 0,1 L(C(γ),C(γ)) 35 ˜ ˜ ˜ Step2: An estimate of the remainder term Kf +δf − Kf − (Kf ) (δf ). For any ϕ ∈ C(γ), by Taylor’s formula, = ˜ ˜ ˜ [Kf +δf − Kf − (Kf ) (δf )]ϕ(x1 ) ˜ 1 ∂ 2 Φ(x1 − y1 , f (x1 ) − f (y1 ) + t(δf (x1 ) − δf (y1 )) ) ∂x2 2 γ 0 dt [δf (x1 ) − δf (y1 )]2 ϕ(y1 )dy1 ˜ 1 ∂ 2 Φ(x1 − y1 , f (x1 ) + f (y1 ) + t(δf (x1 ) + δf (y1 )) ) − dt ∂x2 γ 0 2 2 ϕ(y )dy [δf (x1 ) + δf (y1 )] 1 1 = B1 + B2 . δf 0,1 is sufficiently small. For a fixed small constant τ0 , if |y1 − x1 | ≥ τ0 (away from By a similar argument as in Step 1, we can estimate B1 . Assume that singularity), then ˜ 1 ∂ 2 Φ(x1 − y1 , f (x1 ) − f (y1 ) + t(δf (x1 ) − δf (y1 )) ) dt ∂x2 γ\{|y1 −x1 |<τ0 } 0 2 2 [δf (x1 ) − δf (y1 )]2 |ϕ(y1 )| dy1 ≤ M δf ϕ ∞. 0,1 If |y1 − x1 | < τ0 , ˜ ∂ 2Φ 1 ∼ O( ) for τ0 sufficiently small. Thus 2 ∂x2 r1 2 ˜ 1 ∂ 2 Φ(x1 − y1 , f (x1 ) − f (y1 ) + t(δf (x1 ) − δf (y1 )) ) dt ∂x2 {|y1 −x1 |<τ0 } 0 2 2 M ϕ ∞ |x1 − y1 |2 dy1 δf [δf (x1 ) − δf (y1 )]2 |ϕ(y1 )| dy1 ≤ 2 0,1 {|y1 −x1 |<τ0 } r1 2 ϕ ∞. ≤ M δf 0,1 36 1 2 Next, the term B2 can also be split into two parts B2 and B2 : ˜ 1 ∂ 2 Φ(x1 − y1 , f (x1 ) + f (y1 ) + t(δf (x1 ) + δf (y1 )) ) dt ∂x2 γ 0 2 2 ϕ(y )dy [δf (x1 ) − δf (y1 )] 1 1 1 B2 = and ˜ 1 ∂ 2 Φ(x1 − y1 , f (x1 ) + f (y1 ) + t(δf (x1 ) + δf (y1 )) ) dt ∂x2 γ 0 2 [δf (x1 )δf (y1 )]ϕ(y1 )dy1 . 2 B2 = 4 2 It suffices to estimate B2 . For x1 ∈ γ\γ δ , 2 2 2 B2 ≤ Mδ δf ϕ ∞, 0,1 where Mδ = min k 2 x1 ∈γ\γ δ 2 (1) 2k ) H0 ( f (x1 ) + 2k f (x1 ) (1) 2k ) H0 ( f (x1 ) is sufficiently small. if δf 0,1 For x1 ∈ γ δ , 2 2 B2 ≤ γ\[x1 −δ/2,x1 +δ/2] δf (x1 ) + [x1 −δ/2,x1 +δ/2] ϕ ∞ δf 0,1 37 1 ∂ 2Φ ˜ dt dy1 0 ∂x2 2 x1 +δ/2 2 1 ˜ ˜ ≤ Mδ δf ϕ ∞+M dy1 0,1 x1 −δ/2 (y1 − x1 )2 + ( /2)2 x1 − xb 2 δf (x1 ) δf ϕ ∞ 0,1 ˜ 2 M ˜ δ (x ) ≤ Mδ δf ϕ ∞+ 0,1 /2 x1 − xb f 1 2 ≤ M ( , δ) δf ϕ ∞. 0,1 δf ϕ ∞. 0,1 Therefore 2 ˜ ˜ ˜ Kf +δf − Kf − (Kf ) (δf ) = O( δf ), L(C(γ),C(γ)) 0,1 for sufficiently small δf , which completes the proof. 0,1 0,1 ˜ 0,1 Let h0 ∈ C0 (¯ ) be defined as in (3.26). For f ∈ P1 ⊂ C0 (γ), let ϕ ∈ C(γ) γ satisfy f ϕ = h0 . The following lemma concerns the Fr´chet derivative of the mapping e 0,1 f → ϕ ( C0 (γ) → C(γ) ). 0,1 ˜ 0,1 Lemma 3.3.2. If f ∈ C0 (γ) ⊂ C0 (γ), then the mapping f → ϕ defined as the 0,1 above is Fr´chet differentiable from C0 (γ) to C(γ). Moreover, its Fr´chet derivative e e 0,1 is the linear mapping δf → ϕ for δf ∈ C0 (γ), where ϕ ∈ C(γ) and ϕ(x1 ) ϕ (x1 ) = − δf (x1 ) for x1 ∈ γ. f (x1 ) Proof. It is easy to show that 1 ϕ (x1 ) ≤ (Mδ + ) ϕ ∞ δf 0,1 for x1 ∈ γ , 1 where Mδ = minx ∈γ\γ is a constant. Therefore 1 δ f (x1 ) ϕ 1 ≤ (Mδ + ) ϕ ∞ δf 0,1 ∞ 0,1 and the mapping δf → ϕ is bounded from C0 (γ) → C(γ). 38 0,1 For a perturbation of f with δf ∈ C0 (γ), a perturbation of ϕ satisfies (ϕ + δϕ )(f + δf ) = h0 . If δf is sufficiently small, for x1 ∈ γ, the following estimate for the high order 0,1 term holds: (ϕ + δϕ )(x1 ) − ϕ(x1 ) − ϕ (x1 ) 1 = |ϕ(x1 )| δf (x1 ) (f (x1 ) + δf (x1 ))f (x1 ) 2 1 1 2 ≤ (Mδ + ) ϕ ∞ δf . 2 2 0,1 Thus (ϕ + δϕ ) − ϕ − ϕ 2 = O( δf ), ∞ 0,1 for sufficiently small δf . 0,1 By Taylor’s expansion, it is easily seen that the mapping f → gf (:= g(x1 , f (x1 )) ) ˜ 0,1 e is also Fr´chet differentiable from C0 (γ) to C(γ). We denote its Fr´chet derivative e as the mapping δf → gf . Combining Lemma 3.3.1 and Lemma 3.3.2 and using the ˜ product rule, we have the following theorem: 0,1 ˜ 0,1 ˜ Theorem 3.3.3. If f ∈ C0 (γ) ⊂ C0 (γ), F (f ) := Kf ϕ − gf is Fr´chet dif˜ e 0,1 ferentiable from C0 (γ) to C(γ). Moreover, the Fr´chet derivative maps δf to e ˜ ˜ ˜ DF (δf ) = [(K)f (δf )]ϕ + Kf (ϕ ) − gf . 3.4 Numerical Examples First, let us consider the solution of the forward scattering problem. By Proposition ˜ 3.2.3, if k 2 is not an eigenvalue of −∆ in D, to solve the forward scattering problem (4.1)-(4.3) efficiently in our numerical simulation, we can firstly solve the integral 39 Figure 3.4: Real part (left) and the imaginary part (right) of the scattered field. equation Kψ = g and substitute ψ into (3.7). If k 2 is an eigenvalue, then the forward scattering problem can be solved by introducing the artificial boundary (e.g. half circle) and solving the problem in a bounded domain. Since our focus is on inverse problem, without loss of generality, we assume that k 2 is not an eigenvalue in our numerical examples. In the following examples, an incident wave ui = eikq·x /1002 with normal incident direction impinges on the obstacle. The wavenumber k = 100, λ ≈ 6.28 cm, q = (0, −1)T . In all the figures, the plots are rescaled with respect to the wavelength λ. Example 4.1. The real surface displacement is represented by two bumps, each one with the size of order λ. The two bumps are close to each other, and separated with distance λ/10. The scattered field in the region [−3λ, 3λ] × [0, 3λ] is plotted in Figure 3.4. Data are collected above the obstacle with distance d = λ/5 (nearfield). We also assume that 5% noise is added to the simulation data. It follows from 40 y/λ 0.1 0.05 0 −1 −0.6 −0.2 0 0.2 0.6 1 0.2 0.6 1 x/λ y/λ 0.1 0.05 0 −1 −0.6 −0.2 0 x/λ Relative error Figure 3.5: Near-field (top) and far-field (bottom) images. The solid line represents the real image, and the dotted line is the reconstruction. 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 10 Iterations 20 Figure 3.6: Relative error with respect to Newton iterations. 41 30 y/λ 0.1 0.05 0 −1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1 x/λ y/λ 0.1 0.05 0 −1 −0.8 −0.6 −0.4 −0.2 0 x/λ Figure 3.7: Near-field (left) and far-field (right) images. The solid line represents the real image, and the dotted line is the reconstruction. Y/λ 0.2 0.1 0 −10 −8 −6 −4 −2 0 2 4 6 8 X/λ Figure 3.8: Real profile, the distance between neighboring bumps is λ/10. 42 10 Y/λ 0.2 0.1 0 −10 −8 −6 −4 −2 0 2 4 6 8 10 X/λ Figure 3.9: Comparison of the reconstruction (dotted line) with the real profile (solid line) for the near-field case. (3.22) that kc ≈ 2.6k. The near-field image fn and the real image f are plotted in Figure 3.5 (left). Though two bumps are close to each other (λ/10), they are clear distinguishable. Therefore, super-resolution is achieved with near-field measurements. To confirm the convergence of Newton’s method for solving the minimization problem (3.27), the relative error is shown with respect to the iteration number in Figure 3.6. Here the relative error is defined as 1/2 N f (xj ) − f (xj ) 2 n 1 j=0 1 N f (xj )2 1/2 j=0 1 . The reconstruction converges fast, which leads to the real surface displacement after the first 20 iterations. To compare with the far-field image, we collect the data again with d = 5λ and 5% noise. It is obvious that kc = k in the far-field case. The image fn is shown in Figure 3.5 (right). It is clear that the two bumps can not be distinguished, which is due to the fact that the high spatial frequency information of the surface displacement is lost. Example 4.2 We consider a non-smooth profile in this example. Two bumps are separated with distance λ/10. The measurement is polluted with 5% noise. Figure 3.7 is the reconstructed near-field and far-field images when d = λ/5 and 5λ respec43 tively. Super-resolution is also achieved via near-field imaging. On the other hand, it is also observed that the accuracy of the reconstruction is deteriorated when data is collected at far field. Example 4.3 The real surface displacement is a long periodic structure (see Figure 3.8). Each period is a bump with size of order λ, and two neighboring bumps are separated with distance λ/10. The measurement distance d again is λ/5, where 5% noise is added to the simulation data. We compare the near-field image and the real image in Figure 3.9. The periodic structure is also accurately reconstructed with super-resolution. 3.5 Discussions The proposed numerical method in the near-field regime, which recovers high spatial frequency modes of the surface displacement by taking account of the evanescent modes, yields super-resolution for the reconstructed image. Numerically, the Newton iteration for the minimization problem (3.27) converges fast to the real solution. However, no rigorous theoretical convergence analysis is presently available. Another issue is denoising in the near-field regime. Noise will be exponentially amplified in near-field regime. The cutoff wavenumber kc , which is the critical parameter for the resolution of the reconstructed image, strongly depends on the noise level. A denosing technique based on the broadband signal has recently been proposed in [29] for the linearized model. For the nonlinear imaging, the problem becomes much more challenging and is completely open. f To reconstruct more general local surface displacement (without assumption ( )2 λ 1), we propose to use multiple frequency data. This is discussed fully in the next Chapter. 44 Chapter 4 Imaging with Multiple Frequency Data 4.1 4.1.1 Introduction Problem Formulations We consider the very general case without specific restriction on the surface displacement in this Chapter. Figure 4.1 is the schematic setup of the scattering problem and data collection. As in Chapter 3, we let x = (x1 , x2 ), and f (x1 ) be a function defined on the real line R with compact support γ. The curve ∂D := { x | x1 ∈ R, x2 = f (x1 ) } represents the boundary of the whole obstacle on which the electromagnetic wave impinges, and the domain above ∂D is denoted as D. Let R be some positive constant such that γ ⊂⊂ (−R, R) and Γ := { x | x1 ∈ (−R, R), x2 = f (x1 ) } represent the local disturbance. Let BR be a disk with radius + R centered at origin. ∂BR := ∂BR ∩ D is the half circle above the ground plane, where ∂BR is the boundary of the disk BR . Without loss of generality, we assume + that Γ lies below ∂BR ( see Figure 4.1). We also denote the domain bounded by 45 Figure 4.1: Problem geometry. + ∂BR ∪ Γ as DR . The mathematical model for the forward scattering problem is the same as in Chapter 3. The total field u satisfies the Helmholtz equation ∆u + k 2 u = 0 in D. (4.1) u = 0 on ∂D. (4.2) and vanishes on the boundary: In addition, at infinity the scattered field us satisfies the Sommerfeld radiation condition: lim r→∞ √ r ∂us − ikus ∂r = 0, r = |x| . (4.3) Now we are ready to present the inverse scattering problem. The total field u is + collected on the half circle ∂BR , and the measurements are assumed to be available for a set of wavenumbers { kj | j = 1, 2, · · · , M ; kj < kj+1 }. The inverse problem is to reconstruct the local disturbance Γ from the measured multiple frequency data + on ∂BR . 46 4.1.2 Why Multiple Frequency Data In the presence of noise, the resolution of the reconstructed image is usually limited by the operating frequency of the electromagnetic wave. When the height of the displacement is small compared to the wavelength, it is already observed in Chapter 3 that a high spatial frequency mode of the scattered wave contains exactly the high spatial frequency information (fine features) of the profile. Theoretical studies also reveal that higher frequency information may yield increased stability (less illposedness). In fact, in our recent paper on a multiple frequency inverse source problem for the Helmholtz equation [12], it is proved that a H¨lder type stability estimate may o be obtained with sufficiently high wavenumber. In [40], Isakov also proved increased stability for reconstructing the Schr¨dinger potential from the Dirichlet-to-Neumann o map with higher wavenumber. On the other hand, when only single frequency data is available, classical regularized iterative optimization methods such as Newton’s method ( see for example [31] ) applied to the inverse scattering problem usually fail to compute the global minimizer. Multiple frequency data overcomes the difficulty of reaching some local minimum. In [7, 8, 9, 22], a stable recursive linearization method is developed for the inverse medium scattering problem with multiple frequency data. The method applies the Born approximation at the lowest frequency to obtain the initial guess of the medium, and sequentially updates at higher frequency until the dominant modes of the medium are essentially recovered. The convergence of the recursive linearization method has been analyzed recently in [13]. 4.2 4.2.1 Analysis of the Forward Scattering Problem Notations 47 We begin with some standard notations that will be used throughout this Chapter. Let H1 (DR ) := { u | u ∈ L2 (DR ), ∂j u ∈ L2 (DR ) } be the standard Sobolev space equipped with the norm u 1,D = u 0,D + |u|1,D , R R R where 2 u 2 0,DR = D |u(x)| dx, R |u|2 1,DR = 2 j=1 DR ∂j u 2 dx. Let (r, θ) represent the polar coordinates. Define the Sobolev space H1/2 (∂BR ) as H1/2 (∂BR ) := { ϕ ∈ L2 (∂BR ) | ∞ 1 + n2 ϕs 2 + ϕc 2 n n < +∞ }, n=0 where ϕs and ϕc are the Fourier coefficients of the function ϕ on the circle ∂BR : n n ϕs = n ϕc = 0 1 π ϕ(θ) sin(nθ)dθ π −π π 1 ϕ(θ)dθ, 2π −π ϕc = n (n ≥ 0), 1 π ϕ(θ) cos(nθ)dθ π −π (n ≥ 1). + For a function ϕ defined on the half circle ∂BR , it is extended to ∂BR by odd reflection:    ϕ(θ) θ ∈ (0, π), ϕ(θ) = ˜  −ϕ(−θ) θ ∈ (−π, 0).  Define + + ˜ ˜ H1/2 (∂BR ) := { ϕ ∈ H1/2 (∂BR ) | ϕ ∈ H1/2 (∂BR ), where ϕ is the odd extension of ϕ }. ˜ 48 Let ϕn = 2 π ϕ(θ) sin(nθ)dθ. π 0 Clearly ϕs = ϕn and ϕc = 0. Hence ˜n ˜n + + ˜ H1/2 (∂BR ) = { ϕ ∈ H1/2 (∂BR ) | ∞ 1 + n2 |ϕn |2 < +∞ }, n=1 with the norm  ϕ 4.2.2  + = ˜ H1/2 (∂BR ) ∞ 1/2 1 + n2 |ϕn |2  . n=1 Dirichlet-to-Neumann Map We reformulate the forward scattering model (4.1) - (4.3) in the bounded domain + DR by introducing the Dirichlet-to-Neumann map on ∂BR . A similar Dirichlet-toNeumann map is also introduced in [63] for the scattering from an overfilled cavity. In fact, the idea of using pseudo-differential operators to reduce the infinite domain to a bounded domain has been applied to various wave simulation problems. See for example [3, 4, 33]. Let ϕ(θ) = us | + . From (4.1) - (4.3), it follows that the scattered filed us ∂BR satisfies   ∆us + k 2 us = 0  in D\BR ,      s +   u =ϕ on ∂BR ,  us = 0 on Γ0 ,     s    limr→∞ √r ∂u − ikus = 0, r = |x| .  ∂r The solution of the scattering problem takes the form us (r, θ) = ∞ π 2 (1) ϕ(θ) sin(nθ)dθ Hn (kr) sin(nθ), (1) n=1 πHn (kR) 0 49 (1) where Hn is the first kind Hankel function of order n. Let ν be the unit normal on + + ∂BR directed into the infinite domain D. Then the normal derivative on ∂BR can be written as ∂us 2k (R, θ) = ∂r π (1) (Hn ) (kR) π ϕ(θ) sin(nθ)dθ sin(nθ). (1) Hn (kR) 0 n=1 ∞ Define the Dirichlet-to-Neumann map T ( ϕ → 2k (T ϕ)(θ) = π ∂us ) as ∂r (1) (Hn ) (kR) π ϕ(θ) sin(nθ)dθ sin(nθ). (1) Hn (kR) 0 n=1 ∞ (4.4) + ˜ Lemma 4.2.1. The Dirichlet-to-Neumann map T is bounded from H1/2 (∂BR ) to + ˜ its dual space H1/2 (∂BR ) . (1) ˜ 1/2 (∂B + ), let cn = (Hn ) (kR) . Then Proof. For any ϕ, ψ ∈ H R (1) Hn (kR) ∞ π ∞ ¯ dθ = kπR ¯ ¯ ds = kR cn ϕn sin(nθ)ψ cn ϕn ψn . < T ϕ, ψ >= +Tϕψ 2 0 n=1 ∂BR n=1 Here, ϕn = 2 π 2 π ϕ(θ) sin(nθ)dθ and ψn = ψ(θ) sin(nθ)dθ. π 0 π 0 From the recurrence relation of Hankel functions [2], (1) (1) (1) (1) −Hn+1 (kR) + n Hn (kR) −Hn+1 (kR) (Hn ) (kR) n kR cn = = = + , (1) (1) (1) kR Hn (kR) Hn (kR) Hn (kR) Thus |cn | ≤ 1 + n ≤c kR 1 + n2 for some positive constant c. 50 Now |< T ϕ, ψ >| ≤  ckπR  2 ≤ ∞ ckπR 2 ∞ n=1 ¯ 1 + n2 ϕn ψn 1/2  1 + n2 |ϕn |2  ∞  1/2 1 + n2 |ψn |2  , n=1 n=1 i.e., |< T ϕ, ψ >| ≤ ckπR ϕ 1/2 ψ 1/2 . ˜ ˜ 2 H H The proof is now complete. 4.2.3 Well-posedness of the Forward Scattering Problem in Bounded Domain The normal derivative of the total field is given by ∂u ∂us ∂ui ∂ur ∂ui ∂ur = + + = T (u) − T (ui + ur ) + + . ∂ν ∂ν ∂ν ∂ν ∂ν ∂ν ∂ui ∂ur + − T (ui + ur ), then the total field satisfies the following ∂ν ∂ν boundary value problem: Denote g =   ∆u + k 2 u = 0 in D ,   R   u=0 on Γ,   ∂u   +  = T (u) + g on ∂BR . ∂ν (4.5) Next, we address the well-posedness of the scattering problem (4.5), which also provides the basis for the analysis of the domain derivative of the forward scattering map in the next section. Define a subspace of H1 (DR ) 51 + ˜1 ˜ H0 (DR ) := { u ∈ H1 (DR ) | u = 0 on Γ, u| + ∈ H1/2 (∂BR ) }. ∂BR By introducing the bilinear form a(u, w) = DR u· w − k 2 uw dx− < T u, w > ¯ ¯ ˜1 ˜1 for u, w ∈ H0 (DR ), then u ∈ H0 (DR ) is a weak solution of the boundary value problem (4.5) if ˜1 for all w ∈ H0 (DR ). a(u, w) = < g, w > (4.6) Theorem 4.2.2. The variational problem (4.6) attains a unique solution. Moreover, u 1,D ≤ C g R + ˜ H1/2 (∂BR ) for some positive constant C. Proof. First we prove a G˚ arding type inequality. It is easy to show that kπR < T u, u >= 2 where ∞ cn |un |2 , (4.7) n=1 (1) (Hn ) (kR) 2 π u(θ) sin(nθ)dθ. cn = , and un = (1) π 0 Hn (kR) (1) Note that Hn = Jn + i Yn , where Jn and Yn are the first and second kind Bessel functions respectively, and the modulus of the Hankel function is a decreasing function [2], then cn = Here 2 2 1 (Jn ) (kR) + (Yn ) (kR) Jn (kR)Jn (kR) + Yn (kR)Yn (kR) = < 0. 2 2 2 2 2 Jn (kR) + Yn (kR) Jn (kR) + Yn (kR) denotes the real part of a function. From (4.7) and (4.8) it follows that (< T u, u >) < 0. Consequently 2 a(u, u) ≥ u 2 ˜ 1,DR − c u 0,DR 52 (4.8) for some positive constant c. ˜ To prove the existence of a weak solution, from the Fredholm alternative we only ˜1 need to prove the uniqueness. If a(u, w) = 0 for any w ∈ H0 (DR ), then the imaginary part of a(u, u) kπR a(u, u) = − < T u, u >= − 2 ∞ cn |un |2 = 0. n=1 By the Wronskian formula [2], cn = Jn (kR)Yn (kR) − Yn (kR)Jn (kR) 2 1 = > 0. 2 (kR) + Y 2 (kR) 2 (kR) + Y 2 (kR) kπR Jn Jn n n + ˜1 Hence un = 0 for all n. Note that u ∈ H0 (DR ), therefore u = 0 on ∂BR , and ∂u + = T u = 0 on ∂BR . T u = 0 by definition. On the other hand, g = 0 implies that ∂ν Now u ≡ 0 in DR follows from the unique continuation result [41]. The continuous dependence of the solution on g follows from the Lax-Milgram lemma and the Fredholm alternative. 4.3 Domain Derivative of the Forward Scattering Map The theory of shape sensitivity analysis has been studied extensively for various shape optimization problems. We refer the reader to [52, 56] for detailed discussions and references. For the inverse scattering problem, Kirsch rigorously derived the domain derivative of the far-field operator for a C 2 bounded obstacle in [44]. In this section, we derive the domain derivative of the operator that maps Γ to the measurement u| + (Theorem 4.3.1). ∂BR 53 Define the forward mapping M : Γ → u| + . It maps the local surface displace∂BR + ment to the measurement on the half circle ∂BR . The vector field V (x) is defined on 2 Γ. It is assumed that V (x) ∈ C0 (Γ; R2 ), i.e., twice continuously differentiable with compact support suppV ⊂⊂ Γ. For a given vector field V (x), denote 2 Γδ = {x + δ · V (x) | x ∈ Γ, V (x) ∈ C0 (Γ; R2 ) } as the perturbation of Γ with V (x). Now the domain derivative of the forward mapping M with respect to the direction V (x) is defined as M (Γ) := lim δ→0 M (Γδ ) − M (Γ) . δ (4.9) 2 Theorem 4.3.1. Let u be the solution to (4.1)-(4.3). If Γ is C 2 , V ∈ C0 (Γ; R2 ), then the domain derivative M (Γ) exists. Moreover, M (Γ) = u | + , where u ∂BR solves   ∆u + k 2 u = 0 in D ,   R   ∂u u = −(V · ν) on Γ, (4.10)  ∂ν   ∂u  +  = T (u ) on ∂BR . ∂ν Here ν is the unit normal on Γ directed into the infinite domain D. Proof. Let δ be a small real number. The scattered field uδ produced by the scattering from the perturbed profile Γδ satisfies the boundary value problem    ∆uδ + k 2 uδ = 0 in Dδ ,   R   uδ = 0 on Γδ ,    ∂uδ   +  = T (uδ ) + g on ∂BR , ∂ν + δ where DR is the domain bounded by Γδ and ∂BR . The weak solution of this bound54 ˜1 δ ary value problem uδ ∈ H0 (DR ) satisfies aδ (uδ , wδ ) = < g, wδ > ˜1 δ for all wδ ∈ H0 (DR ), (4.11) wδ − k 2 uδ wδ dx− < T uδ , wδ > . (4.12) where aδ (uδ , wδ ) = δ DR uδ · 2 We extend the definition of V ∈ C0 (Γ; R2 ) to the closure of the bounded domain DR , which is denoted by DR , such that V (x) is C 2 for x ∈ DR and V (x) = [0, 0]T if |x| > R − α for some small positive constant α. Define the mapping Ψ(y) by letting x = Ψ(y) = y + δV (y) for y ∈ DR . δ Clearly, Ψ is a C 2 mapping from DR → DR . Denote the inverse map of Ψ(y) as δ Φ(x), which maps DR → DR . ∂uδ ∂ uδ ∂Φm ˜ Let uδ (y) = uδ (Ψ(y)), wδ (y) = wδ (Ψ(y)). It follows that ˜ ˜ = 2 m=1 ∂ym ∂x , ∂xi i where Φm is the mth component of Φ. Therefore, the bilinear form (4.12) can be written as  aδ (uδ , wδ ) = Here, bmn = 2  DR m,n=1  δ ∂ wδ ∂u ˜ ˜ bmn − k 2 uδ wδ  J dy− < T uδ , wδ > . ˜ ˜ ˜ ˜ ∂ym ∂yn ¯ 2 ∂Φm ∂ Φn , the Jacobian J = det DΨ. i=1 ∂x ∂x i i Define a new bilinear form aδ (˜δ , w) by letting ˜ u  aδ (˜δ , w) = ˜ u 2  DR m,n=1  δ ∂w ∂u ˜ ¯ bmn − k 2 uδ w J dy− < T uδ , w > ˜ ¯ ˜ ∂ym ∂yn 55 ˜1 for uδ , w ∈ H0 (DR ). Then the variational problem (4.11) is equivalent to finding ˜ ˜1 uδ ∈ H0 (DR ) such that ˜ aδ (˜δ , w) = < g, w > ˜ u ˜1 for all w ∈ H0 (DR ). (4.13) From (4.6) and (4.13), it is easily seen that uδ − u satisfies ˜ a(˜δ − u, w) = −(˜δ (˜δ , w) − a(˜δ , w)). u a u u (4.14) On the right hand side,  aδ (˜δ , w) − a(˜δ , w) = ˜ u u 2  DR m,n=1 − uδ · ˜ DR  δ ∂w ¯ ∂u ˜ − k 2 uδ w J dy ˜ ¯ bmn ∂ym ∂yn w − k 2 uδ w dy. ¯ ˜ ¯ (4.15) The coefficient matrix (bmn ) and the Jacobian J can be further written as J = 1+δ · V + O(δ 2 ), (bm,n )J = I − δ(˜mn ) + O(δ 2 ), b (4.16) where I is the 2 × 2 identity matrix and the matrix (˜mn ) = b V + ( V )T − ( · V )I. (4.17) Therefore, from (4.14) - (4.17) and the continuous dependence of uδ − u on the right ˜ hand side, it follows that uδ − u ˜ 1,DR → 0 as δ → 0. 56 Now a( uδ − u ˜ 1 , w) = − (˜δ (˜δ , w) − a(˜δ , w)) a u u δ δ 2 ˜δ ¯ ˜mn ∂ u ∂ w + k 2 ( = b ∂ym ∂yn DR m,n=1 Since uδ − u ˜ 1,DR · V )˜δ w dy + O(δ). u ¯ → 0 as δ → 0, 2 uδ − u ˜ ¯ ˜mn ∂u ∂ w + k 2 ( a( , w) → b δ ∂ym ∂yn DR m,n=1 · V )uw dy ¯ as δ → 0. uδ − u ˜ exists, which we denote as u0 . Hence the limit limδ→0 δ Clearly, u0 is the solution to the following variational problem: 2 a(u0 , w) = DR m,n=1 ¯ ˜mn ∂u ∂ w + k 2 ( b ∂ym ∂yn · V )uw dy ¯ (4.18) ˜1 for any w ∈ H0 (DR ). By the assumption that Γ is C 2 , the scattered field u ∈ ¯ ˜1 C 2 (D) ∩ C(D) [61]. Apply the formula (4.17), then for any w ∈ H0 (DR ) ∩ H2 (DR ), 2 ¯ ˜mn ∂u ∂ w b ∂ym ∂yn m,n=1 = uT ( V + ( V )T − ( = [ (V · w) · ¯ −( ( u)V ) · = (V · w) · ¯ · V )I) w ¯ u − ( ( w)V ) · ¯ w] − ¯ u+ uT ( (V · 57 u] + [ (V · u) · w ¯ · V )I w ¯ u) · w− ¯ · [( u · w)V ]. ¯ + By the Green’s formula and noting that V = 0 on ∂BR , (4.18) can be reduced to a(u0 , w) = DR − Γ −(V · (V · w)∆u + ¯ w) ¯ (V · ∂u −( u· ∂ν u) · w + k2( ¯ · V )uw dy ¯ w)(V · ν)dsy . ¯ On the other hand, u = w = 0 on Γ, we have (V · w) ¯ ∂u −( u· ∂ν w)(V · ν) = 0 on ¯ Γ. Therefore, a(u0 , w) = = where D k 2 R a(u0 , w) = DR DR DR k 2 u(V · (V · w) + ¯ u) · (V · u) · w − k 2 (V · ¯ w + k2( ¯ u)w dy + ¯ · V )uw dy ¯ DR k2 · (uwV ) dy, ¯ · (uwV ) dy = 0 by the Divergence theorem. We finally obtain ¯ (V · u) · w − k 2 (V · u)w dy ¯ ¯ ˜1 for any w ∈ H0 (DR ) ∩ H2 (DR ). (4.19) It is easy to check that u0 defined above is the weak solution of the following boundary value problem   ∆u + k 2 u = (∆ + k 2 )(V ·   0 0   u =0  0  ∂u   0 = T (u )  0 ∂ν Let u = u0 − V · u) in DR , on Γ, + on ∂BR , + u, then u solves (4.10). Further, note that V = 0 on ∂BR , therefore u = u0 = lim δ→0 uδ − u ˜ + = M (Γ) on ∂BR . δ The proof is now complete. 58 4.4 Imaging with Multiple Frequency Data 4.4.1 Descent Direction for the Cost Functional + For a fixed wavenumber k, denote um as the measurement collected on ∂BR . For a given curve Γ, define the cost functional F (Γ) := 1 M(Γ) − um 2 2 + . 2 L (∂BR ) A descent direction on Γ is a vector field V such that the cost functional decreases, i.e., F (Γδ ) < F (Γ) if δ ∈ (0, t0 ] for some small positive constant t0 . We also call this descent direction V descent velocity. The perturbation of Γ may be viewed as an evolution process of assigning each point x on Γ a descent velocity V (x) and moving Γ with the velocity V (x) in a small artificial time interval [0, t0 ]. The characterization of a descent velocity for the cost functional F (Γt ) is established in the next theorem. Theorem 4.4.1. Let Γ be C 2 , ν be the unit normal on Γ directed into the infinite domain D. u is the solution to the forward scattering problem (4.1)-(4.3), and u∗ is the solution of the following boundary value problem:   ∆u∗ + k 2 u∗ = 0  in DR ,    u∗ = 0 on Γ,    ∂u∗  +  = T ∗ (u∗ ) + u − um on ∂BR . ∂ν (4.20) 2 where T ∗ is the adjoint operator of T . If V ∈ C0 (Γ; R2 ), and satisfies − then dF (Γt ) < 0 . Here dt t=0 Γ (V · ν) · ∂u ∂u∗ ds < 0, ∂ν ∂ν denotes the real part of a function. 59 (4.21) Proof. Clearly, u also solves the boundary value problem (4.5). Let u and u∗ be the solution of (4.10) and (4.20), respectively. By the Green’s formula, ∂u∗ ∂u ∂u ∂u∗ u − u∗ ds = u − u∗ ds. + ∂ν ∂ν ∂ν ∂BR Γ ∂ν (4.22) Using the boundary conditions in (4.10) and (4.20), the identity (4.22) is reduced to ∂u∗ ∂u T ∗ (u∗ )u ds+ (u−um )u ds− u∗ T (u )ds = − (V ·ν) ds. + + + ∂ν ∂ν ∂BR ∂BR ∂BR Γ Since T ∗ is the adjoint operator of T , we have ∂u∗ ∂u (u − um )u ds = − (V · ν) ds. + ∂ν ∂ν ∂BR Γ Therefore by Theorem 4.3.1, dF (Γt ) = dt t=0 m + (u − u )u ds ∂BR =− Γ (V · ν) · ∂u ∂u∗ ds. ∂ν ∂ν The assertion of the theorem holds. From (4.21), it is easily seen that in practice there are many possible choices for the descent direction V defined on Γ. In our case, since the local surface displacement 2 is represented by the function f , we let V = [0, v]T , where v ∈ C0 (Γ) is a compactly supported function on Γ satisfying the inequality − Γ (v · ν2 ) · ∂u ∂u∗ ds < 0. ∂ν ∂ν (4.23) Here ν2 is the second component of the unit normal ν. In particular, if v · ν2 = ∂u ∂u∗ , then V is the steepest descent direction for the cost functional F (Γt ). ∂ν ∂ν 60 ∂u ∂u∗ may not be a smooth and compactly supported function on ∂ν ∂ν Γ. Therefore, we need to modify its definition to make it smooth and compactly However, supported. This can be accomplished by the cubic spline interpolation [28]. We denote the new smooth, compactly supported function as vs . The choice of the smooth version vs of v is a regularization procedure in the regularization theory for ill-posed problems. 4.4.2 Reconstruction Scheme Suppose that multiple frequency measurements { um | j = 1, 2, · · · , M } on the half kj + are available for a set of wavenumbers { k | j = 1, 2, · · · , M }, where circle ∂BR j kj > ki if j > i. Starting from the flat surface as our initial guess, the proposed reconstruction method marches from k1 to kM . At the fixed wavenumber k = kj , by the cubic spline interpolation, a smooth version of the descent vector field V is defined, where V = [0, v]T and v satisfies (4.23). The reconstructed profile evolves with the chosen descent velocity at the fixed wavenumber. This evolution process continues until k = kM . Let Γrec be the reconstruction, Drec is the domain above the curve Γrec ∪ Γ0 . + rec DR := Drec ∩ BR is the domain bounded by Γrec and ∂BR . The reconstruction method with multiple frequency data is proposed as follows: (1) (Initialization) Let k = k1 , initially set Γrec = { (x1 , x2 ) | x1 ∈ (−R, R), x2 = 0 } (flat surface). (2) (Update the reconstruction by marching along the wavenumbers) FOR j = 1, 2, 3, · · · , M Let k = kj , 61 FOR n = 1, 2, 3, · · · , N (Evolution process at fixed wavenumber k ): 2 Choose the descent direction Vn = [0, vn ] s.t. vn ∈ C0 (Γ) and − Γ (vn · ν2 ) ·   ∆u + k 2 u = 0 in Drec ,    R  u=0 on Γrec ,   ∂u   +  = T (u) + g on ∂BR . ∂ν ∂u ∂u∗ ds < 0, ∂ν ∂ν where   ∆u∗ + k 2 u∗ = 0 rec  in DR ,    u∗ = 0 on Γrec ,    ∂u∗  +  = T ∗ (u∗ ) + u − um on ∂BR . k ∂ν Update: Γrec ←− Γrec + αn Vn , αn is the step length. END END Remark 5.1 The lowest wavenumber k1 has to be small to guarantee that the main features of the profile are captured. Based on the numerical investigation, one basic rule to follow is 1/k1 ≥ R, i.e., the wavelength λ1 corresponding to the lowest wavenumber k1 is at least comparable to R. The restriction may be due to the reason that the reconstruction with the measurement um usually captures the feature of order k1 λ1 , which needs to be sufficiently large in order that the main feature of the profile (of order R) is recovered in our approach. Without loss of generality, it is assumed that R = 1 and k1 = 1. On the other hand, 1/kM has to be smaller than the scale of the fine features of the profile such that the small details of the target are accurately reconstructed as well. Remark 5.2 Theorem 4.4.1 does not require that Γ be parameterized by some function f . In fact the assertion of the theorem is valid for any C 2 local disturbance Γ. Therefore, our proposed method can be modified to deal with more general cases. The corresponding descent direction may be chosen to be parallel to ν such that V = v · ν 62 ∂u ∂u∗ . The evolution process of Γ with the descent velocity V can be ∂ν ∂ν simulated by the level set method. and v = At low wavenumbers, the reconstruction captures the main features of the local disturbance (low frequency modes). It also serves as the starting point for the reconstruction at the next higher wavenumber, where the evolution process continues to recover the fine features of the local disturbance (high frequency modes). Ideally, the smaller the increment kj+1 −kj is, the better reconstruction image would be. In fact, the increment parameter kj+1 − kj depends on the scale feature of the real profile. If kj+1 − kj is too large, the reconstruction procedure may fail to recover the Fourier modes of the real profile between kj and kj+1 . From numerical simulations discussed in the next section, kj+1 − kj = 2 is sufficient to obtain accurate reconstruction. It is also worth pointing out that the reconstruction with only the single highest wavenumber k = kM will not yield a satisfactory image, as the convergence to the global minimizer of the cost functional strongly relies on the initial guess. In practice, a good initial guess is hard to choose without a priori information of the imaging target. This may be confirmed by numerical simulations presented in the next section. However, the reconstruction with multiple frequency data overcomes the difficulty of reaching some local minimum. Staring from the flat surface as an initial guess, the reconstructed image captures the large scale features of the profile at a lower wavenumber kj , which serves as the initial guess for the reconstruction at a higher wavenumber kj+1 . 63 4.5 Numerical Examples 4.5.1 Numerical approximation of the forward scattering problem ∂u ∂u∗ and on Γrec to define a descent direction at a fixed ∂ν ∂ν wavenumber, the solution u and u∗ in the interior domain need not be calculated. We need to calculate only Hence the boundary integral equation method may be applied to solve the forward problem, which is fast and reduces the complexity of the computation by solving a much smaller linear system compared to the finite element method. In particular, the boundary integral equation method can handle the high wavenumber problem. i (1) H (k |x − y|) be the fundamental solution for the Helmholtz 4 0 equation in R2 . The solution u to (4.5) satisfies the following system of integral Let Φ(x, y) = equations: 1 ∂u(y) u(x)+ Φ(x, y)dsy − + (T u)(y)Φ(x, y)dsy 2 Γrec ∂νy ∂BR ∂Φ(x, y) + + + ∂νy u(y)dsy = ∂B + g(y)Φ(x, y)dsy x ∈ ∂BR , ∂BR R ∂u(y) ∂Φ(x, y) Φ(x, y)dsy − + (T u)(y)Φ(x, y)dsy + ∂B + ∂νy u(y)dsy Γrec ∂νy ∂BR R = + ∂BR g(y)Φ(x, y)dsy 64 x ∈ Γrec . Similarly, we calculate ∂u∗ + on ∂BR by solving the system ∂ν 1 ∗ ∂u∗ (y) ∗ ∗ u (x)+ Φ(x, y)dsy − + (T u )(y)Φ(x, y)dsy 2 Γrec ∂νy ∂BR ∂Φ(x, y) ∗ m + + ∂νy u (y)dsy = ∂B + (u(y) − u (y))Φ(x, y)dsy ∂BR R + x ∈ ∂BR , ∂u∗ (y) ∂Φ(x, y) ∗ Φ(x, y)dsy − (T ∗ u∗ )(y)Φ(x, y)dsy + u (y)dsy + + Γrec ∂νy ∂BR ∂BR ∂νy m = + (u(y) − u (y))Φ(x, y)dsy x ∈ Γrec . ∂BR 4.5.2 Imaging of the local surface displacement Several numerical examples are presented to illustrate the efficiency of the proposed method. Example 1 considers a smooth upward profile (f ≥ 0), the convergence of the reconstruction is highlighted. We also compare the reconstruction with the image when only the single highest wavenumber is used. In Example 2, we remove the restriction that f ≥ 0. The surface displacement is represented by a general smooth function f , where Γ+ := { (x1 , f ) | x1 ∈ (−R, R), f (x1 ) > 0 } and Γ− := { (x1 , f ) | x1 ∈ (−R, R), f (x1 ) < 0 } are both nonempty sets. Example 3 discusses the reconstruction for the piecewise linear (nonsmooth) local surface displacement. In the last example, the reconstruction of a multiscale profile is presented. In all simulations, we use noisy data by adding 5% additive noise to the measurements. Example 1. Smooth upward profile 65    0.2 + 0.2 cos(4πx + π) x ∈ [−0.5, 0),  1 1   f (x1 ) = 0.1 + 0.1 cos(4πx1 + π) x1 ∈ [0, 0.5],      0 elsewhere. The boundary of the whole obstacle is C 1 (Figure 4.2 left). An incident wave ui = eikd·x with normal incident direction impinges on the obstacle. The wavenumbers k1 = 1 and kM = 20. The reconstruction fn when k = 16 and the real profile f are plotted in Figure 4.2 (right). It is observed that the reconstruction is accurate even though noise is present in measurements. To test the convergence of the proposed method, the relative error with respect to the wavenumber is shown in Figure 4.3. Here the relative error is defined as 1/2 R 2 −R |f (x1 ) − fn (x1 )| dx1 . R |f (x )|2 dx 1/2 1 1 −R It is clear that the relative error decreases until the main Fourier modes of f are recovered. Figure 4.4 illustrates the reconstructions for various wavenumbers. At low wavenumbers, the reconstruction captures the main features of the real profile, while the fine features of the profile is recovered as the evolution process continues. Next is the case when only the single highest frequency measurement is used. M(Γrec ) − um kM 2 + L (∂BR ) Figure 4.5 (left) presents the relative residual for the m u kM L2 (∂B + ) R evolution process at k = kM when only the single frequency is applied. It is clear that the reconstruction reaches some local minimum of the cost functional. The residual decreases and stagnates around 0.2 after 40 iterations. The corresponding image is shown in Figure 5 (right, dash line), which deviates greatly from the real profile. As pointed out previously, in this case the convergence to the global minimizer of the 66 0.6 0.4 0.2 0 −1 −0.5 0 0.5 1 −0.5 0 0.5 1 0.6 0.4 0.2 0 −1 Figure 4.2: Top: real profile. Bottom: reconstruction (dash line) at k = 16 compared with the real curve (solid line). cost functional strongly relies on the initial guess, which is hard to choose without a priori information of the imaging target. Multiple frequency data overcomes the difficulty, since the reconstruction at each lower frequency serves as the initial guess for the reconstruction at the higher frequency. Example 2. General smooth profile In this example, the surface displacement is represented by a general smooth function f = f1 + f2 (solid line in Figure 4.6), where    0.2(cos( 5π x ) − 1) + 0.05(1 + cos( 5π x + π)) x ∈ [− 4 , 0),  1  2 1 2 1 5  4 5π x ) − 1) f1 (x1 ) = x1 ∈ [0, 5 ], 0.2(cos( 2 1      0 elsewhere 67 1 0.9 relative error 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 wavenumber Figure 4.3: Relative error with respect to the wavenumber. and    exp  f2 (x1 ) =    0 1 5 ( 4 x1 )2 − 1 4 |x1 | < , 5 elsewhere. Here Γ+ := { (x1 , f ) | x1 ∈ (−R, R), f (x1 ) > 0 } and Γ− := { (x1 , f ) | x1 ∈ (−R, R), f (x1 ) < 0 } are both nonempty sets. We set kM = 30, and the final reconstruction with 5% noise in measurements is plotted in Figure 4.6 (dash line). Figure 4.7 shows the relative error for various wavenumbers. The proposed reconstruction method with multiple frequency data again gives a stable and accurate reconstruction of the general profile. However, compared with the profile considered in Example 1, the convergence rate decreases when the nonnegativity or non-positivity assumption is removed from the imaging target. (see Figure 4.3 and Figure 4.7). 68 0.6 0.6 0.4 0.4 0.2 0.2 0 0 −1 −0.5 0 0.5 1 −1 0.6 1 −0.5 0 0.5 1 −0.5 0 0.5 1 0.2 0 0.5 0.4 0.2 0 0.6 0.4 −0.5 0 −1 −0.5 0 0.5 1 −1 0.6 0.6 0.4 0.4 0.2 0.2 0 −1 −0.5 0 0.5 1 0 −1 Figure 4.4: Evolution of the reconstruction at k = 1, 5, 8, 10, 12, 16 (dash line, left to right, top to bottom). 69 0.8 Residual 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 10 20 30 40 50 60 70 Iterations 0.4 0.2 0 −0.2 −0.4 −1 −0.5 0 0.5 1 Figure 4.5: Top: relative residual with respect to the iteration number at k = kM when only single frequency data is used. Bottom: reconstruction (dash line) with single frequency only. 70 0.4 0.2 0 −0.2 −0.4 −1 −0.5 0 0.5 1 Figure 4.6: Reconstruction (dash line) at k = 30 compared with the real curve (solid line). Example 3. Piecewise linear profile The previous examples assume some regularity on the local surface displacement. Next we consider a piecewise linear profile (the solid line in Figure 4.8). By using our method, the reconstruction fn captures the position and height of bumps (where f > 0) accurately when kM = 16 (the dash line in Figure 4.8). The corner of the profile is also approximated with reasonable accuracy by the C 2 smooth reconstructed function. It will be interesting to develop a regularization strategy to approximate the corner with some non-smooth function, which will be investigated in our future work. Example 4. Multiscale profile The multiscale profile is represented by    f (x1 ) = 0.13 + 0.1 cos(   0 8π 3 1 x1 + π) + 0.03 cos(16πx1 + π) x1 ∈ [− , ], 5 4 2 elsewhere. It consists of two scales. The macroscale feature of the profile is represented by the 71 1.2 relative error 1 0.8 0.6 0.4 0.2 0 0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 wavenumber Figure 4.7: Relative error with respect to the wavenumber. 0.6 0.4 0.2 0 −1 −0.5 0 0.5 1 Figure 4.8: Reconstruction (dash line) at k = 16 compared with the real curve (solid line). 72 0.4 0.2 0 −1 −0.5 0 0.5 1 −0.5 0 0.5 1 −0.5 0 0.5 1 −0.5 0 0.5 1 0.4 0.2 0 −1 0.4 0.2 0 −1 0.4 0.2 0 −1 73 Figure 4.9: Reconstruction at k = 10, 20, 33, 40 (dash line, top to bottom) compared with the real curve (solid line). 8π x + π) ; while the microscale is represented by function 5 1 1 0.03 cos(16πx1 + π) with period . The reconstruction captures the macroscale fea8 tures when k = 10 (Figure 4.9, top left). To recover the fine details of the profile with 1 the period , k needs to be sufficiently high. Here, microscale features are captured 8 when k = 40 (Figure 4.9, bottom right). The whole local disturbance is accurately function 0.13 + 0.1 cos( reconstructed with noisy data. On the other hand, it is observed that the resolution of the reconstruction does not improve much from k = 10 to k = 33. This is due to the fact that other than the macroscale feature, no scale length of the profile is comparable with the corresponding wavelength for k ∈ (10, 33). 74 Chapter 5 Conclusion This thesis is the initial endeavor in the whole picture of the near-field imaging. We focus on one specific problem, where the local surface displacement on an infinite ground plane is the imaging target. The main contribution is the explicit formulation of the connection between the evanescent wave modes and the high frequency components of the surface displacement, and a new numerical scheme to reconstruct the surface displacement that extracts the information carried by the evanescent modes effectively. Numerical examples show that images with a resolution of λ/10 are obtained. For the general local surface displacement, a reconstruction scheme with multiple frequency data is proposed that captures the main (large scale) features at low frequencies and recovers the fine details at high frequencies. Numerical evidences confirm that the resolution in the near-field regime can be significantly improved. However, theoretical study on the uniqueness and the stability estimates remain open. In particular, the stability estimate which is able to incorporates explicitly the dependence on the distance d would help to understand the ill-posedness of the inverse scattering problem better in the near field. There are also many related imaging problems that arise in biological and nanosciences. For example, the imaging of the human cancer cell in the near-field regime, 75 which turns out to be an inverse medium scattering problem. Though the evanescent wave modes carried by the scattered field is available in the near-field measurement, the relation between such evanescent wave modes and the high frequency components of the medium is completely open. Such relation is the key to design efficient numerical algorithms that make full use of the evanescent waves at hand. On the other hand, when multiple frequency data is available, one theoretically challenging problem is to investigate the stability for the inverse scattering problem. It is believed that a H¨lder type stability estimate can be obtained with sufficiently o large band of frequencies. 76 BIBLIOGRAPHY 77 BIBLIOGRAPHY [1] E. Abbe, Beitr¨ge zur Theorie des Mikroskops und der mikroskopischena Wahrnehmung, Archivf. Miroskop. Anat., 9, (1873), 413. [2] M. Abramowitz and I. Stegun (Eds.), Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables, NBS Applied Mathematics Series 55, National Bureau of Standards, Washington, DC (1964). [3] H. Ammari, G. Bao, and A.W. Wood, An integral equation method for the electromagnetic scattering from cavities, Math. Meth. Appl. Sci. 23 (2000), 10571072. [4] H. Ammari, G. Bao, and A.W. Wood, Analysis of the electromagnetic scattering from a cavity, Japan J. Indust. Appl. Math. 19 (2002), 301-310. [5] K. Belkebir, P Chaumet, and A. Sentenac, Superresolution in total internal reflection tomography, J. Opt. Soc. Am. A, 22 (2005), 1889-1897. [6] K. Belkebir and A. Sentenac, High-resolution optical diffraction microscopy, J. Opt. Soc. Am. A, 20 (2003), 1223-1229. [7] G. Bao and P. Li, Inverse medium scattering for three-dimensional time harmonic Maxwell equations, Inverse Problems, 20 (2004), L1-L7. [8] G. Bao and P. Li, Inverse medium scattering problems for electromagnetic waves, SIAM J. Appl. Math., 65 (2005), 2049-2066. [9] G. Bao and P. Li, Inverse medium scattering problems in near-field optics, J. Comput. Math., 25 (2007), 252-265. [10] G. Bao and J. Lin, Near-field imaging of the surface displacement on an infinite ground plane, submitted. [11] G. Bao and J. Lin, Imaging of local surface displacement on an infinite ground plane: the multiple frequency case, submitted. 78 [12] G. Bao, J. Lin, and F. Triki, A multi-frequency inverse source problem, J. Differential Equations, 249 (2010), 3443-3465. [13] G. Bao and F. Triki, Error estimates for the recursive linearization for solving inverse medium problems, J. Comput. Math., 28 (2010), 725-744. [14] J. Baumeister, Stable Soltions of Inverse Problems, Vieweg-Verlag, Braunschweig, (1987). [15] M. Bertero, P. Boccacci and M. Piana, Resolution and super-resolution in inverse diffraction, Lecture Notes in Physics 486 (1997), 1-17. [16] H. Bialy, Iterative Behandlung linearer Funktionalgleichungen, Arch. Rat. Mech. Analy. 4 (1959), 166. [17] M. Born and E. Wolf, Principles of Optics (6th ed.), Cambridge Univ. Press, 1980. [18] R. Carminati and J. Greffet, Influence of dielectric contrast and topography on the near field scattered by an inhomogeneous surface, J. Opt. Soc. Am. A, 12 (1995), 2716-2725. [19] P. Carney and J. Schotland, Inverse scattering for near-field microscopy, Appl. Phys. Lett. 77 (2000), 2798-800. [20] P. Carney and J. Schotland, Three-dimensional total internal reflection microscopy, Opt. Lett. 26 (2001), 1072-1074. [21] P Chaumet, K. Belkebir, and A. Sentenac, Numerical study of grating-assisted optical diffraction tomography, Phy. Rev. A, 76 (2007), 013814. [22] Y. Chen, Inverse scattering via Heisenberg’s uncertainty principle, Inverse Problems, 13 (1997), 1-13. [23] D. Colton and R. Kress, Integral Equation Methods in Scattering Theory, Pure and Applied Mathematics, Wiley, New York (1983). [24] D. Colton and R. Kress, Inverse Acoustic and Electromagnetic Scattering Theory, Applied Mathematical Sciences, Vol. 93, Springer-Verlag, Berlin (1998). [25] D. Courjon and C. Bainier, Near field microscopy and near field optics, Rep. Prog. Phys. 57 (1994), 989-1028. [26] D. Courjon, K. Sarayeddine, and M. Spajer, Scanning tunneling optical microsocpy, Opt. Commun. 71 (1989), 23-28. [27] J. W. Daniel, The conjugate gradient emthod for linear and nonlinear operator equaitons, SIAM J. Numer. Anal. 4 (1967), 10-26. 79 [28] C. de Boor, A Practical Guide to Splines, Applied Mathematical Sciences, vol. 27, Springer-Verlag, New York (2001). [29] G. Derveaux, G. Papanicolaou and C. Tsogka, Resolution and denoising in near-field imaging, Inverse Problems 22 (2006), 1437-1456. [30] R. Dunn, Near-Field Scanning Optical Microscopy, Chem. Rev. 99 (1999), 28912927. [31] H. W. Engl, M. Hanke, and A. Neubauer, Regularization of Inverse Problems, Mathematics and Its Application, Kluwer Academic Pubishers, New York (1996). [32] H. W. Engl, K. Kunisch, and A. Neubauer,, Convergence rates for Tikhonov Regularization of nonliear ill-posed problems, Inverse Problems, 5 (1989) 541547. [33] B. Engquist and A. Majda, Absorbing boundary conditions for the numerical simulation of waves, Math. Comput., 31 (1977), 629-651. [34] L. C. Evans, Partial Differential Equations, Graduate Studies in Mathematics, Vol. 19, American Mathematical Society (1997). [35] V. Fridman, A method of successive approximations for Fredholm integral equations of the first kind, Uspki Mat. Nauk., 11 (1956), 233-234. [36] J. Graves and P. M. Prenter, Numerical iterative filters applied to first Freholm integral equations, Numer. Math., 30 (1978), 281-299. [37] C. W. Groetsch, The Theory of Tikhonov Regularization for Fredholm Equations of the First Kind. Pitman, Boston (1984). [38] J. Hadamard, Lectures on Cauchy’s problem in Linear Partial Differential Equations, Yale University Press, New Haven (1923). [39] M. Hanke, A. Neubauer, and O. Scherzer A convergence analysis of the Landweber iteration for nonlinear ill-posed problems, Numer. Math., 72 (1995), 21-37. [40] V. Isakov, Increasing stability for the Schr¨dinger potential from the Dirichleto to-Neumann map, Discrete Contin. Dynam. Systems-S, to appear. [41] D. Jerison and C. Kenig, Unique continuation and absence of positive eigenvalues for Schr¨dinger operators. Ann. Math., 121(1985), 463-488. o [42] J. Jin, The Finite Element Method in Electromagnetics (2nd ed.), Jonh Wiley & Sons (2002). [43] A. Kirsch, An Introduction to the Mathematical Theory of Inverse Problems, Applied Mathematical Sciences, Vol. 120, Springer-Verlag, New York, 1996. 80 [44] A. Kirsch, The domain derivative and two applications in inverse scattering theory, Inverse Problems, 16 (1993), 81-96. [45] R. Kress, Linear Integral Equations, Applied Mathematical Sciences, Vol. 82, Sringer-verlag, New York (1999). [46] R. Kress and T. Tran, Inverse scattering for a locally perturbed half-plane, Inverse Problems 16 (2000), 1541-1559. [47] D. Labeke and D. Barchiesi, Scanning-tunneling optical microscopy: a theoretical macroscopic approach, J. Opt. Soc. Am. A 9 (1992), 732-739. [48] L. Landweber, An iteration formula for Freholm integral equations of the first kind, Am. J. Math. 73 (1951), 615-624. [49] V. A. Morozov, On the soltuion of functional equations by the method of regularization, Soviet Math. Dokl 7 (1966), 414-417. [50] A. Neubauer, Tikhonov Regularization for nonliear ill-posed problems: optimal convergence and finite-dimensional approximation, Inverse Problems, 5 (1989) 541-547. [51] L. Novotny and B. Hecht, Principles of Nano-Optics, Cambridge University Press (2006). [52] O. Pironneau, Optimal Shape Design for Elliptic Systems, Springer-Verlag, New York (1984). [53] L. Rayleigh, On the theory of optical images with special reference to the optical microscope, Phil. Mag. 5 (1896), 167-195. [54] R. C. Reddick, R. J Warmack, D. W. Chilcott, S. L. Sharp , and T. L. Ferrel, Photon scanning tunneling microscopy, Rev. Sci. Instrum. 61 (1990) 3669-3677. [55] T. I. Seidman and C. R. Vogel, Well-posedness and convergence of some regularization methods for nonlinear ill-posed problems, Inverse Problems, 5 (1989) 227-238. [56] J. Sokolowski and J. Zolesio, Introduction to Shape Optimization: Shape Sensitivity Analysis, Springer-Verlag, Berlin (1992). [57] J. Sun, P. Carney, and J. Schotland, Near-field scanning optical tomography: a nondestructive method for three-dimensional nanoscale imaging, IEEE J. Sel. Top. Quant. 12 (2006), 1072-1082. [58] S. Twomey, The application of numerical filtering to the solution of integral equations encountered in indirect sensing measurements, J. Franklin Inst. 279 (1965), 95-109. 81 [59] A. V. Tikhonov, On the solution of incorrectly formulated problems and the regularization method, Soviet Math. Doklady, 4 (1963), 1035-1038. [60] A. V. Tikhonov, Regularization of incorectly posed problems, Soviet Math. Doklady, 4 (1963), 1624-1627. [61] A. Willers, The Helmholtz equation in disturbed half-spaces, Math. Meth. Appl. Sci. 9 (1987), 312-323. [62] E. Wolf (Ed.) Progress in Optics 50, Elsevier, Amsterdam, The Netherlands (2007). [63] A. Wood, Analysis of electromagnetic scattering from an overfilled cavity in the ground plane, J. Comput. Phys., 215 (2006), 630-641. [64] B. Zhang and S. N. Chandler-Wilde, Integral equation methods for scattering by infinite rough surfaces, Math. Meth. Appl. Sci. 26 (2003), 463-488. 82