0712.0175/QR1.tex
1: 
2: \documentclass[12pt]{article}
3: \usepackage{amssymb}
4: 
5: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6: \usepackage{graphicx}
7: \usepackage{amsmath}
8: \usepackage{subfigure}
9: \usepackage{calc}
10: \usepackage{float}
11: 
12: %TCIDATA{OutputFilter=LATEX.DLL}
13: %TCIDATA{Created=Fri Sep 07 08:44:56 2001}
14: %TCIDATA{LastRevised=Thu Nov 29 20:47:32 2007}
15: %TCIDATA{<META NAME="GraphicsSave" CONTENT="32">}
16: %TCIDATA{<META NAME="DocumentShell" CONTENT="General\Blank Document">}
17: %TCIDATA{Language=American English}
18: %TCIDATA{CSTFile=LaTeX article (bright).cst}
19: 
20: \setcounter{MaxMatrixCols}{10}
21: \newtheorem{theorem}{Theorem}
22: \newtheorem{acknowledgement}[theorem]{Acknowledgement}
23: \newtheorem{algorithm}[theorem]{Algorithm}
24: \newtheorem{axiom}[theorem]{Axiom}
25: \newtheorem{case}[theorem]{Case}
26: \newtheorem{claim}[theorem]{Claim}
27: \newtheorem{conclusion}[theorem]{Conclusion}
28: \newtheorem{condition}[theorem]{Condition}
29: \newtheorem{conjecture}[theorem]{Conjecture}
30: \newtheorem{corollary}[theorem]{Corollary}
31: \newtheorem{criterion}[theorem]{Criterion}
32: \newtheorem{definition}[theorem]{Definition}
33: \newtheorem{example}[theorem]{Example}
34: \newtheorem{exercise}[theorem]{Exercise}
35: \newtheorem{lemma}{Lemma}
36: \newtheorem{notation}[theorem]{Notation}
37: \newtheorem{problem}[theorem]{Problem}
38: \newtheorem{proposition}[theorem]{Proposition}
39: \newtheorem{remark}[theorem]{Remark}
40: \newtheorem{solution}[theorem]{Solution}
41: \newtheorem{summary}[theorem]{Summary}
42: \newenvironment{proof}[1][Proof]{\textbf{#1.} }{\ \rule{0.5em}{0.5em}}
43: \input{tcilatex}
44: \oddsidemargin=0in
45: \evensidemargin=0in
46: \textwidth=6.5in
47: \textheight=8.5in
48: \topmargin=0in
49: \input{tcilatex}
50: 
51: \begin{document}
52: 
53: \title{\textbf{The Quasi-Reversibility Method for the Thermoacoustic Tomography and
54: a Coefficient Inverse Problem}}
55: \author{Michael V. Klibanov$^{\ast }$, Andrey V. Kuzhuget$^{\ast },$ \and Sergey I.
56: Kabanikhin$^{\ast \ast }$, and Dmitriy V.\ Nechaev$^{\bigtriangledown }$ \\
57: %EndAName
58: $^{\ast }$Department of Mathematics and Statistics\\
59: University of North Carolina at Charlotte,\\
60: Charlotte, NC 28223, USA\\
61: $^{\ast \ast }$ Sobolev Institute of Mathematics\\
62: of the Siberian Branch \\
63: of the Russian Academy of Science\\
64: Prospect Acad. Koptyuga 2,\\
65: Novosibirsk, 630090, Russia\\
66: $^{\bigtriangledown }$ Lavrent'ev Institute of Hydrodynamics \\
67: of the Siberian Branch \\
68: of the Russian Academy of Science\\
69: Prospect Acad. Lavrent'eva 15\\
70: Novosibirsk, 63090, Russia\\
71: E-mails: mklibanv@uncc.edu; akuzhuge@uncc.edu; \\
72: kabanikh@math.nsc.ru; nechaev@hydro.nsc.ru}
73: \maketitle
74: 
75: \begin{abstract}
76: An inverse problem of the determination of an initial condition in a
77: hyperbolic equation from the lateral Cauchy data is considered. This problem
78: has applications to the thermoacoustic tomography, as well as to linearized
79: coefficient inverse problems of acoustics and electromagnetics. A new
80: version of the quasi-reversibility method is described.\ This version
81: requires a new Lipschitz stability estimate, which is obtained via the
82: Carleman estimate. Numerical results are presented.
83: \end{abstract}
84: 
85: \bigskip \textbf{KEY WORDS}: Quasi-reversibility method, Carleman estimate,
86: numerical results, imaging of sharp peaks
87: 
88: \bigskip \textbf{AMS subject classification:} 65N21, 65D10, 65F10
89: 
90: \section{Introduction}
91: 
92: \bigskip In this paper we propose a new version of the Quasi-Reversibility
93: Method (QRM) for the inverse problem of the determination of an initial
94: condition in a hyperbolic equation from the lateral Cauchy data.\ We discuss
95: applications of this inverse problem to thermoacoustic tomography, as well
96: as to linearized coefficient inverse problems of acoustics and
97: electromagnetics. Using the Carleman estimate, we prove convergence of our
98: version of the QRM.\ We also present numerical results. In particular, we
99: show that this version of the QRM enables one to image $\delta -$ like
100: functions, i.e., narrow high peaks.
101: 
102: Let $\Omega \subset \mathbb{R}^{n}$ be a convex domain with a piecewise
103: smooth boundary $\partial \Omega $ and $2R$ be the diameter of $\Omega
104: ,2R=\max_{x,y\in \Omega }$ $\left| x-y\right| .$ Let $T=const.>R.$ Denote $%
105: Q_{T}=\Omega \times \left( 0,T\right) .$ Consider the elliptic operator $%
106: L(x,t)$ of the form
107: \begin{equation*}
108: L(x,t)u=\Delta u+\sum\limits_{j=1}^{n}b_{j}\left( x,t\right)
109: u_{j}+b_{0}\left( x,t\right) u_{t}+c\left( x,t\right) u,
110: \end{equation*}
111: where $u_{j}:=\partial _{x_{j}}u.$ We assume that all coefficients of the
112: operator $L$ belong to $C\left( \overline{Q}_{T}\right) .$ Let the function $%
113: u\in H^{2}\left( Q_{T}\right) $ be a solution of the hyperbolic equation in
114: the cylinder $Q_{T},$%
115: \begin{equation}
116: u_{tt}=L(x,t)u+F\left( x,t\right) \text{ in }Q_{T},  \tag{1.1}
117: \end{equation}
118: $F\in L_{2}\left( Q_{T}\right) $ with initial conditions
119: \begin{equation}
120: u\left( x,0\right) =\varphi \left( x\right) ,u_{t}\left( x,0\right) =\psi
121: \left( x\right) ,\varphi \in H^{1}\left( \Omega \right) ,\psi \in
122: L_{2}\left( \Omega \right) .  \tag{1.2}
123: \end{equation}
124: We consider the following
125: 
126: \textbf{Inverse Problem 1.} Let one of functions $\varphi $ or $\psi $ be
127: known and another one be unknown. Determine that unknown function assuming
128: that the following functions $f$ and $g$ are given
129: \begin{equation}
130: u\mid _{S_{T}}=f\left( x,t\right) ,\text{ }\frac{\partial u}{\partial \nu }%
131: \mid _{S_{T}}=g\left( x,t\right) ,\text{ }S_{T}=\partial \Omega \times
132: \left( 0,T\right) ,  \tag{1.3}
133: \end{equation}
134: where $\nu $ is the unit outward normal vector at $\partial \Omega .$ We
135: call the problem of the determination of the function $\varphi $ the ``$%
136: \varphi -$problem'' and the problem of the determination of the function $%
137: \psi $ the ``$\psi -$problem''.
138: 
139: In principle, in the case $T>2R$ one should not assume that one of functions
140: $\varphi $ or $\psi $ is known.\ This is because for $T>2R$ the following
141: Lipschitz stability estimate takes place (see [4], [10], [11] and Theorem
142: 2.4.1 in [13])
143: \begin{equation}
144: \left\| u\right\| _{H^{1}\left( Q_{T}\right) }\leq C\left( \left\| f\right\|
145: _{H^{1}\left( S_{T}\right) }+\left\| g\right\| _{L_{2}\left( S_{T}\right)
146: }+\left\| F\right\| _{L_{2}\left( Q_{T}\right) }\right) .  \tag{1.4}
147: \end{equation}
148: Here and below $C$ denotes different positive constants depending only on $%
149: \Omega ,T$ and $C\left( \overline{Q}_{T}\right) $ norms of coefficients of
150: the operator $L$. However, since numerical studies for the case of the
151: finite domain were conducted in previous publications [4], [12], we are
152: interested here in solving the Inverse Problem 1 in an \emph{unbounded}
153: domain, which was not done before. This leads us to the case $T>R.$ Namely,
154: we want to solve an analogue of the Inverse Problem 1 in a quadrant,
155: assuming that the lateral Cauchy data are given only on parts of two
156: coordinate axis. We are motivated by the publication [14], where the
157: Lipschitz stability was proven for an analogue of Inverse Problem 1 for the
158: case of either a quadrant in $\mathbb{R}^{2}$ or a an octant in $\mathbb{R}%
159: ^{3}$, assuming that one of initial conditions (1.2) is zero, and the second
160: one has a finite support.
161: 
162: We now specify conditions of our numerical study. Suppose that equation
163: (1.1) is homogeneous with $F\left( x,t\right) \equiv 0$ and it is satisfied
164: in $D_{T}^{3}=\mathbb{R}^{2}\times \left( 0,T\right) .$ Consider the
165: quadrant $QU=\left\{ x_{1},x_{2}>0\right\} .$ And also consider the square $%
166: SQ\subset QU,$%
167: \begin{equation*}
168: SQ\left( a\right) =\left\{ 0<x_{1},x_{2}<a\right\} .
169: \end{equation*}
170: Suppose that
171: \begin{equation}
172: \varphi (x)=\psi \left( x\right) =0\text{ outside of }SQ\left( a\right) .
173: \tag{1.5}
174: \end{equation}
175: Then the energy estimate implies that
176: \begin{equation}
177: u\left( x,t\right) =0\text{, }\forall \left( x,t\right) \in \left\{ x\mid
178: x\in QU,dist\left( x,SQ\left( a\right) \right) >T\right\} \times \left(
179: 0,T\right) .  \tag{1.6}
180: \end{equation}
181: Denote
182: \begin{equation*}
183: \Gamma _{1T}=\left\{ x_{1}\in \left( 0,a+T\right) ,x_{2}=0\right\} \times
184: \left( 0,T\right) ,
185: \end{equation*}
186: \begin{equation*}
187: \Gamma _{2T}=\left\{ x_{2}\in \left( 0,a+T\right) ,x_{1}=0\right\} \times
188: \left( 0,T\right) ,
189: \end{equation*}
190: \begin{equation*}
191: \Gamma _{3T}=\left\{ x_{1}=a+T,x_{2}\in \left( 0,a+T\right) \right\} \times
192: \left( 0,T\right) ,
193: \end{equation*}
194: \begin{equation*}
195: \Gamma _{4T}=\left\{ x_{2}=a+T,x_{1}\in \left( 0,a+T\right) \right\} \times
196: \left( 0,T\right) ,
197: \end{equation*}
198: see Figure $\ref{fig:geometry}$. Then by (1.6)
199: \begin{figure}[tbp]
200: \begin{center}
201: \includegraphics[scale=0.5]{geometry.eps}
202: \end{center}
203: \caption{Geometry for the Inverse problem 2.}
204: \label{fig:geometry}
205: \end{figure}
206: \begin{equation}
207: u=\frac{\partial u}{\partial \nu }=0\text{ on }\Gamma _{3T}\cup \Gamma _{4T}.
208: \tag{1.7}
209: \end{equation}
210: Hence, we focus our numerical study on
211: 
212: \textbf{Inverse Problem 2.} Let equation (1.1) be satisfied in $D_{T}^{3}$
213: with initial conditions (1.2) satisfying (1.4). In this case $\Omega
214: :=SQ\left( a+T\right) .$Suppose that one of these initial conditions is
215: zero. Determine the second initial condition, assuming that functions $f$
216: and $g$ are known, where
217: \begin{equation}
218: u\mid _{\Gamma _{1T}\cup \Gamma _{2T}}=f\left( x,t\right) ,\text{ }\frac{%
219: \partial u}{\partial \nu }\mid _{\Gamma _{1T}\cup \Gamma _{2T}}=g\left(
220: x,t\right) .  \tag{1.8}
221: \end{equation}
222: 
223: Suppose for a moment that only the function $f\left( x,t\right) $ is given.\
224: Then one can solve the boundary value problem for equation (1.1) with $%
225: F\equiv 0$ outside of the square $SQ\left( a+T\right) $ with the following
226: initial and boundary data
227: \begin{equation*}
228: u\left( x,0\right) =u_{t}\left( x,0\right) =0,x\in R^{2}\diagdown SQ(a+T),
229: \end{equation*}
230: \begin{equation*}
231: u\mid _{\Gamma _{1T}\cup \Gamma _{2T}}=f\left( x,t\right) ,u\mid _{\Gamma
232: _{3T}\cup \Gamma _{4T}}=0\text{.}
233: \end{equation*}
234: This gives one in a \emph{stable way} the normal derivative $g\left(
235: x,t\right) $ on $\Gamma _{1T}\cup \Gamma _{2T}.$ Thus, we arrive at Inverse
236: Problem 2. It was proven in [14] that if
237: \begin{equation}
238: T>\frac{a\sqrt{2}}{2-\sqrt{2}}  \tag{1.9}
239: \end{equation}
240: and one of functions $\varphi $ or $\psi $ equals zero, then the following
241: Lipschitz stability estimate is valid
242: \begin{equation}
243: \left\| u\right\| _{H^{1}\left( G_{T}\right) }\leq C\left( \left\| f\right\|
244: _{H^{1}\left( \Gamma _{T}\right) }+\left\| g\right\| _{L_{2}\left( \Gamma
245: _{T}\right) }\right) ,  \tag{1.10}
246: \end{equation}
247: where $\Gamma _{T}=\Gamma _{1T}\cup \Gamma _{2T}$ and $\left\| f\right\|
248: _{H^{1}\left( \Gamma _{T}\right) }=\left\| f\right\| _{H^{1}\left( \Gamma
249: _{1T}\right) }+\left\| f\right\| _{H^{1}\left( \Gamma _{2T}\right) }.$ The
250: estimate (1.10) implies a similar estimate for the unknown initial condition
251: [14]. The knowledge of the fact that one of initial conditions was zero was
252: used in [14] for either odd or even extension with respect to $t$ of the
253: function $u(x,t)$ in $\left\{ t<0\right\} ,$ depending on which of initial
254: conditions was assumed to be unknown$.$ The proof of [14] is based on the
255: Carleman estimate. The method of Carleman estimates was first applied in
256: [11] to obtain the Lipschitz stability for the hyperbolic problem with the
257: lateral Cauchy data, also see [10] and Theorem 2.4.1 in [13]. Prior to [11]
258: the Lipschitz stability for the hyperbolic problem with the lateral Cauchy
259: data was obtained in [20] by the method of multipliers, but only for the
260: case when lower order terms in (1.1) are absent. The use of the Carleman
261: estimate enables one to incorporate lower order terms and also to extend to
262: the case of hyperbolic inequalities. Recently the method of [10], [11], [13]
263: was extended to hyperbolic equations with the non-constant principal part,
264: see, e.g. [22]-[24]. The method of multipliers was recently extended to the
265: case of non-zero lower order terms in Theorem 3.5 of the book [6].
266: 
267: The above problems were previously solved numerically in [4], [7], [12] and
268: [15]. The work [7] was the first one, where the problem of thermoacoustic
269: tomography was formulated and solved numerically as Inverse Problem 1, i.e.,
270: as the hyperbolic Cauchy problem with the lateral data. The QRM for the
271: latter problem was used in [4]. The QRM was first proposed in the book [19]
272: for a variety of ill-posed boundary value problems. Its convergence rates
273: were established in [9] and [11] for the cases of Laplace and hyperbolic
274: equations respectively and in section 2.5 of [13] for elliptic, parabolic
275: and hyperbolic equations. In particular, it was shown in [13] that one can
276: work with weak $H^{2}$ solutions of QRM instead of strong $H^{4}$ solutions
277: of the original book [19]. Also, see [2] and [3] for the recent results for
278: the QRM for the elliptic case and [8] for the application of the QRM to
279: linearized coefficient inverse problems for parabolic equations. The main
280: tool of works [4], [8], [9], [11] and [13] is the tool of Carleman estimates.
281: 
282: There are three main differences between the current paper and the previous
283: works on the QRM for hyperbolic equations. First, we take into account
284: boundary conditions via including them in the Tikhonov regularizing
285: functional $J_{\varepsilon }$. Unlike this, boundary conditions were made
286: zero in [4] via subtracting off a corresponding function, and they were
287: treated via integration by parts in [12]. Second, we incorporate in $%
288: J_{\varepsilon }$ a penalizing term, which reflects our knowledge of one of
289: initial conditions. We show numerically that without this term we cannot
290: image well maximal values of the unknown initial condition inside of narrow
291: peaks. On the other hand, since cancerous tumors can be modeled as narrow
292: peaks, it is interesting to image those peaks in the application to
293: thermoacoustic tomography considered below. These first two ideas for $%
294: J_{\varepsilon }$ are taken from [15]. Mainly because of the second
295: difference we cannot apply previously derived convergence results and thus,
296: need to prove convergence of our new version of the QRM. In particular, we
297: need to prove a \ new Lipschitz stability estimate (Theorem 4.1). Finally,
298: the third difference is that while $H^{2}$ finite elements were used in [4]
299: and [12], we use finite differences now. Note that while smooth slowly
300: varying functions were reconstructed numerically in [4] and [12], our
301: numerical experiments reconstruct both those functions and $\delta -$like
302: functions. $\delta -$like functions were also reconstructed in [15] for the
303: Inverse Problem 2. However, the numerical technique of [15] is different
304: from one of the current paper. The method of [7] and [15] is based on the
305: representation of the function $u(x,t)$ via truncated Fourier series and
306: minimization of the resulting residual least squares functional.
307: 
308: In section 2 we describe applications of Inverse Problems 1 and 2. In
309: section 3 we describe the version of the QRM we use here. In section 4 we
310: prove a new Lipschitz stability estimate. In section 5 we prove convergence
311: of our method, based on the result of section 4. In section 6 we describe
312: our numerical implementation. In section 7 numerical results are presented.
313: Conclusions are drawn in section 8.
314: 
315: \section{Applications}
316: 
317: In this section we discuss two applications of above inverse problems
318: 
319: \subsection{Thermoacoustic tomography}
320: 
321: Inverse Problems 1 and 2 arise in thermoacoustic tomography [4], [16], [25].
322: In \ this case the target is subjected to a short electromagnetic impulse.
323: The electromagnetic energy is absorbed. As a result, temperature is
324: increased and the target is expanded. This causes a pressure wave, which is
325: measured as a change in the acoustic field at the boundary of the sample.
326: Assuming that the absorption of the electromagnetic energy is spatially
327: varying inside the sample, the resulting wave field is carrying the
328: signature of the inhomogeneity. On the other hand, cancerous regions absorb
329: more than surroundings. This leads to applications in medical imaging.
330: Hence, the problem is to calculate the absorption coefficient $\alpha (x)$
331: of the sample using time dependent measurements at its boundary. Let $\beta $
332: be the thermal expansion coefficient, $c_{p}$ be the specific capacity of
333: the medium and $I_{0}$ be the power of the source. Usually $\beta ,c_{p}$
334: and $I_{0}$ are known. Also, assume that the speed of sound in the medium is
335: constant and equals 1. Let $u(x,t)$ be the pressure wave. It was shown in
336: e.g., [4] that
337: \begin{equation}
338: u_{tt}=\Delta u,\left( x,t\right) \in \mathbb{R}^{3}\times \left( 0,T\right)
339: ,  \tag{2.1}
340: \end{equation}
341: \begin{equation}
342: u\left( x,0\right) =\alpha (x)I_{0}\frac{\beta }{c_{p}},\text{ \ }%
343: u_{t}\left( x,0\right) =0.  \tag{2.2}
344: \end{equation}
345: Suppose that we measure the function $u(x,t)$ at the boundary of the domain $%
346: \Omega $ and $\alpha (x)=0$ outside of $\Omega .$ Then those measurements
347: give us the boundary value problem for equation (2.1) outside of $\Omega $
348: with zero initial conditions. Solving this problem, we uniquely determine
349: the normal derivative of the function $u(x,t)$ at $\partial \Omega .$ Thus,
350: we arrive at the $\varphi -$ problem.
351: 
352: A different approach to the problem of thermoacoustic tomography is
353: currently actively developed in a number of publications. In this approach
354: the solution of the problem (2.1), (2.2) is presented via the
355: Poisson-Kirchhoff formula, which leads to the problem of integral geometry
356: of recovering a function via its integrals over certain spheres, whose
357: centers run over a surface and radii vary. Then uniqueness theorems are
358: proven for this case and inversion formulas are derived, see, e,g., [1],
359: [5], [16], and [17]. In particular, works [1] and [16] include the case of a
360: variable speed and [16] and [17] include numerical examples. A survey of
361: these developments can be found in [16]. Also, see \S 1 of Chapter 6 of the
362: book [18] for an example of the ill-posedness of this integral geometry
363: problem for the case when centers of spheres run over a plane in $\mathbb{R}%
364: ^{3}$.
365: 
366: \subsection{Linearized inverse acoustic and electromagnetic problems}
367: 
368: There is also another application, in which Inverse Problems 1 and 2 can be
369: considered as linearized inverse acoustic and inverse electromagnetic
370: problems. The idea of this subsection is motivated by \S 1 of Chapter 7 of
371: [18] and \S 3 of Chapter 2 of [21]. We present this application now without
372: discussing delicate details about the validity of the linearization. In this
373: setting the point source is running along a surface and time dependent
374: measurements of back-reflected data are performed at the positions of the
375: source. In [11] the Newton-Kantorovich method was presented for the case
376: when the source position is fixed and the time dependent measurements are
377: performed at a surface.
378: 
379: Let the function $\alpha \left( x\right) \in C\left( \mathbb{R}^{3}\right) $
380: be strictly positive, $\alpha \left( x\right) \geq const.>0.$ Consider the
381: Cauchy problem for the hyperbolic equation
382: \begin{equation}
383: \alpha \left( x\right) w_{tt}=\Delta _{x}w+4\pi \delta \left(
384: x-x_{0},t\right) ,\left( x,t\right) \in \mathbb{R}^{3}\times \left(
385: 0,T\right) ,  \tag{2.3}
386: \end{equation}
387: \begin{equation}
388: w\left( x,x_{0},0\right) =w_{t}\left( x,x_{0},0\right) =0,  \tag{2.4}
389: \end{equation}
390: where $x_{0}\in \mathbb{R}^{3}$ is the source position. It is well known
391: that in acoustics $\alpha \left( x\right) =c^{-2}\left( x\right) ,$ where $%
392: c\left( x\right) $ is the speed of sound in the medium, and in some
393: situations of the electromagnetics $\alpha \left( x\right) =\left( \mu
394: \epsilon \right) \left( x\right) ,$ where $\mu $ and $\epsilon $ are
395: respectively magnetic permeability and electric permittivity of the medium.
396: We pose the following
397: 
398: \textbf{Inverse Problem 3.} Suppose that the function $\alpha \left(
399: x\right) =1$ outside of the domain $\Omega $ and it is unknown inside of
400: this domain. Determine this function for $x\in \Omega ,$ assuming that the
401: following function $p\left( x_{0},t\right) $ is known
402: \begin{equation}
403: w\left( x_{0},x_{0},t\right) \mid _{x_{0}\in \partial \Omega }=p\left(
404: x_{0},t\right) .  \tag{2.4}
405: \end{equation}
406: 
407: The full Inverse Problem 3 is difficult to address because of its
408: nonlinearity. Hence, we consider now a linearized problem. Similarly with \S
409: 3 of Chapter 2 of [21], suppose that the function $\alpha \left( x\right) $
410: can be represented in the form
411: \begin{equation*}
412: \alpha \left( x\right) =1-\xi a\left( x\right) ,
413: \end{equation*}
414: where $\xi \in \left( 0,1\right) $ is a small parameter. Hence, the term $%
415: \xi a\left( x\right) $ is a small perturbation of 1. We assume that this
416: perturbation is unknown, i.e., the function $a\left( x\right) $ is unknown.
417: Again, similarly with [21], we can formally set at $\xi \rightarrow 0$
418: \begin{equation}
419: w\left( x,x_{0},t\right) =w_{0}\left( x,x_{0},t\right) +\xi w_{1}\left(
420: x,x_{0},t\right) +O\left( \xi ^{2}\right) ,  \tag{2.5}
421: \end{equation}
422: where functions $w_{0}$ and $w_{1}$ are independent on $\xi .$ This setting
423: was rigorously justified in \S 3 of Chapter 2 of [21] for the case of the
424: telegraph equation
425: \begin{equation}
426: w_{tt}=\Delta w+\left( a_{0}\left( x\right) +\xi a_{1}\left( x\right)
427: \right) w.  \tag{2.6}
428: \end{equation}
429: It was also justified in \S 1 of Chapter 7 of [18]for equation (2.6) without
430: the introduction of the parameter $\xi ,$ which is actually introduced here
431: for convenience only. Indeed, instead, one can assume that $\alpha \left(
432: x\right) =1-a\left( x\right) ,$ where $\left| a\left( x\right) \right| <<1.$
433: 
434: Substituting (2.6) in (2.3) and (2.4) and dropping the term with $O\left(
435: \xi ^{2}\right) ,$ we obtain that functions $w_{0}$ and $w_{1}$ are
436: solutions of the following Cauchy problems
437: \begin{equation}
438: w_{0tt}=\Delta _{x}w_{0}+4\pi \delta \left( x-x_{0},t\right) ,\left(
439: x,t\right) \in \mathbb{R}^{3}\times \left( 0,T\right) ,  \tag{2.7}
440: \end{equation}
441: \begin{equation}
442: w_{0}\left( x,x_{0},0\right) =w_{0t}\left( x,x_{0},0\right) =0,  \tag{2.8}
443: \end{equation}
444: \begin{equation}
445: w_{1tt}=\Delta _{x}w_{1}+a\left( x\right) w_{0tt}\left( x,x_{0},t\right)
446: ,\left( x,t\right) \in \mathbb{R}^{3}\times \left( 0,T\right) ,  \tag{2.9}
447: \end{equation}
448: \begin{equation}
449: w_{0}\left( x,x_{0},0\right) =w_{t}\left( x,x_{0},0\right) =0.  \tag{2.10}
450: \end{equation}
451: Consider the function $h\left( x,x_{0},t\right) ,$%
452: \begin{equation*}
453: h\left( x,x_{0},t\right) =\int\limits_{0}^{t}d\tau \int\limits_{0}^{\tau
454: }w_{1}\left( x,x_{0},s\right) ds.
455: \end{equation*}
456: Integrating (2.9) with respect to $t$ twice and using (2.8) and (2.10), we
457: obtain
458: \begin{equation}
459: h_{tt}=\Delta _{x}h+a\left( x\right) w_{0}\left( x,x_{0},t\right) ,\left(
460: x,t\right) \in \mathbb{R}^{3}\times \left( 0,T\right) ,  \tag{2.11}
461: \end{equation}
462: \begin{equation}
463: h\left( x,x_{0},0\right) =h_{t}\left( x,x_{0},0\right) =0,  \tag{2.12}
464: \end{equation}
465: It follows from (2.7), (2.11), (2.12) and the formula (7.13) of \S 1 of
466: Chapter 7 of [18] that the function $h\left( x,x_{0},t\right) $ is
467: \begin{equation}
468: h\left( x,x_{0},t\right) =\frac{1}{2\pi \left( t^{2}-\left| x-x_{0}\right|
469: ^{2}\right) }\int\limits_{S\left( x,x_{0},t\right) }\left| y-x_{0}\right|
470: ^{2}a\left( y\right) d\omega _{y},  \tag{2.13}
471: \end{equation}
472: where $d\omega _{y}=\sin \theta d\varphi d\theta ,\left( \varphi ,\theta
473: \right) \in \left[ 0,2\pi \right] \times \left[ 0,\pi \right] $ are angles
474: in the spherical coordinate system with the center at $\left\{ x_{0}\right\}
475: $ and $S\left( x,x_{0},t\right) $ is the following ellipsoid with foci at $%
476: \left\{ x\right\} $ and $\left\{ x_{0}\right\} $%
477: \begin{equation*}
478: S\left( x,x_{0},t\right) =\left\{ y\in \mathbb{R}^{3}:\left| x-y\right|
479: +\left| x_{0}-y\right| =t\right\} .
480: \end{equation*}
481: 
482: Setting in (2.13) $x_{0}:=x$ and denoting $v\left( x,t\right) =h\left(
483: x,x,t\right) ,$ we obtain that the function $v$ is the spherical Radon
484: transform of the function $a,$%
485: \begin{equation}
486: v\left( x,t\right) =\frac{1}{4\pi }\int\limits_{\left| x-y\right|
487: =t/2}a\left( y\right) d\omega _{y}.  \tag{2.14}
488: \end{equation}
489: On the other hand, (2.14) implies that the function $\widetilde{v}\left(
490: x,t\right) =v\left( x,2t\right) \cdot t$ is the solution of the following
491: Cauchy problem
492: \begin{equation}
493: \widetilde{v}_{tt}=\Delta \widetilde{v},\left( x,t\right) \in \mathbb{R}%
494: ^{3}\times \left( 0,2T\right) .  \tag{2.15}
495: \end{equation}
496: \begin{equation}
497: \widetilde{v}\mid _{t=0}=0,\widetilde{v}_{t}\mid _{t=0}=a\left( x\right) .
498: \tag{2.16}
499: \end{equation}
500: Also, using the above linearization one can ``translate'' the data $p\left(
501: x_{0},t\right) $ in (2.4) for the Inverse Problem 3 in the following
502: function $\widetilde{p}\left( x,t\right) $%
503: \begin{equation}
504: \widetilde{v}\mid _{S_{T}}=\widetilde{p}\left( x,t\right) ,t\in \left(
505: 0,2T\right) .  \tag{2.17}
506: \end{equation}
507: Since the function $a\left( x\right) =0$ outside of the domain $\Omega ,$
508: then solving the initial boundary value problem (2.15)-(2.17) for $\left(
509: x,t\right) \in \left( \mathbb{R}^{3}\diagdown \Omega \right) \times \left(
510: 0,T\right) ,$ we obtain the normal derivative $q\left( x,t\right) ,$%
511: \begin{equation}
512: \frac{\partial \widetilde{v}}{\partial \nu }\mid _{S_{T}}=q\left( x,t\right)
513: ,t\in \left( 0,2T\right)  \tag{2.18}
514: \end{equation}
515: 
516: In conclusion, we have reduced the linearized Inverse Problem 3 to the $\psi
517: -$problem, which consists in the recovery of the function $a\left( x\right) $
518: from conditions (2.15)-(2.18). A similar derivation is valid for a similar
519: inverse problem for the telegraph equation (2.6) at $a_{0}\equiv 0,$ see
520: [18] and [21].
521: 
522: \section{The Method}
523: 
524: We consider Inverse Problem 1, because it is more general than Inverse
525: Problem 2. Denote $Mu=u_{tt}-Lu.$ To solve the Inverse Problem 1
526: numerically, consider the Tikhonov regularizing functional
527: \begin{equation*}
528: J_{\varepsilon }\left( u\right) =\left\| Mu-F\right\| _{L_{2}\left(
529: Q_{T}\right) }^{2}+\varepsilon \left\| u\right\| _{H^{2}\left( Q_{T}\right)
530: }^{2}
531: \end{equation*}
532: \begin{equation}
533: +\left\| D^{\beta }u\mid _{S_{T}}-D^{\beta }f\right\| _{L_{2}\left(
534: S_{T}\right) }^{2}+\left\| u_{\nu }\mid _{S_{T}}-g\right\| _{L_{2}\left(
535: S_{T}\right) }^{2}  \tag{3.1}
536: \end{equation}
537: \begin{equation*}
538: +\chi _{\varphi }\left\| u_{t}(x,0)-\psi \right\| _{L_{2}\left( \Omega
539: \right) }^{2}+\chi _{\psi }\left\| u(x,0)-\varphi \right\| _{H^{1}\left(
540: \Omega \right) }^{2},\forall u\in H^{2}\left( Q_{t}\right) .
541: \end{equation*}
542: Here $\varepsilon >0$ is the regularization parameter,
543: \begin{equation*}
544: u_{\nu }\mid _{S_{T}}:=\frac{\partial u}{\partial \nu }\mid _{S_{T}}
545: \end{equation*}
546: and $D^{\beta },\left| \beta \right| \leq 1$ is the operator of $\left(
547: x,t\right) $ derivatives with$,$ where $x$-derivatives are those, which are
548: taken in directions orthogonal to the normal vector. Also,
549: \begin{equation*}
550: \chi _{\psi }=\left\{
551: \begin{array}{c}
552: 1\text{ for the }\psi -\text{problem} \\
553: 0\text{ for the }\varphi -\text{problem}
554: \end{array}
555: \right\} ,\chi _{\varphi }=\left\{
556: \begin{array}{c}
557: 1\text{ for the }\varphi -\text{problem} \\
558: 0\text{ for the }\psi -\text{problem}
559: \end{array}
560: \right\} .
561: \end{equation*}
562: Hence, $\chi _{\varphi }\chi _{\psi }=0,\chi _{\varphi }+\chi _{\psi }=1.$
563: In previous works on the QRM terms in the second line of (3.1) were absent
564: because of subtracting off boundary conditions from the original function $u$%
565: . Terms in the third line of (3.1) were absent also, and they are
566: incorporated now to emphasize the knowledge of one of initial conditions.
567: 
568: To find the minimizer of $J_{\varepsilon }\left( u\right) ,$ we set the
569: Fr\'{e}chet derivative of this functional to zero and obtain for all $v\in
570: H^{2}\left( Q_{T}\right) $
571: \begin{equation*}
572: \int\limits_{Q_{T}}MuMvdxdt+\int\limits_{S_{T}}\left( D^{\beta }vD^{\beta
573: }u+vu\right) \mid _{S_{T}}dS+\int\limits_{S_{T}}\left( v_{\nu }u_{\nu
574: }\right) \mid _{S_{T}}dS
575: \end{equation*}
576: \begin{equation}
577: +\chi _{\psi }\int\limits_{\Omega }\left[ \nabla u\nabla v+uv\right] \left(
578: x,0\right) dx+\chi _{\varphi }\int\limits_{\Omega
579: }u_{t}(x,0)v_{t}(x,0)dx+\varepsilon \left[ u,v\right]  \tag{3.2}
580: \end{equation}
581: \begin{equation*}
582: =\int\limits_{Q_{T}}FMvdxdt+\int\limits_{S_{T}}\sum\limits_{\left| \beta
583: \right| \leq 1}\left( D^{\beta }v\mid _{S_{T}}\right) D^{\beta
584: }fdS++\int\limits_{S_{T}}\left( v_{\nu }\mid _{S_{T}}\right) \cdot gdS
585: \end{equation*}
586: \begin{equation*}
587: +\chi _{\psi }\int\limits_{\Omega }\left[ \nabla \varphi \nabla v\left(
588: x,0\right) +\varphi v\left( x,0\right) \right] dx+\chi _{\varphi
589: }\int\limits_{\Omega }\psi v_{t}(x,0)dx.
590: \end{equation*}
591: Riesz theorem and (3.2) imply
592: 
593: \textbf{Lemma 3.1.} \emph{For any vector function }$\left( F,f,g\right) \in
594: L_{2}\left( Q_{T}\right) \times H^{1}\left( S_{T}\right) \times L_{2}\left(
595: S_{T}\right) $\emph{\ there exists unique solution }$u_{\varepsilon }\in
596: H^{2}\left( Q_{T}\right) $\emph{\ of the problem (3.2) and}
597: \begin{equation*}
598: \left\| u_{\varepsilon }\right\| _{H^{2}\left( Q_{T}\right) }\leq \frac{C}{%
599: \sqrt{\varepsilon }}\left( \left\| F\right\| _{L_{2}\left( Q_{T}\right)
600: }+\left\| f\right\| _{H^{1}\left( S_{T}\right) }+\left\| g\right\|
601: _{L_{2}\left( S_{T}\right) }+\chi _{\psi }\left\| \varphi \right\|
602: _{H^{1}\left( \Omega \right) }+\chi _{\varphi }\left\| \psi \right\|
603: _{L_{2}\left( \Omega \right) }\right) .
604: \end{equation*}
605: 
606: Setting in (3.2) $v:=u,$ we obtain that the unique minimizer of the
607: functional $J_{\varepsilon }\left( u\right) $ satisfies the following
608: estimate
609: \begin{equation*}
610: \left\| Mu\right\| _{L_{2}\left( Q_{T}\right) }^{2}+\chi _{\psi }\left\|
611: u(x,0)\right\| _{H^{1}\left( \Omega \right) }^{2}+\chi _{\varphi }\left\|
612: u_{t}(x,0)\right\| _{L_{2}\left( \Omega \right) }^{2}
613: \end{equation*}
614: \begin{equation}
615: +\left\| u\mid _{S_{T}}\right\| _{H^{1}\left( S_{T}\right) }^{2}+\left\|
616: u_{\nu }\mid _{S_{T}}\right\| _{L_{2}\left( S_{T}\right) }^{2}  \tag{3.3}
617: \end{equation}
618: \begin{equation*}
619: \leq \left\| F\right\| _{L_{2}\left( Q_{T}\right) }^{2}+\left\| f\right\|
620: _{H^{1}\left( S_{T}\right) }^{2}+\left\| g\right\| _{L_{2}\left(
621: S_{T}\right) }^{2}+\chi _{\psi }\left\| \varphi \right\| _{H^{1}\left(
622: \Omega \right) }^{2}+\chi _{\varphi }\left\| \psi \right\| _{H^{1}\left(
623: \Omega \right) }^{2}.
624: \end{equation*}
625: To prove convergence of our method, we need to derive from (3.3) the
626: Lipschitz stability estimate for the function $u$ in the $H^{1}\left(
627: Q_{T}\right) $-norm. This in turn requires a modification of the proofs of
628: [4], [10], [11] and [13]. We specifically refer to the proofs of Theorem
629: 2.4.1 in [13] and Theorem 4.4 in [4]. The main difference with previous
630: proofs is that now either $\left\| u(x,0)\right\| _{H^{1}\left( \Omega
631: \right) }^{2}$ or $\left\| u_{t}(x,0)\right\| _{L_{2}\left( \Omega \right)
632: }^{2}$ can be estimated via the right hand side of (3.3), which was not done
633: before. This is because terms in the third line of (3.1) were not included
634: in the Tikhonov functional for QRM. So, if $\chi _{\psi }=0,$ then $\chi
635: _{\varphi }=1$ and we estimate $\left\| u(x,0)\right\| _{L_{2}\left( \Omega
636: \right) }^{2}$ in the $\varphi -$problem. If, however, $\chi _{\psi }=1,$
637: then $\chi _{\varphi }=0$ and we estimate $\left\| u_{t}(x,0)\right\|
638: _{L_{2}\left( \Omega \right) }^{2}$ in the $\psi $ problem. It is because of
639: the incorporation of these terms that we can assume that $T>R$, as it is
640: required in Inverse Problem 2 (see (1.9)), instead of $T>2R$ of previous
641: works. The required modification is done in the next section.
642: 
643: \section{Lipschitz Stability Estimate}
644: 
645: \textbf{Theorem 4.1.} \emph{Let }$\Omega \subset \mathbb{R}^{n}$\emph{\ be a
646: convex bounded domain with the piecewise smooth boundary\ and let }$T>R$%
647: \emph{. Suppose that the function }$u\in H^{2}\left( Q_{T}\right) $\emph{\
648: satisfies the inequality}
649: \begin{equation}
650: \left\| Mu\right\| _{L_{2}\left( Q_{T}\right) }+\chi _{\psi }\left\|
651: u(x,0)\right\| _{H^{1}\left( \Omega \right) }+\chi _{\varphi }\left\|
652: u_{t}(x,0)\right\| _{L_{2}\left( \Omega \right) }  \tag{4.1}
653: \end{equation}
654: \begin{equation*}
655: +\left\| u\mid _{S_{T}}\right\| _{H^{1}\left( S_{T}\right) }+\left\| u_{\nu
656: }\mid _{S_{T}}\right\| _{L_{2}\left( S_{T}\right) }\leq K,
657: \end{equation*}
658: \emph{where }$K=const.>0$\emph{. Then }
659: \begin{equation}
660: \left\| u\right\| _{H^{1}\left( Q_{T}\right) }+\chi _{\varphi }\left\|
661: u\left( x,0\right) \right\| _{H^{1}\left( \Omega \right) }+\chi _{\psi
662: }\left\| u_{t}\left( x,0\right) \right\| _{L_{2}\left( \Omega \right) }\leq
663: CK.  \tag{4.2}
664: \end{equation}
665: 
666: \textbf{Proof}. Choose a pair of points $x^{\prime }$,$y^{\prime }\in
667: \partial \Omega $ such that $\left| x^{\prime }-y^{\prime }\right| =2R.$ Put
668: the origin at the point $\left( x+y\right) /2.$ Choose a constant $\eta \in
669: \left( 0,1\right) $. Consider the function $p\left( x,t\right) ,$
670: \begin{equation*}
671: p\left( x,t\right) =\left| x\right| ^{2}-\eta t^{2}.
672: \end{equation*}
673: Consider the Carleman Weight Function (CWF) $\mathcal{C}(x,t),$%
674: \begin{equation*}
675: \mathcal{C}(x,t)=\exp \left( 2\lambda p\left( x,t\right) \right) ,
676: \end{equation*}
677: where $\lambda >1$ is a parameter. For any positive number $b$ denote
678: 
679: $G_{b}=\left\{ \left( x,t\right) \mid p\left( x,t\right) >b,x\in \Omega
680: ,t>0\right\} .$ Choose a sufficiently small number $c\in \left(
681: 0,R^{2}\right) .$ Since $T>R$, then in $G_{c}$%
682: \begin{equation*}
683: t^{2}<\frac{R^{2}-c}{\eta }<T^{2},\forall \eta \in \left( \eta _{0},1\right)
684: ,
685: \end{equation*}
686: where $\eta _{0}=\eta _{0}\left( T,R\right) \in \left( 0,1\right) .$ Hence, $%
687: G_{c}\subset Q_{T}.$ \ Choose $\delta \in \left( 0,c\right) $ so small that $%
688: G_{c+4\delta }\neq \varnothing .$ Note that
689: \begin{equation}
690: G_{c+4\delta }\subset G_{c+3\delta }\subset G_{c+2\delta }\subset
691: G_{c+\delta }\subset G_{c}.  \tag{4.3}
692: \end{equation}
693: 
694: Denote $M_{0}u=u_{tt}-\Delta u.$ The following pointwise Carleman estimate
695: takes place
696: \begin{equation}
697: \left( M_{0}u\right) ^{2}\mathcal{C}^{2}\geq C\lambda \left( \left| \nabla
698: _{x,t}u\right| ^{2}+\lambda ^{2}u^{2}\right) \mathcal{C}+\nabla \cdot U+V_{t}%
699: \text{ in }G_{c},\forall u\in C^{2}\left( \overline{G}_{c}\right) ,\forall
700: \lambda >\lambda _{0},  \tag{4.4}
701: \end{equation}
702: where
703: \begin{equation}
704: \left| U\right| +\left| V\right| \leq C\lambda \left( \left| \nabla
705: _{x,t}u\right| ^{2}+\lambda ^{2}u^{2}\right) \mathcal{C}\text{ in }G_{c}
706: \tag{4.5}
707: \end{equation}
708: and $\lambda _{0}\left( G_{c},\eta \right) >1$ is sufficiently large. In
709: addition, the function $V$ is estimated as
710: \begin{equation}
711: \left| V\right| \leq C\lambda ^{3}t\left( \left| \nabla _{x,t}u\right|
712: ^{2}+u^{2}\right) \mathcal{C}+C\lambda ^{3}\left| u_{t}\right| \left( \left|
713: \nabla u\right| +\left| u\right| \right) \mathcal{C}\text{ in }G_{c}.
714: \tag{4.6}
715: \end{equation}
716: This Carleman estimate was proven in Theorem 2.2.4 of [13]. It was derived
717: earlier in \S 4 of Chapter 4 of [18].
718: 
719: Consider the cut-off function $\rho \left( x,t\right) \in C^{2}\left(
720: \overline{G}_{c}\right) $ such that
721: \begin{equation}
722: \rho \left( x,t\right) =\left\{
723: \begin{array}{c}
724: 1\text{ in }G_{c+2\delta } \\
725: 0\text{ in }G_{c}\diagdown G_{c+\delta } \\
726: \text{ between 0 and 1 otherwise}
727: \end{array}
728: \right\} .  \tag{4.7}
729: \end{equation}
730: The existence of such functions is well known. For an arbitrary function $%
731: u\in C^{2}\left( \overline{G}_{c}\right) $ denote $v=v(u):=\rho u.$ Using
732: (4.4)-(4.6), we obtain
733: \begin{equation*}
734: \int\limits_{G_{c}}\left( M_{0}v\right) ^{2}\mathcal{C}dxdt\geq C\lambda
735: \int\limits_{G_{c}}\left( \left| \nabla _{x,t}v\right| ^{2}+\lambda
736: ^{2}v^{2}\right) \mathcal{C}dxdt
737: \end{equation*}
738: \begin{equation*}
739: -C\lambda ^{3}\int\limits_{\Omega }\left[ \left| u_{t}\right| \left( \left|
740: \nabla u\right| +\left| u\right| \right) \right] \left( x,0\right) \exp
741: \left( 2\lambda \left| x\right| ^{2}\right) dx-C\lambda \int\limits_{S_{T}}%
742: \left[ \left( D^{\beta }u\right) ^{2}+\lambda ^{2}u_{\nu }^{2}\right]
743: \mathcal{C}dS.
744: \end{equation*}
745: Because by (4.3) $G_{c+2\delta }\subset G_{c},$ then (4.7) implies that the
746: last inequality can be rewritten as
747: \begin{equation}
748: \int\limits_{G_{c}}\left( M_{0}v\right) ^{2}\mathcal{C}dxdt\geq C\lambda
749: \int\limits_{G_{c+2\delta }}\left( \left| \nabla _{x,t}u\right| ^{2}+\lambda
750: ^{2}u^{2}\right) \mathcal{C}dxdt  \tag{4.8}
751: \end{equation}
752: \begin{equation*}
753: -C\lambda ^{3}\int\limits_{\Omega }\left[ \left| u_{t}\right| \left( \left|
754: \nabla u\right| +\left| u\right| \right) \right] \left( x,0\right) \exp
755: \left( 2\lambda \left| x\right| ^{2}\right) dx-C\lambda \int\limits_{S_{T}}%
756: \left[ \left( D^{\beta }u\right) ^{2}+u_{\nu }^{2}\right] \mathcal{C}dS.
757: \end{equation*}
758: By (4.7) the left hand side of (4.8) can be estimated from the above as
759: \begin{equation*}
760: \int\limits_{G_{c}}\left( M_{0}v\right) ^{2}\mathcal{C}dxdt\leq
761: \int\limits_{G_{c+2\delta }}\left( M_{0}u\right) ^{2}\mathcal{C}%
762: dxdt+C\int\limits_{G_{c}\diagdown G_{c+2\delta }}\left( \left| \nabla
763: _{x,t}u\right| ^{2}+u^{2}\right) \mathcal{C}dxdt
764: \end{equation*}
765: \begin{equation*}
766: \leq \int\limits_{G_{c+2\delta }}\left( M_{0}u\right) ^{2}\mathcal{C}%
767: dxdt+C\lambda ^{3}\left\| u\right\| _{H^{1}\left( Q_{T}\right) }^{2}\exp
768: \left[ 2\lambda \left( c+2\delta \right) \right]
769: \end{equation*}
770: \begin{equation*}
771: \leq \int\limits_{G_{c+2\delta }}\left( Mu\right) ^{2}\mathcal{C}%
772: dxdt+C\int\limits_{G_{c+2\delta }}\left( \left| \nabla _{x,t}u\right|
773: ^{2}+u^{2}\right) \mathcal{C}dxdt+C\left\| u\right\| _{H^{1}\left(
774: Q_{T}\right) }\exp \left[ 2\lambda \left( c+2\delta \right) \right] .
775: \end{equation*}
776: Substituting this in (4.8), recalling that $\lambda $ is sufficiently large
777: and using (4.3), we obtain
778: \begin{equation*}
779: \int\limits_{G_{c+2\delta }}\left( Mu\right) ^{2}\mathcal{C}dxdt+\lambda
780: ^{3}\left\| u\right\| _{H^{1}\left( Q_{T}\right) }^{2}\exp \left[ 2\lambda
781: \left( c+2\delta \right) \right]
782: \end{equation*}
783: \begin{equation*}
784: +\lambda \int\limits_{S_{T}}\left[ \left( D^{\beta }u\right) ^{2}+\lambda
785: ^{2}u_{\nu }^{2}\right] \mathcal{C}dS+\lambda ^{3}\int\limits_{\Omega }\left[
786: \left| u_{t}\right| \left( \left| \nabla u\right| +\left| u\right| \right)
787: \right] \left( x,0\right) \exp \left( 2\lambda \left| x\right| ^{2}\right) dx
788: \end{equation*}
789: \begin{equation*}
790: \geq C\lambda \int\limits_{G_{c+2\delta }}\left( \left| \nabla
791: _{x,t}u\right| ^{2}+\lambda ^{2}u^{2}\right) \mathcal{C}dxdt\geq C\lambda
792: ^{3}\exp \left[ 2\lambda \left( c+3\delta \right) \right] \int%
793: \limits_{G_{c+3\delta }}\left( \left| \nabla _{x,t}u\right|
794: ^{2}+u^{2}\right) dxdt.
795: \end{equation*}
796: Let $m=\max_{\overline{G}_{c}}p(x,t).$ Dividing the last inequality by $%
797: C\lambda ^{3}\exp \left[ 2\lambda \left( c+3\delta \right) \right] ,$ we
798: obtain
799: \begin{equation*}
800: \int\limits_{G_{c+3\delta }}\left( \left| \nabla _{x,t}u\right|
801: ^{2}+u^{2}\right) dxdt\leq
802: \end{equation*}
803: \begin{equation}
804: Ce^{2\lambda m}\left( \left\| Mu\right\| _{L_{2}\left( Q_{T}\right)
805: }^{2}+\left\| u\mid _{S_{T}}\right\| _{H^{1}\left( S_{T}\right)
806: }^{2}+\left\| u_{\nu }\mid _{S_{T}}\right\| _{L_{2}\left( S_{T}\right)
807: }^{2}\right) +C\left\| u\right\| _{H^{1}\left( Q_{T}\right)
808: }^{2}e^{-2\lambda \delta }  \tag{4.9}
809: \end{equation}
810: \begin{equation*}
811: +Ce^{2\lambda m}\int\limits_{\Omega }\left[ \left| u_{t}\right| \left(
812: \left| \nabla u\right| +\left| u\right| \right) \right] \left( x,0\right) dx.
813: \end{equation*}
814: 
815: The last term of (4.9) was not present in previous publications, and we will
816: analyze it now. Consider the $\varphi -$problem first. That is, consider the
817: case $\chi _{\varphi }=1,\chi _{\psi }=0.$ Let $\gamma >0$ be a small number
818: which we will choose later. We estimate the last term of (4.9) as
819: \begin{equation*}
820: Ce^{2\lambda m}\int\limits_{\Omega }\left[ \left| u_{t}\right| \left( \left|
821: \nabla u\right| +\left| u\right| \right) \right] \left( x,0\right) dx
822: \end{equation*}
823: \begin{equation}
824: \leq C\gamma \int\limits_{\Omega }\left( \left| \nabla u\right|
825: ^{2}+u^{2}\right) \left( x,0\right) dx+\frac{Ce^{4\lambda m}}{\gamma }%
826: \int\limits_{\Omega }u_{t}^{2}\left( x,0\right) dx  \tag{4.10$\varphi $}
827: \end{equation}
828: \begin{equation*}
829: \leq C\gamma \left\| u\left( x,0\right) \right\| _{H^{1}\left( \Omega
830: \right) }^{2}+\frac{Ce^{4\lambda m}}{\gamma }K^{2}.
831: \end{equation*}
832: We have used (4.1) to estimate the last term in the second line of (4.10$%
833: \varphi $). Consider now the $\psi -$problem. Then similarly with (4.10$%
834: \varphi $)
835: \begin{equation}
836: Ce^{2\lambda m}\int\limits_{\Omega }\left[ \left| u_{t}\right| \left( \left|
837: \nabla u\right| +\left| u\right| \right) \right] \left( x,0\right) dx\leq
838: C\gamma \left\| u_{t}\left( x,0\right) \right\| _{L_{2}\left( \Omega \right)
839: }^{2}+\frac{Ce^{4\lambda m}}{\gamma }K^{2}.  \tag{4.10$\psi $}
840: \end{equation}
841: Consider now the set
842: \begin{equation*}
843: F_{1}\left( c,\delta \right) =G_{c+3\delta }\cap \left\{ t\in \left(
844: 0,\delta \right) \right\} .
845: \end{equation*}
846: Then
847: \begin{equation}
848: \left\{ \left( x,t\right) :\left| x\right| >\sqrt{c+3\delta +\eta \delta ^{2}%
849: },x\in \Omega ,t\in \left( 0,\delta \right) \right\} \subset F_{1}\left(
850: c,\delta \right) .  \tag{4.11}
851: \end{equation}
852: Then (4.9), (4.10$\varphi )$ and (4.10$\psi )$ imply that
853: \begin{equation*}
854: \int\limits_{F_{1}\left( c,\delta \right) }\left( \left| \nabla
855: _{x,t}u\right| ^{2}+u^{2}\right) dxdt\leq \frac{Ce^{4\lambda m}}{\gamma }%
856: K^{2}+C\left\| u\right\| _{H^{1}\left( Q_{T}\right) }^{2}e^{-2\lambda \delta
857: }
858: \end{equation*}
859: \begin{equation*}
860: +Ce^{2\lambda m}\left( \left\| u\mid _{S_{T}}\right\| _{H^{1}\left(
861: S_{T}\right) }^{2}+\left\| u_{\nu }\mid _{S_{T}}\right\| _{L_{2}\left(
862: S_{T}\right) }^{2}\right)
863: \end{equation*}
864: \begin{equation*}
865: +\chi _{\varphi }C\gamma \left\| u\left( x,0\right) \right\| _{H^{1}\left(
866: \Omega \right) }^{2}+\chi _{\psi }C\gamma \left\| u_{t}\left( x,0\right)
867: \right\| _{L_{2}\left( \Omega \right) }^{2}.
868: \end{equation*}
869: Hence, by (4.1)
870: \begin{equation}
871: \int\limits_{F_{1}\left( c,\delta \right) }\left( \left| \nabla
872: _{x,t}u\right| ^{2}+u^{2}\right) dxdt\leq \frac{Ce^{4\lambda m}}{\gamma }%
873: K^{2}+C\left\| u\right\| _{H^{1}\left( Q_{T}\right) }^{2}e^{-2\lambda \delta
874: }  \tag{4.12}
875: \end{equation}
876: \begin{equation*}
877: +\chi _{\varphi }C\gamma \left\| u\left( x,0\right) \right\| _{H^{1}\left(
878: \Omega \right) }^{2}+\chi _{\psi }C\gamma \left\| u_{t}\left( x,0\right)
879: \right\| _{L_{2}\left( \Omega \right) }^{2}.
880: \end{equation*}
881: 
882: Choose numbers $c$ and $\delta $ so small that $3\sqrt{c+3\delta +\eta
883: \delta ^{2}}<R.$ Hence, we can choose $x_{0}\in \Omega $ such that $\left|
884: x_{0}\right| =3\sqrt{c+3\delta +\eta \delta ^{2}}.$\ Next, we ``shift'' the
885: function $p\left( x,t\right) $ to the point $x_{0},$ thus considering the
886: function
887: \begin{equation*}
888: p\left( x-x_{0},t\right) =\left| x-x_{0}\right| ^{2}-\eta t^{2}.
889: \end{equation*}
890: For $b>0$ let $G_{b}\left( x_{0}\right) =\left\{ \left( x,t\right) \mid
891: p\left( x-x_{0},t\right) >b,x\in \Omega ,t>0\right\} .$ Similarly with the
892: above denote
893: \begin{equation*}
894: F_{2}\left( c,\delta \right) =G_{c+3\delta }\left( x_{0}\right) \cap \left\{
895: t\in \left( 0,\delta \right) \right\} .
896: \end{equation*}
897: Then
898: \begin{equation}
899: \left\{ \left( x,t\right) :\left| x-x_{0}\right| >\sqrt{c+3\delta +\eta
900: \delta ^{2}},x\in \Omega ,t\in \left( 0,\delta \right) \right\} \subset
901: F_{2}\left( c,\delta \right) .  \tag{4.13}
902: \end{equation}
903: Consider an arbitrary point $x$ such that $x\in \left\{ \left| x\right| <2%
904: \sqrt{c+3\delta +\eta \delta ^{2}}\right\} .$ Then
905: \begin{equation*}
906: \left| x_{0}-x\right| \geq \left| x_{0}\right| -\left| x\right| \geq 3\sqrt{%
907: c+3\delta +\eta \delta ^{2}}-2\sqrt{c+3\delta +\eta \delta ^{2}}=\sqrt{%
908: c+3\delta +\eta \delta ^{2}}.
909: \end{equation*}
910: Hence, by (4.13)
911: \begin{equation*}
912: \left\{ \left( x,t\right) :\left| x\right| <2\sqrt{c+3\delta +\eta \delta
913: ^{2}},t\in \left( 0,\delta \right) \right\} \subset F_{2}\left( c,\delta
914: \right) .
915: \end{equation*}
916: Combining this with (4.11), we see that
917: \begin{equation}
918: \Omega \times \left( 0,\delta \right) =Q_{\delta }\subset F_{1}\left(
919: c,\delta \right) \cup F_{2}\left( c,\delta \right) .  \tag{4.14}
920: \end{equation}
921: Using function $p\left( x-x_{0},t\right) $ instead of $p\left( x,t\right) ,$
922: we obtain similarly with (4.12)
923: \begin{equation*}
924: \int\limits_{F_{2}\left( c,\delta \right) }\left( \left| \nabla
925: _{x,t}u\right| ^{2}+u^{2}\right) dxdt\leq \frac{Ce^{4\lambda m}}{\gamma }%
926: K^{2}+C\left\| u\right\| _{H^{1}\left( Q_{T}\right) }^{2}e^{-2\lambda \delta
927: }
928: \end{equation*}
929: \begin{equation*}
930: +\chi _{\varphi }C\gamma \left\| u\left( x,0\right) \right\| _{H^{1}\left(
931: \Omega \right) }^{2}+\chi _{\psi }C\gamma \left\| u_{t}\left( x,0\right)
932: \right\| _{L_{2}\left( \Omega \right) }^{2}.
933: \end{equation*}
934: Combining this with (4.12) and (4.14), we obtain
935: \begin{equation*}
936: \left\| u\right\| _{H^{1}\left( Q_{\delta }\right) }\leq \frac{Ce^{4\lambda
937: m}}{\gamma }K^{2}+C\left\| u\right\| _{H^{1}\left( Q_{T}\right)
938: }^{2}e^{-2\lambda \delta }
939: \end{equation*}
940: \begin{equation*}
941: +\chi _{\varphi }C\gamma \left\| u\left( x,0\right) \right\| _{H^{1}\left(
942: \Omega \right) }^{2}+\chi _{\psi }C\gamma \left\| u_{t}\left( x,0\right)
943: \right\| _{L_{2}\left( \Omega \right) }^{2}.
944: \end{equation*}
945: Hence, there exists a number $t^{\ast }\in \left( 0,\delta \right) $ such
946: that
947: \begin{equation*}
948: \int\limits_{\Omega }\left( \left| \nabla _{x,t}u\right| ^{2}+u^{2}\right)
949: \left( x,t^{\ast }\right) dx\leq \frac{Ce^{4\lambda m}}{\delta \gamma }K^{2}+%
950: \frac{C}{\delta }\left\| u\right\| _{H^{1}\left( Q_{T}\right)
951: }^{2}e^{-2\lambda \delta }
952: \end{equation*}
953: 
954: \begin{equation*}
955: +\frac{1}{\delta }\left[ \chi _{\varphi }C\gamma \left\| u\left( x,0\right)
956: \right\| _{H^{1}\left( \Omega \right) }^{2}+\chi _{\psi }C\gamma \left\|
957: u_{t}\left( x,0\right) \right\| _{L_{2}\left( \Omega \right) }^{2}\right] .
958: \end{equation*}
959: This inequality combined with (4.1) and the standard energy estimates
960: implies that
961: \begin{equation*}
962: \left\| u\right\| _{H^{1}\left( Q_{T}\right) }^{2}+\chi _{\varphi }\left\|
963: u\left( x,0\right) \right\| _{H^{1}\left( \Omega \right) }^{2}+\chi _{\psi
964: }\left\| u_{t}\left( x,0\right) \right\| _{L_{2}\left( \Omega \right)
965: }^{2}\leq C\left\| u\right\| _{H^{1}\left( Q_{T}\right) }^{2}e^{-2\lambda
966: \delta }
967: \end{equation*}
968: \begin{equation}
969: +\frac{Ce^{4\lambda m}}{\gamma }K^{2}+\chi _{\varphi }C\gamma \left\|
970: u\left( x,0\right) \right\| _{H^{1}\left( \Omega \right) }^{2}+\chi _{\psi
971: }C\gamma \left\| u_{t}\left( x,0\right) \right\| _{L_{2}\left( \Omega
972: \right) }^{2}.  \tag{4.15}
973: \end{equation}
974: Note that $\delta $ is independent on $\lambda .$ Choose sufficiently large $%
975: \lambda _{0}$ such that
976: \begin{equation*}
977: 1-Ce^{-2\lambda _{0}\delta }>\frac{1}{2}
978: \end{equation*}
979: and set $\lambda :=\lambda _{0}.$ Choose $\gamma $ so small that $C\gamma
980: <1/2.$ Then we obtain (4.2) from (4.15). $\square $
981: 
982: \section{Convergence}
983: 
984: Theorem 4.1 enables us to prove convergence of our method. Following the
985: Tikhonov concept for ill-posed problems [18], we first introduce an
986: ``ideal'' exact solution of either $\varphi $ or $\psi $ problem without an
987: error in the data. Next, we assume the existence of the error in \ the
988: boundary data $f$ and $g$ and prove that our solution tends to the exact one
989: as the level of error in the data tends to zero. We consider the more
990: general Inverse Problem 1. Let $f^{\ast }\in H^{1}\left( S_{T}\right) $ and $%
991: g^{\ast }\in L_{2}\left( S_{T}\right) $ be the exact boundary data (1.3), $%
992: F^{\ast }\in L_{2}\left( Q_{T}\right) $ be the exact right hand side of
993: equation (1.1) and $\varphi ^{\ast }$ and $\psi ^{\ast }$ be exact initial
994: conditions. We assume that there exists an exact function $u^{\ast }\in
995: H^{2}\left( Q_{T}\right) $ satisfying
996: \begin{equation}
997: u_{tt}^{\ast }=L(x,t)u^{\ast }+F^{\ast }\left( x,t\right) \text{ in }Q_{T},
998: \tag{5.1}
999: \end{equation}
1000: with initial conditions
1001: \begin{equation}
1002: u^{\ast }\left( x,0\right) =\varphi ^{\ast }\left( x\right) ,u_{t}^{\ast
1003: }\left( x,0\right) =\psi ^{\ast }\left( x\right) ,\varphi ^{\ast }\in
1004: H^{1}\left( \Omega \right) ,\psi ^{\ast }\in L_{2}\left( \Omega \right) ,
1005: \tag{5.2}
1006: \end{equation}
1007: \begin{equation}
1008: u^{\ast }\mid _{S_{T}}=f^{\ast }\left( x,t\right) ,\frac{\partial u^{\ast }}{%
1009: \partial \nu }\mid _{S_{T}}=g^{\ast }\left( x,t\right) ,  \tag{5.3}
1010: \end{equation}
1011: where $\varphi ^{\ast }$ and $\psi ^{\ast }$ are exact initial conditions.
1012: We assume that the real boundary data in (1.3) have an error, so as the
1013: given initial condition. In other words, we assume that
1014: \begin{equation}
1015: \left\| f-f^{\ast }\right\| _{H^{1}\left( S_{T}\right) }+\left\| g-g^{\ast
1016: }\right\| _{L_{2}\left( S_{T}\right) }+\left\| F-F^{\ast }\right\|
1017: _{L_{2}\left( Q_{T}\right) }  \tag{5.4}
1018: \end{equation}
1019: \begin{equation*}
1020: +\chi _{\psi }\left\| \varphi -\varphi ^{\ast }\right\| _{H^{1}\left( \Omega
1021: \right) }+\chi _{\varphi }\left\| \psi -\psi ^{\ast }\right\| _{L_{2}\left(
1022: \Omega \right) }\leq \delta ,
1023: \end{equation*}
1024: where $\delta >0$ is a small number. The following convergence theorem holds
1025: 
1026: \textbf{Theorem 5.1.} \emph{Suppose that }$T>R.$\emph{\ Let }$u_{\varepsilon
1027: \delta }\in H^{2}\left( Q_{T}\right) $\emph{\ be the solution of the QRM
1028: problem (3.2), which is guaranteed by Lemma 2.1. Let conditions (5.1)-(5.4)
1029: be satisfied. Then the following estimate is valid}
1030: \begin{equation*}
1031: \left\| u-u^{\ast }\right\| _{H^{1}\left( Q_{T}\right) }+\chi _{\varphi
1032: }\left\| \varphi -\varphi ^{\ast }\right\| _{H^{1}\left( \Omega \right)
1033: }+\chi _{\psi }\left\| \psi -\psi ^{\ast }\right\| _{L_{2}\left( \Omega
1034: \right) }\leq C\left( \delta +\sqrt{\varepsilon }\right) .
1035: \end{equation*}
1036: 
1037: \textbf{Proof. }Since the functional $J_{0}\left( u\right) $ with the exact
1038: data (5.2), (5.3) achieves its minimal zero value at $u:=u^{\ast },$ then
1039: the function $u^{\ast }$ satisfies equation (3.2) with $\varepsilon =0$ and
1040: with the exact data (5.2), (5.3). Subtracting that equation for $u^{\ast }$
1041: from equation \ (3.2) for $u:=u_{\varepsilon \delta },$ denoting $%
1042: w=u_{\varepsilon \delta }-u^{\ast }$, setting in resulting equation $v:=w$
1043: and using (5.4), we obtain similarly with (3.3)
1044: \begin{equation*}
1045: \int\limits_{Q_{T}}\left( Mw\right) ^{2}dxdt+\chi _{\psi }\left\|
1046: w(x,0)\right\| _{H^{1}\left( \Omega \right) }^{2}+\chi _{\varphi }\left\|
1047: w_{t}(x,0)\right\| _{L_{2}\left( \Omega \right) }^{2}
1048: \end{equation*}
1049: \begin{equation*}
1050: +\left\| w\mid _{S_{T}}\right\| _{H^{1}\left( S_{T}\right) }^{2}+\left\|
1051: w_{\nu }\mid _{S_{T}}\right\| _{L_{2}\left( S_{T}\right) }^{2}\leq 4\delta
1052: ^{2}+\varepsilon .
1053: \end{equation*}
1054: The \ rest of the proof follows immediately from Theorem 4.1. $\square $
1055: 
1056: \section{Numerical Implementation}
1057: 
1058: In our numerical study we have considered the Inverse Problem 2. To generate
1059: the data for the inverse problem, we have solved the Cauchy problem
1060: \begin{equation}
1061: u_{tt}=\Delta u,\left( x,t\right) \in \mathbb{R}^{2}\times \left( 0,T\right)
1062: ,  \tag{6.1}
1063: \end{equation}
1064: \begin{equation}
1065: u(x,0)=\varphi \left( x\right) ,u_{t}(x,0)=\psi \left( x\right) .  \tag{6.2}
1066: \end{equation}
1067: In our numerical experiments $\psi \left( x\right) \equiv 0$ for the $%
1068: \varphi -$problem, and $\varphi \left( x\right) \equiv 0$ for the $\psi -$%
1069: problem. Because of (1.5) and the finite speed of propagation, we use in our
1070: solution of the forward problem zero Dirichlet boundary condition at the
1071: boundary of the rectangle $\left( -T,a+T\right) \times \left( -T,a+T\right) $
1072: (Figure 1). Hence, we solve initial boundary value problem inside of this
1073: rectangle for equation (6.1) with initial conditions (6.2) at zero Dirichlet
1074: boundary condition. In all our calculations we took $a=1.$ In tests 1, 2 and
1075: 5, which are concerned with the Inverse Problem 2, we took $T=3.$ Hence,
1076: condition (1.9) is satisfied. Tests 3 and 4 are concerned with the Inverse
1077: Problem 1 and we have taken different values of $T$ in these tests. The
1078: square $SQ(a)$ is $SQ(a)=SQ(1)=\left( 0,1\right) \times \left( 0,1\right) ,$
1079: the domain $\Omega $ in tests 1,2 and 5 is
1080: \begin{equation}
1081: \Omega :=\left( 0,4\right) \times \left( 0,4\right)  \tag{6.3}
1082: \end{equation}
1083: and in all tests
1084: \begin{equation}
1085: \varphi \left( x\right) =\psi \left( x\right) =0\text{ for }x\notin SQ(1).
1086: \tag{6.4}
1087: \end{equation}
1088: 
1089: We have solved the Cauchy problem (6.1), (6.2) via finite differences using
1090: the uniform grid. We set
1091: \begin{equation*}
1092: u(t_{k},x_{1n},x_{2m})\approx
1093: u_{kmn},\,k=0,...,N_{t},\,\,n=0,...,N_{x},m=0,...,N_{y},
1094: \end{equation*}
1095: \begin{equation*}
1096: t_{k}=kh_{t},\,x_{1n}=nh_{x_{1}},\,x_{2m}=mh_{x_{2}},
1097: \end{equation*}
1098: step sizes $h_{x_{1}}=h_{x_{2}}=0.1,h_{t}=1/15$ and $N_{x}=N_{y}=10,N_{t}=45$%
1099: . This solution has generated the boundary data (1.8). Next, we have
1100: introduced noise in these data as
1101: \begin{equation}
1102: f_{n}\left( x^{i},t_{j}\right) =f\left( x^{i},t_{j}\right) \left( 1+\gamma
1103: N\left( t_{j}\right) \right) ,g_{n}\left( x^{i},t_{j}\right) =g\left(
1104: x^{i},t_{j}\right) \left( 1+\gamma N\left( t_{j}\right) \right) ,  \tag{6.5}
1105: \end{equation}
1106: where $\left( x^{i},t_{j}\right) $ is the grid point at the boundary. Here $%
1107: N\in \left( -1,1\right) $ is a pseudo random variable, which is given by
1108: function $Math.random()$ in Java and $\gamma \in \left[ 0.05,0.5\right] $ is
1109: the noise level. We have chosen the grid points the same as ones in the
1110: finite difference scheme we have solved the problem (6.1), (6.2). The
1111: presence of the random noise in the date prevents us from committing
1112: ``inverse crime''. In (6.5) points $x^{i}\in \Gamma _{1T}\cup \Gamma _{2T}.$
1113: As to $\Gamma _{3T}\cup \Gamma _{4T},$ we simply set $f=g=0$ on this part of
1114: the boundary, because of (1.7).
1115: 
1116: To find the minimizer of the functional $J_{\varepsilon },$ we have also
1117: used finite differences. We have used in (3.1) the finite difference
1118: approximations for $Mu=u_{tt}-\Delta u$ and $u_{\nu }\mid _{S_{T}}$ and have
1119: minimized the resulting functional $\widetilde{J}_{\varepsilon }$ with
1120: respect to the vector $\left\{ u_{kmn}\right\} ,$ which approximates values
1121: of the function $u$ at grid points. Here $\widetilde{J}_{\varepsilon }$
1122: means the functional $J_{\varepsilon },$ which\ is expressed via the finite
1123: differences. The norms $\left\| u_{x_{1}}(x,0)\right\| _{L_{2}\left( \Omega
1124: \right) },$ $\left\| u_{x_{2}}(x,0)\right\| _{L_{2}\left( \Omega \right) }$
1125: in $\left\| u(x,0)\right\| _{H^{1}\left( \Omega \right) }$ in the $\psi -$%
1126: problem were calculated via finite differences. As to the term $\left\|
1127: D^{\beta }u\mid _{S_{T}}-D^{\beta }f\right\| _{L_{2}\left( S_{T}\right)
1128: }^{2} $ in (3.1), we have used only $\beta =0,$ thus ending up with $\left\|
1129: u\mid _{S_{T}}-f\right\| _{L_{2}\left( S_{T}\right) }^{2}$ (in the discrete
1130: sense)$.$ Note that since $\beta =0,$ our numerical results seem to be
1131: stronger than Theorem 4.1 predicts. The integrals were calculated as
1132: \begin{equation*}
1133: \int\limits_{\Omega _{T}}(u_{tt}-\Delta u)^{2}dv\approx \frac{%
1134: h_{t}h_{x_{2}}h_{x_{1}}}{h_{t}^{4}}\sum_{k=1}^{N_{t}-1}\sum_{m=1}^{N_{y}-1}%
1135: \sum_{n=1}^{N_{x}-1}M_{kmn}^{2},
1136: \end{equation*}
1137: where
1138: \begin{multline*}
1139: M_{kmn}=(u_{k+1,mn}-2u_{kmn}+u_{k-1,mn})-\lambda
1140: _{y}(u_{k,m+1,n}-2u_{kmn}+u_{k,m-1,n}) \\
1141: -\lambda _{x}\left( u_{km,n+1}-2u_{kmn}+u_{km,n-1}\right) \\
1142: =(u_{k+1,mn}+u_{k-1,mn})-\lambda _{y}(u_{k,m+1,n}+u_{k,m-1,n})-\lambda
1143: _{x}(u_{km,n+1}+u_{km,n-1})-\lambda _{t}u_{kmn},
1144: \end{multline*}
1145: where
1146: \begin{equation*}
1147: \lambda _{x}=\frac{h_{t}^{2}}{h_{x_{1}}^{2}},\quad \lambda _{y}=\frac{%
1148: h_{t}^{2}}{h_{x_{2}}^{2}},\quad \lambda _{t}=2(1-\lambda _{x}-\lambda _{y}).
1149: \end{equation*}
1150: Also,
1151: \begin{equation*}
1152: \int\limits_{0}^{T}\int\limits_{0}^{a+T}(u(t,x_{2},x_{1}^{\ast
1153: })-f(t,x_{2}))^{2}dx_{2}dt\approx
1154: h_{t}h_{x_{2}}\sum_{k=0}^{N_{t}}\sum_{m=0}^{N_{x_{2}}}H_{km}^{2},
1155: \end{equation*}
1156: where
1157: \begin{equation*}
1158: H_{km}=u_{kmn_{\ast }}-h_{km},
1159: \end{equation*}
1160: where $n_{\ast }$ in the layer number (value of $x_{1}^{\ast }$) at which
1161: the grid function $f_{km}$ is given.
1162: 
1163: To minimize the functional $\widetilde{J}_{\varepsilon },$ we have used the
1164: conjugate gradient method. Derivatives with respect to variables $u_{kmn}$
1165: where calculated in closed forms, using the following formula
1166: \begin{equation*}
1167: \frac{\partial u_{kmn}}{\partial u_{\overline{k}\overline{m}\overline{n}}}%
1168: =\delta _{k\overline{k}}\delta _{m\overline{m}}\delta _{n\overline{n}},
1169: \end{equation*}
1170: where $\delta _{k\overline{k}}$ is the Kronecker symbol. This formula can be
1171: conveniently used to obtain closed form expressions for derivatives
1172: \begin{equation*}
1173: \frac{\partial \widetilde{J}_{\varepsilon }\left( u\right) }{\partial u_{kmn}%
1174: }.
1175: \end{equation*}
1176: 
1177: Let $a$ be the vector of unknowns of the functional $\widetilde{J}%
1178: _{\varepsilon }$. We start our iterative process from $a:=a_{0}=0$. It is
1179: well known in the field of ill-posed problems that the number of iterations
1180: can often be taken as a regularization parameter, and it depends, of course
1181: on the range of parameters of a problem one considers. We have found that
1182: the optimal number of iterations for our range of parameters is $300$. Thus,
1183: in all numerical examples below 300 iterations of the conjugate gradient
1184: method were used, thus ending up with $a_{300}$. Figure $\ref
1185: {fig:functl_and_grad}$ displays typical dependencies of the functional $%
1186: \widetilde{J}_{\varepsilon }\left( a_{k}\right) $ and the norm of its
1187: gradient on the iteration number $k$.
1188: 
1189: \begin{figure}[t]
1190: \begin{center}
1191: \begin{tabular}{c@{\hspace{0.05cm}}c}
1192: \includegraphics[width=8.5cm]{J.eps} & \includegraphics[width=8.5cm]{g.eps}
1193: \end{tabular}
1194: \end{center}
1195: \caption{Typical dependence of the functional $J=J_{\protect\varepsilon }$
1196: (left) and $g=\|\protect\nabla J_{\protect\varepsilon }\|^2$ (right) on
1197: number of iterations.}
1198: \label{fig:functl_and_grad}
1199: \end{figure}
1200: 
1201: \section{Numerical Results}
1202: 
1203: In this section we present results of some numerical experiments. We have
1204: always used $\varepsilon =10^{-6}.$ Larger values of $\varepsilon $ such as $%
1205: 10^{-5}$ brought lower quality results. In our numerical experiments we have
1206: imaged both smooth slowly varying functions and the finite difference
1207: analogue of the $\delta -$ function. Let $\left( x_{1k},x_{2r}\right) \in
1208: \Omega $ be a fixed grid point. To obtain the finite difference analogue of $%
1209: \delta \left( x_{1}-x_{1k},x_{2}-x_{2r}\right) $, we consider the following
1210: grid points $\left( x_{1n},x_{2m}\right) $ and we model the function $\delta
1211: \left( x_{1n}-x_{1k},x_{2m}-x_{2r}\right) $ as
1212: \begin{equation*}
1213: \delta \left( x_{1n}-x_{1k},x_{2m}-x_{2r}\right) =\frac{3}{%
1214: 4h_{x_{2}}h_{x_{1}}}\delta _{nk}\delta _{mr},
1215: \end{equation*}
1216: where the multiplier at $\delta _{nk}\delta _{mr}$ is chosen such that the
1217: volume of the pyramid based on $\left( x_{1k-1},x_{2r-1}\right) ,\,\left(
1218: x_{1k-1},x_{2r+1}\right) ,\,\left( x_{1k+1},x_{2r+1}\right) ,\,\left(
1219: x_{1k+1},x_{2r-1}\right) $ equals to $1$. Hence, the support of the function
1220: $\delta \left( x_{1n}-x_{1k},x_{2m}-x_{2r}\right) $ is limited only to the
1221: point $\left( x_{1n},x_{2m}\right) $.
1222: \begin{figure}[tbp]
1223: \begin{center}
1224: \includegraphics[width=16cm, height=8cm]{bad_bndry.eps}
1225: \end{center}
1226: \caption{Exact (red) and calculated (black) functions $\protect\varphi $ in
1227: (7.1) without balancing coefficients with 5\% noise in boundary data.}
1228: \label{fig:bad_bndry}
1229: \end{figure}
1230: \begin{figure}[tbp]
1231: \begin{center}
1232: \includegraphics[width=16cm, height=8cm]{good_bndry.eps}
1233: \end{center}
1234: \caption{Exact (red) and calculated (black) functions $\protect\varphi $ in
1235: (7.1) with balancing coefficients with 5\% noise in boundary data.}
1236: \label{fig:good_bndry}
1237: \end{figure}
1238: 
1239: We have observed that having equal coefficient at all terms of the
1240: functional $\widetilde{J}_{\varepsilon }$ in (3.1) does not lead to good
1241: reconstruction results. This is because not all the terms of (3.1) provide
1242: an equal impact in this functional. For example, for the $\varphi -$problem
1243: with no noise in the data for the function
1244: \begin{equation}
1245: \varphi \left( x\right) =\left\{
1246: \begin{array}{c}
1247: \sin (2\pi x_{1})\sin (2\pi x_{2}),x\in SQ\left( 1\right) \\
1248: 0\text{ otherwise}
1249: \end{array}
1250: \right\}  \tag{7.1}
1251: \end{equation}
1252: we first got the result displayed in Figure $\ref{fig:bad_bndry}$. One can
1253: observe that the error at the boundary is significant. And indeed, the
1254: values of two terms in (3.1) after 300 iterations were for this case
1255: \begin{equation*}
1256: \int\limits_{\Omega _{T}}(u_{tt}-\Delta u)^{2}dxdt\approx 10^{-3},\quad
1257: \Vert u-f\Vert _{L_{2}(S_{T})}\approx 10^{-2}.
1258: \end{equation*}
1259: Hence, the impact of the boundary term in (3.1) is 10 greater than the
1260: impact of the $\left\| Mu\right\| _{L_{2}\left( Q_{T}\right) }^{2}.$ To
1261: minimize the error at the boundary, we took the balancing coefficient $1000$
1262: at $1000\cdot \Vert u-f\Vert _{L_{2}(S_{T})}$ instead of $1\cdot \Vert
1263: u-f\Vert _{L_{2}(S_{T})}$. The other balancing coefficients equal to $1$.
1264: The quality of the resulting image was improved, see Figure $\ref
1265: {fig:good_bndry}$. Thus, in all our tests with the $\varphi -$problem we
1266: have taken the same balancing coefficients. In the case of the $\psi -$%
1267: problem we have taken $100\cdot \chi _{\psi }\left\| u(x,0)-\varphi \right\|
1268: _{H^{1}\left( \Omega \right) }^{2}$ and the other balancing coefficients
1269: equal to $1$.
1270: 
1271: Note that Theorems 4.1 and 5.1 remain the same, including their proofs, if
1272: balancing coefficients are introduced.
1273: 
1274: \textbf{Test 1.} \emph{The }$\varphi -$\emph{problem.} Here $\psi \left(
1275: x\right) \equiv 0$ and the function $\varphi \left( x\right) $ to be
1276: reconstructed is one in (7.1). In Figures $\ref{fig:phi_sin_25}$ and $\ref
1277: {fig:phi_sin_50}$ represent resulting images with 25\% and 50\% noise
1278: respectively. Next, we test our method for the case when the term with $\chi
1279: _{\varphi }$ is absent in the functional $J_{\varepsilon }\left( u\right) $
1280: in (3.1).\ Regardless on the small amount of noise in the data, both maximal
1281: ($1$) and minimal ($-1$) values of the imaged function were missed by about
1282: 22\% in this case, whereas they were not missed in the previous cases with
1283: 25\% and 50\% noise when the term with $\chi _{\varphi }$ was not absent in
1284: (3.1). To see this, we display on Figure $\ref{fig:phi_sin_cross}$ the
1285: 1-dimensional cross-sections by the straight line $\left\{ x_{1}=0.5\right\}
1286: $ of the correct function (7.1), the imaged function with 50\% noise of
1287: Figure $\ref{fig:phi_sin_50}$ and the imaged function with the absent term
1288: with $\chi _{\varphi }$ and 5\% noise. One can observe that the maximal
1289: value of the calculated function is $0.7$, while the maximal absolute value
1290: of the correct function is $0.9$, so as the one of Figure $\ref
1291: {fig:phi_sin_50}$. Here we have $0.9$ instead of $1$ only because the points
1292: with the absolute value of $1$ are not the grid points. \ This emphasizes
1293: the importance of the incorporation of the term with $\chi _{\varphi }.$ We
1294: have observed the same for the $\psi -$ problem (images not shown).
1295: \begin{figure}[tbp]
1296: \begin{center}
1297: \includegraphics[width=16cm, height=8cm]{phi_sin_25.eps}
1298: \end{center}
1299: \caption{Test 1. Exact (red) and calculated (black) functions $\protect%
1300: \varphi $ in (7.1) with 25\% noise in the boundary data.}
1301: \label{fig:phi_sin_25}
1302: \end{figure}
1303: \begin{figure}[tbp]
1304: \begin{center}
1305: \includegraphics[width=16cm, height=8cm]{phi_sin_50.eps}
1306: \end{center}
1307: \caption{Test 1. Exact (red) and calculated (black) functions $\protect%
1308: \varphi $ in (7.1) with 50\% noise in the boundary data.}
1309: \label{fig:phi_sin_50}
1310: \end{figure}
1311: \begin{figure}[tbp]
1312: \begin{center}
1313: \includegraphics[width=16cm, height=16 cm]{phi_sin_cross.eps}
1314: \end{center}
1315: \caption{Test 1. Cross sections of exact (red) and calculated (black, blue)
1316: functions $\protect\varphi $ with 5\%, 50\% noise, ''no integral'' means $%
1317: \protect\chi _{\protect\varphi =0}$. One can see that the maximal value of
1318: the case $\protect\chi _{\protect\varphi =0}$ is $0.7/(-0.7)$. The maximal
1319: value of the exact function is $0.9<1$ only because of the grid step size.}
1320: \label{fig:phi_sin_cross}
1321: \end{figure}
1322: 
1323: \textbf{Test 2}.\emph{\ The }$\psi -$\emph{problem.} In this case $\varphi
1324: \left( x\right) \equiv 0$ and the function $\psi \left( x\right) $ to be
1325: reconstructed is
1326: \begin{equation}
1327: \psi \left( x\right) =\left\{
1328: \begin{array}{c}
1329: \sin (\frac{\pi }{2}\left( x_{1}-0.5\right) \sin (\frac{\pi }{2}\left(
1330: x_{2}-0.5\right) ),x\in SQ\left( 1\right) \\
1331: 0\text{ otherwise.}
1332: \end{array}
1333: \right\} .  \tag{7.2}
1334: \end{equation}
1335: Figures $\ref{fig:psi_sin_5}$, $\ref{fig:psi_sin_25}$ and $\ref
1336: {fig:psi_sin_50}$display resulting images of the function (7.2) with 5\%,
1337: 25\% and 50\% of the noise level in the data respectively.
1338: \begin{figure}[tbp]
1339: \begin{center}
1340: \includegraphics[width=16cm, height=8cm]{psi_sin_5.eps}
1341: \end{center}
1342: \caption{Test 2. Exact (red) and calculated (black) functions $\protect\psi$
1343: in (7.2) with 5\% noise in the boundary data.}
1344: \label{fig:psi_sin_5}
1345: \end{figure}
1346: \begin{figure}[tbp]
1347: \begin{center}
1348: \includegraphics[width=16cm, height=8cm]{psi_sin_25.eps}
1349: \end{center}
1350: \caption{Test 2. Exact (red) and calculated (black) functions $\protect\psi$
1351: in (7.2) with 25\% noise in the boundary data.}
1352: \label{fig:psi_sin_25}
1353: \end{figure}
1354: \begin{figure}[tbp]
1355: \begin{center}
1356: \includegraphics[width=16cm, height=8cm]{psi_sin_50.eps}
1357: \end{center}
1358: \caption{Test 2. Exact (red) and calculated (black) functions $\protect\psi$
1359: with 50\% noise in the boundary data.}
1360: \label{fig:psi_sin_50}
1361: \end{figure}
1362: 
1363: \textbf{Test 3}. \emph{The }$\varphi -$\emph{problem in }$SQ(1)$\emph{\ for }%
1364: $T\in \left( 0.5diam\,SQ\left( 1\right) ,diam\,SQ\left( 1\right) \right) ,$
1365: \emph{where }$diam\,SQ\left( 1\right) =\sqrt{2}$\emph{\ is the diameter of
1366: the square }$SQ\left( 1\right) $. We have decided to see what kind of
1367: results can be obtained if the boundary Cauchy data are given on the entire
1368: boundary of the square $SQ\left( 1\right) $ in the case when $\,T\in \left(
1369: 0.5diam\,SQ\left( 1\right) ,diam\,SQ\left( 1\right) \right) .$ We are
1370: especially interested in the question about the influence of terms with $%
1371: \chi _{\varphi }$ and $\chi _{\psi }.$ We have used
1372: \begin{equation*}
1373: \,T=0.75<diam\,SQ\left( 1\right) =\sqrt{2},N_{x}=N_{y}=20,\,N_{t}=3.
1374: \end{equation*}
1375: and have reconstructed the function (7.1). Figure $\ref{fig:phi_sin_e_t_25}$
1376: displays the resulting image with $25\%$ noise in the case when the term $%
1377: \chi _{\varphi }$ is present in (3.1). This quality of the reconstruction is
1378: good for such a high noise level. Figure $\ref{fig:phi_sin_e_t_25_cross}$
1379: displays the 1-dimensional cross-section of the image by the straight line $%
1380: \left\{ x_{1}=0.5\right\} $, as well as the 1-dimensional cross-section of
1381: the image for the case when the term with $\chi _{\varphi }$ is not present
1382: in (3.1) and 25\% noise in the data is in. One can observe that the minimal
1383: value of $(-0.9)$ is not achieved in the case when the term with $\chi
1384: _{\varphi }$ is not present. The calculated minimal value is $(-0.7)$ in
1385: this case.
1386: \begin{figure}[tbp]
1387: \begin{center}
1388: \includegraphics[width=16cm]{phi_sin_25_entire_t.eps}
1389: \end{center}
1390: \caption{Test 3. Exact (red) and calculated (black) solutions of the problem
1391: $\protect\varphi -$ in SQ(1) with 25\% noise in the boundary data for $%
1392: T=0.75 $.}
1393: \label{fig:phi_sin_e_t_25}
1394: \end{figure}
1395: \begin{figure}[tbp]
1396: \begin{center}
1397: \includegraphics[width=16cm]{phi_sin_25_entire_t_cross.eps}
1398: \end{center}
1399: \caption{Test 3. Cross sections of exact (red) and calculated (black, blue)
1400: functions $\protect\varphi $ with 25\% noise in the boundary data for $%
1401: T=0.75 $, ''no integral'' means $\protect\chi _{\protect\varphi }=0$. The
1402: maximal value of the exact function is $0.9<1$ only because of the grid step
1403: size.}
1404: \label{fig:phi_sin_e_t_25_cross}
1405: \end{figure}
1406: 
1407: \textbf{Test 4}. \emph{The }$\varphi -$\emph{problem in $SQ(1)$ for }$%
1408: T>diam\,SQ\left( 1\right) $. We now test our method for the case when the
1409: boundary Cauchy data are given at the entire boundary of the rectangle $%
1410: SQ\left( 1\right) $ and $T>diam\,SQ\left( 1\right) $. We take
1411: \begin{equation*}
1412: T=2,\,N_{x}=N_{y}=20,\,N_{t}=60.
1413: \end{equation*}
1414: The function (7.1) was reconstructed. Figure $\ref{fig:phi_sin_e_t1_25}$
1415: displays the resulting image with $25\%$ noise and Figure $\ref
1416: {fig:phi_sin_e_t1_25_cross}$ displays the 1-dimensional cross-section of the
1417: image by the straight line $\left\{ x_{1}=0.5\right\} $, as well as the
1418: 1-dimensional cross-section of the image for the case when the term with $%
1419: \chi _{\varphi }$ is not present in (3.1) (with 25\% noise). One can observe
1420: that both images are very close to the correct one. This points towards the
1421: fact, which follows from the theory of above cited publications and also
1422: from Theorem 4.1: the presence of terms with $\chi _{\varphi }$ and $\chi
1423: _{\psi }$ is important only when $T\in \left( R,2R\right) $ and it is
1424: unimportant for $T>2R.$
1425: \begin{figure}[tbp]
1426: \begin{center}
1427: \includegraphics[width=16cm]{phi_sin_25_entire_T1.eps}
1428: \end{center}
1429: \caption{Test 4. Exact (red) and calculated (black) solutions of the $%
1430: \protect\varphi -$ problem in SQ(1) with 25\% noise in the boundary data for
1431: $T=2$.}
1432: \label{fig:phi_sin_e_t1_25}
1433: \end{figure}
1434: \begin{figure}[tbp]
1435: \begin{center}
1436: \includegraphics[width=16cm]{phi_sin_25_entire_T1_cross.eps}
1437: \end{center}
1438: \caption{Test 4. Cross sections of exact (red) and calculated (black, blue)
1439: functions $\protect\varphi $ with 25\% noise in the boundary data for $%
1440: T=0.75 $, ''no integral'' means $\protect\chi _{\protect\varphi }=0$. The
1441: maximal value of the exact function is $0.9<1$ only because of the grid step
1442: size.}
1443: \label{fig:phi_sin_e_t1_25_cross}
1444: \end{figure}
1445: \begin{figure}[tbp]
1446: \begin{center}
1447: \includegraphics[width=16cm]{phi_delta_50.eps}
1448: \end{center}
1449: \caption{Test 5. Exact (red) and calculated (black) function $\protect%
1450: \varphi $ with 50\% noise in the boundary data. The function $\protect\chi _{%
1451: \protect\varphi }$ in (3.1) is present. Scatter plot mode. Squares show
1452: heights. Correct heights are achieved.}
1453: \label{fig:phi_delta_50}
1454: \end{figure}
1455: \begin{figure}[tbp]
1456: \begin{center}
1457: \includegraphics[width=16cm]{phi_delta_5_nogt.eps}
1458: \end{center}
1459: \caption{Test 5. Exact (red) and calculated (black) function $\protect%
1460: \varphi $ with 5\% noise in the boundary data and $\protect\chi _{\protect%
1461: \varphi }=0$. Scatter plot mode. Squares show heights. Correct heights are
1462: not achieved.}
1463: \label{fig:phi_delta_5_nogt}
1464: \end{figure}
1465: 
1466: \textbf{Test 5.} \emph{The }$\varphi -$\emph{problem with two }$\delta -$%
1467: \emph{functions. }We now again consider the Inverse Problem 2 with the
1468: domain $\Omega $ as in (6.3) and with $\psi (x)\equiv 0.$\ The data for the
1469: forward problem were simulated for the case
1470: \begin{equation}
1471: \varphi \left( x_{1},x_{2}\right) =\delta \left( x_{1}-0.4,x_{2}-0.4\right)
1472: +\delta \left( x_{1}-0.7,x_{2}-0.7\right)  \tag{7.3}
1473: \end{equation}
1474: with the above described finite difference analogue of the $\delta -$
1475: function. Figure $\ref{fig:phi_delta_50}$ displays the resulting image of
1476: the function (7.3) for the case of 50\% of the noise in the boundary data,
1477: scatter plot mode was used, squares show exact height. Figure $\ref
1478: {fig:phi_delta_5_nogt}$, on the other hand, shows the image when the term
1479: with $\chi _{\varphi }$ is absent in (3.1) and only 5\% noise in the data is
1480: present. One can see that the correct height is not reached on Figure $\ref
1481: {fig:phi_delta_5_nogt}$, unlike Figure $\ref{fig:phi_delta_50}$. This again
1482: shows the importance of the introduction of terms in the third line of (3.1).
1483: 
1484: Very similar results (not shown) were obtained for the $\psi -$problem with
1485: exactly the same $\delta -$ functions as ones in (7.3).
1486: 
1487: \section{Conclusions}
1488: 
1489: We have considered the inverse problems of the determination of one of
1490: initial condition in a hyperbolic equation using the lateral Cauchy data.\
1491: We have presented applications of these problems to the thermoacoustic
1492: tomography, as well as to linearized inverse acoustic and inverse
1493: electromagnetic problems. The problems we consider are very close ones with
1494: the Cauchy problems for hyperbolic equations with the lateral data, and we
1495: have actually solved the latter numerically in Tests 3 and 4. We have
1496: focused on the inverse problem in an infinite domain (octant), whereas only
1497: finite domains were considered in previous numerical studies. Nevertheless,
1498: we are able to reduce our inverse problem to one in a finite domain due to
1499: the finite speed of propagation of waves. Since one initial condition is
1500: known, we were able to decrease the observation time $T$ by twofold. We have
1501: shown numerically that it is important to know one of initial conditions if $%
1502: T<diameter\left( \Omega \right) ,$ as it is required by stability and
1503: uniqueness results. However, if $T>diameter\left( \Omega \right) ,$ then
1504: both the theory and our numerical result of Test 4 show that one does not
1505: need to know the initial condition.
1506: 
1507: We have proposed a new version of the Quasi-Reversibility method. The main
1508: new element of this version is the inclusion of the terms characterizing
1509: \emph{a priori} knowledge of one of initial conditions. Two other new
1510: elements are incorporation of boundary terms in the Tikhonov functional
1511: instead of subtracting off boundary conditions and the use of finite
1512: differences instead of finite elements in the inverse solver. To prove
1513: convergence of this new version, we have modified the technique of previous
1514: works, which is based on Carleman estimates. A comprehensive numerical study
1515: of the proposed numerical method was conducted. This study has demonstrated
1516: robustness of our technique with respect up to 50\% random noise in the
1517: data, similarly with previous publications [4], [12], [15]. This study has
1518: also demonstrated that this method is capable to image sharp peaks, which is
1519: important for the application to thermoacoustic tomography, for example.
1520: 
1521: \begin{center}
1522: \textbf{Acknowledgment}
1523: \end{center}
1524: 
1525: The research of M.V. Klibanov and A. V. Kuzhuget was supported by the U.S.
1526: Army Research Laboratory and U.S. Army Research Office under contract/grant
1527: number W911NF-05-1-0378. The first author has performed a part of this work
1528: during the Special Semester on Quantative Biology Analyzed by Mathematical
1529: Methods, October 1$^{\text{st}}$, 2007 - January 27$^{\text{th}}$, 2008,
1530: organized by RICAM, Austrian Academy of Sciences.
1531: 
1532: \begin{thebibliography}{99}
1533: \bibitem{Agr}  M.\ Agranovsky and P. Kuchment, \emph{Uniqueness of
1534: reconstruction and an inversion procedure for thermoacoustic tomography with
1535: variable sound speed}, Inverse Problems, 23, pp. 2089-2102, 2007.
1536: 
1537: \bibitem{Burgeois05}  \textsc{L. Burgeois}, \textit{A mixed formulation of
1538: quasi-reversibility to solve the Cauchy problem for Laplace's equation},
1539: Inverse Problems, 21, pp.~1087--1104, 2005.
1540: 
1541: \bibitem{Burgeois06}  \textsc{L. Burgeois}, \textit{Convergence rates for
1542: the quasi-reversibility method to solve the Cauchy problem for Laplace's
1543: equation}, Inverse Problems, 22, pp.~413--430, 2006.
1544: 
1545: \bibitem{CK}  \textsc{C. Clason and M.V. Klibanov}, \emph{The
1546: quasi-reversibility method for thermoacoustic tomography in a heterogeneous
1547: medium}, SIAM J. Sci. Comp., 30, pp. 1-23, 2007.
1548: 
1549: \bibitem{FPR}  \textsc{D.\ Finch, S. Patch and Rakesh,} \emph{Determining a
1550: function from its mean values over a family of spheres}, SIAM J.\ Math.
1551: Anal., 35, pp. 1213-1240, 2004.
1552: 
1553: \bibitem{Isakov06}  \textsc{V. Isakov}, \textit{Inverse Problems for Partial
1554: Differential Equations}, Springer, New York, 2006.
1555: 
1556: \bibitem{KBN05}  \textsc{S.I. Kabanikhin, M.A. Bektemesov and D.V.\ Nechaev}%
1557: , \emph{Numerical solution of the 2D thermoacoustic problem}, J. Inverse and
1558: Ill-Posed Problems, 13, pp.~265--276, 2005.
1559: 
1560: \bibitem{KD90}  \textsc{M.V. Klibanov and P.G. Danilaev}, \textit{On the
1561: solution of coefficient inverse problems by the method of quasi-inversion},
1562: Soviet Math. Doklady, 41, pp.~83--87, 1990.
1563: 
1564: \bibitem{KS91}  \textsc{M.V. Klibanov and F.\ Santosa}, \emph{A
1565: computational quasi-reversibility method for Cauchy problems for Laplace's
1566: equation}, SIAM J. Appl. Math., 31, pp.~1653--1675, 1991.
1567: 
1568: \bibitem{KK93}  \textsc{M. Kazemi and M.V. Klibanov}, \textit{Stability
1569: estimates for ill-posed problems involving hyperboilc equations and
1570: inequalities}, Applicable Analysis, 50 (1993), pp.~93--102.
1571: 
1572: \bibitem{KM91}  \textsc{M.V. Klibanov and J. Malinsky}, \textit{%
1573: Newton-Kantorovich method for 3-dimensional inverse scattering problem and
1574: stability of the hyperbolic Cauchy problem with time dependent data},
1575: Inverse Problems, 7, pp.~577--595, 1991.
1576: 
1577: \bibitem{KR92}  \textsc{M.V. Klibanov and Rakesh}, \emph{Numerical solution
1578: of a timelike Cauchy problem for the wave equation}, Math.\ Methods in Appl.
1579: Science, 15, pp.~559--570, 1992.
1580: 
1581: \bibitem{KT04}  \textsc{M. V. Klibanov and A. Timonov}, \emph{Carleman
1582: Estimates for Coefficient Inverse Problems and Numerical Applications}, VSP,
1583: Utrecht, 2004.
1584: 
1585: \bibitem{K05}  \textsc{M. V. Klibanov}, \emph{Lipschitz stability for
1586: hyperbolic inequalities in octants with the lateral Cauchy data and
1587: refocusing in time reversal}, J. Inverse and Ill-Posed Problems, 13,
1588: pp.~353--363, 2005.
1589: 
1590: \bibitem{KKN06}  \textsc{M.V. Klibanov, S.I. Kabanikhin and D.V.\ Nechaev},
1591: \emph{Numerical solution of the problem of computational time reversal in a
1592: quadrant}, Waves in Random and Complex Media, 16, pp.~473--494, 2006.
1593: 
1594: \bibitem{KK07}  \textsc{P. Kuchment and L.\ Kunyansky,} \emph{Mathematics of
1595: thermoacoustic tomography}, preprint, available at arXiv: 0704.0286v2
1596: [math.AP] 21 Oct 2007.
1597: 
1598: \bibitem{Kunyan}  \textsc{L.\ Kunyansky, }\emph{Explicit inversion formulae
1599: for the spherical mean Radon transform, }Inverse Problems, 23, pp. 373-383,
1600: 2007.
1601: 
1602: \bibitem{LRS}  \textsc{M.M. Lavrent'ev, V.G.\ Romanov and S.P. Shishatskii},
1603: \emph{Ill-Posed Problems of Mathematicl Physics and Analysis}, AMS,
1604: Providence, RI, 1986.
1605: 
1606: \bibitem{LJ69}  \textsc{R. Lattes \ and J.-L. Lions}, \emph{The Method of
1607: Quasi-Reversibility}. Applications to Partial Differential Equations,
1608: Elsevier, New York, 1969.
1609: 
1610: \bibitem{L86}  \textsc{Lop Fat Ho}, \emph{Observabilit\`{e} frontier\`{e} de
1611: l'\`{e}quation des ondes}, C.R. Acad. Sci. Paris, 302, Ser. I, No 12,
1612: pp.~443--446, 1986.
1613: 
1614: \bibitem{Ro74}  \textsc{V.G. Romanov}, I\emph{ntegral Geometry and Inverse
1615: Problems for Hyperbolic Equations}, Springer, New York, 1974.
1616: 
1617: \bibitem{R06_1}  \textsc{V.G. Romanov}, \emph{Carleman estimates for second
1618: order hyperbolic equations}, Sib. Math. J., Vol. 47, No. 1, pp.~135--151,
1619: 2006.
1620: 
1621: \bibitem{R06_2}  \textsc{V.G. Romanov}, \emph{Stability estimates in inverse
1622: problems for hyperbolic equations}, Milan J. of Mathematics, 74,
1623: pp.~357--385, 2006.
1624: 
1625: \bibitem{TY}  \textsc{R. Triggiani and P.F. Yao}, \emph{Carleman estimates
1626: with no lower order terms for general Riemann wave equations. Global
1627: uniqueness and stability in one shot}, Appl.\ Meth. Optim., 46, No. 2/3,
1628: pp.~334--375.
1629: 
1630: \bibitem{XFW03}  \textsc{M. Xu, D.\ Feng, and L.V. Wang}, \emph{Time-domain
1631: reconstruction algorithms and numerical simulations for thermoacoustic
1632: tomography in various geometries}, IEEE Trans. Biomed. Eng., 50,
1633: pp.~1086--1099, 2003.
1634: \end{thebibliography}
1635: 
1636: \end{document}
1637: