This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Carleman contraction mapping for a 1D inverse scattering problem with experimental time-dependent data

Thuy T. Le11footnotemark: 1    Micheal V. Klibanov Department of Mathematics and Statistics, University of North Carolina at Charlotte, Charlotte, NC, 28223, USA, tle55@uncc.edu, mklibanv@uncc.edu (corresponding author) loc.nguyen@uncc.edu.    Loc H. Nguyen11footnotemark: 1    Anders Sullivan US Army Research Laboratory, 2800 Powder Mill Road, Adelphi, MD 20783-1197, USA, (anders.j.sullivan.civ@mail.mil, lam.h.nguyen2civ@mail.mil)    Lam Nguyen22footnotemark: 2
Abstract

It is shown that the contraction mapping principle with the involvement of a Carleman Weight Function works for a Coefficient Inverse Problem for a 1D hyperbolic equation. Using a Carleman estimate, the global convergence of the corresponding numerical method is established. Numerical studies for both computationally simulated and experimentally collected data are presented. The experimental part is concerned with the problem of computing dielectric constants of explosive-like targets in the standoff mode using severely underdetermined data.

Key words: one-dimensional wave equation, Carleman estimate, iterative method, contraction principle, numerical solution, experimental data, global convergence.

AMS subject classification: 35R25, 35R30

1 Introduction

We consider in this paper a 1D Coefficient Inverse Problem (CIP) for a hyperbolic PDE and construct a globally convergent numerical method for it. Unlike all previous works of this group on the convexification method for CIPs, which are cited below in this section, we construct in this paper a sequence of linearized boundary value problems with overdetermined boundary data. Each of these problems is solved using a weighted version of the Quasi-Reversibility Method (QRM). The weight is the Carleman Weight Function (CWF), i.e. this is the function, which is involved as the weight in the Carleman estimate for the corresponding PDE operator, see, e.g. [7, 4, 5, 22, 18, 31, 38, 53] for Carleman estimates. Thus, we call this method “Carleman Quasi-Reversibility Method” (CQRM). The key convergence estimate for this sequence is similar with the one of the conventional contraction mapping principle, see the first item of Remarks 8.1 in section 8. This explains the title of our paper. That estimate implies the global convergence of that sequence to the true solution of our CIP. Furthermore, that estimate implies a rapid convergence. As a result, computations are done here in real time, also, see Remark 11.1 in section 11 as well as section 12. The real time computation is obviously important for our target application, which is described below. The numerical method of this paper can be considered as the second generation of the convexification method for CIPs.

The convexification method ends up with the minimization of a weighted globally strictly convex Tikhonov functional, in which the weight is the CWF. The convexification has been developed by our research group since the first two publications in 1995 [20] and 1997 [21]. Only the theory  was presented in those two originating papers. The start for numerical studies of the convexification was given in the paper [1]. Thus, in follow up publications [24, 25, 30, 16, 14, 15, 24, 25, 27, 28, 29, 32, 33, 30, 40, 50, 51] the theoritical results for the convexification are combined with the numerical ones. In particular, some of these papers work with experimentally collected backscattering data [14, 15, 26, 32, 33]. We also refer to the recently published book of Klibanov and Li [31], in which many of these results are described.

The CIP of our paper has a direct application in the problem of the standoff detection and identification of antipersonnel land mines and improvised explosive devices (IEDs). Thus, in the numerical part of this publication, we present results of the numerical performance of our method for both computationally simulated and experimentally collected data for targets mimicking antipersonnel land mines and IEDs. The experimental data were collected in the field, which is more challenging than the case of the data collected in a laboratory [14, 15, 26, 33]. The data used in this paper were collected by the forward looking radar of the US Army Research Laboratory [46]. Since these data were described in the previous publications of our research group [13, 23, 24, 25, 30, 35, 36, 51], then we do not describe them here. Those previous publications were treating these experimental data by several different numerical methods for 1D CIPs. We also note that the CIP of our paper was applied in [32, 33] to the nonlinear synthetic aperture (SAR) imaging, including the cases of experimentally collected data.

Our goal is to compute approximate values of dielectric constants of targets. We point out that our experimental data are severely underdetermined. Indeed, any target is a 3D object. On the other hand, we have only one experimentally measured time resolved curve for each target. Therefore, we can compute only a sort of an average value of the dielectric constant of each target. This explains the reason of our mathematical modeling here by a 1D hyperbolic PDE rather than by its 3D analog. This mathematical model, along with its frequency domain analog, was considered in all above cited previous publications of our research group, which were working with the experimental data of this paper by a variety of globally convergent inversion techniques.

An estimate of the dielectric constant of a target cannot differentiate with a good certainty between an explosive and a non explosive. However, we hope that these estimates, combined with current targets classification procedures, might potentially result in a decrease of the false alarm rate for those targets. We believe therefore that our results for experimental data might help to address an important application to the standoff detection and identification of antipersonnel land mines and IEDs.

Given a numerical method for a CIP, we call this method globally convergent if:

  1. 1.

    There is a theorem claiming that this method delivers at least one point in a sufficiently small neighborhood of the true solution of that CIP without relying on a good initial guess about this solution.

  2. 2.

    This theorem is confirmed numerically.

CIPs are both highly nonlinear and ill-posed. As a result, conventional least squares cost functionals for them are non convex and typically have many local minima and ravines, see, e.g. [49] for a numerical example of this phenomenon. Conventional numerical methods for CIPs are based on minimizations of those functionals, see, e.g. [8, 11, 12]. The goal of the convexification is to avoid local minima and ravines and to provide accurate and reliable solutions of CIPs.

In the convexification, a CWF is a part of a certain least squares cost functional JλJ_{\lambda}, where λ>1\lambda>1 is the parameter of the CWF. The presence of the CWF ensures that, with a proper choice of the parameter λ,\lambda, the functional JλJ_{\lambda} is strictly convex on a certain bounded convex set SS of an arbitrary diameter d>0d>0 in a Hilbert space. It was proven later in [1] that the existence and uniqueness of the minimizer of JλJ_{\lambda} on SS are also guaranteed. Furthermore, the gradient projection method of the minimization of JλJ_{\lambda} on SS converges globally to that minimizer if starting at an arbitrary point of SS. In addition, if the level of noise δ>0\delta>0 in the data tends to zero, then the gradient projection method delivers a sequence, which converges to the true solution. Thus, since d>0d>0 is an arbitrary number, then this is the global convergence, as long as numerical studies confirm this theory. Furthermore, it was recently established in [33, 40] that the gradient descent method of the minimization of JλJ_{\lambda} on SS also converges globally to the true solution.

First, we come up with the same boundary value problem (BVP) for a nonlinear and nonlocal hyperbolic PDE as we did in the convexification method for the same 1D CIP in [30]. The solution of this BVP immediately provides the target unknown coefficient. However, when solving this BVP, rather than constructing a globally strictly convex cost functional for this BVP, as it was done in [30] and other works on the convexification, we construct a sequence of linearized BVPs. Just as the original BVP, each BVP of this sequence involves a hyperbolic operator with non local terms and overdetermined boundary conditions.

Because of non local terms and the overdetermined boundary conditions, we solve each BVP of this sequence by the new version of the QRM, the Carleman Quasi-Reversibility Method (CQRM). Indeed, the classical QRM solves ill-posed/overdetermined BVPs for linear PDEs, see [37] for the originating work as well as, e.g. [19, 31] for updates. Convergence of the QRM-solution to the true solution is proven via an appropriate Carleman estimates [19, 31]. However, a CWF is not used in the optimizing quadratic functional of the conventional QRM of [19, 31]. Unlike this, a CWF is involved in CQRM. It is exactly the presence of the CWF, which enables us to derive the above mentioned convergence estimate, which is an analog of the one of the contraction principle. This allows us, in turn to establish convergence rate of the sequence of CQRM-solutions to the true solution of that BVP. The latter result ensures that our method is a globally convergent one.

Various versions of the second generation of the convexification method involving CQRM were previously published in [2, 3, 45, 44, 39]. In these publications, globally convergent numerical methods for some nonlinear inverse problems are presented. In [2, 3] CIPs for hyperbolic PDEs in n\mathbb{R}^{n} are considered. The main difference between [2, 3] and our work is that it is assumed in [2, 3] that one of initial conditions is not vanishing everywhere in the closed domain of interest. In other words, papers [2, 3] work exactly in the framework of the Bukhgeim-Klibanov method, see [6] for the originating work on this method and, e.g. [4, 5, 17, 22, 18, 31, 53] for some follow up publications. On the other hand, in all above cited works on the convexification, including this one, either the initial condition is the δ\delta-function, or a similar requirement holds for the Helmholtz equation. The only exception is the paper [29], which also works within the framework of the method of [6]. We refer to papers [10, 34] for some numerical methods for CIPs with the Dirichlet-to-Neumann map data. In these works the number mm of free variables in the data exceeds the number nn of free variables in the unknown coefficient, m>n.m>n. In our paper m=n=1.m=n=1. Also, m=nm=n in all other above cited works on the convexification.

There is a classical Gelfand-Levitan method [41] for solutions of 1D CIPs for some hyperbolic PDEs. This method does not rely on the optimization, and, therefore, avoids the phenomenon of local minima. It reduces the original CIP to a linear integral equation of the second kind. This equation is called “Gelfand-Levitan equation” (GL). However, the questions about uniqueness and stability of the solution of GL for the case of noisy data is open, see, e.g. Lemma 2.4 in the book [48, Chapter 2]. This lemma is valid only in the case of noiseless data. However, the realistic data are always noisy. In addition, it was demonstrated numerically in [13] that GL cannot work well for exactly the same experimental data as the ones we use in the current paper. On the other hand, it was shown in [24, 25, 51] that the convexification works well with these data. The same is true for the method of the current paper.

Uniqueness and Lipschitz stability theorems of the CIP considered here are well known. Indeed, it was shown in, e.g. [51] that, using the same change of variables as the one considered in section 2 below, one can reduce our CIP to a similar CIP for the equation vtt=vyy+r(y)v,yv_{tt}=v_{yy}+r\left(y\right)v,y\in\mathbb{R} with the unknown coefficient r(y).r\left(y\right). We refer to [48, Theorem 2.6 of Chapter 2] for the Lipschitz stability estimate for the latter CIP. In addition, both uniqueness and Lipschitz stability results for our CIP actually follow from Theorem 8.1 below as well as from the convergence analysis in two other works of our research group on the convexification method for this CIP [30, 51].

This paper is arranged as follows. In section 2 we state both forward and inverse problems. In section 3 we derive a nonlinear BVP with non local terms. In section 4 we describe iterations of our method to solve that BVP. In section 5 we formulate the Carleman estimate for the principal part of the PDE operator of that BVP. In section 6 we prove the strong convexity of a functional of section 5 on an appropriate bounded set. In section 7 we formulate two methods for finding the unique minimizer of that functional: gradient descent method and gradient projection method and prove the global convergence to the minimizer for each of them. In section 8 we establish the contraction mapping property and prove the global convergence of the method of section 4. In section 9 we formulate two more global convergence theorems, which follow from results of sections 7 and 8. Numerical results with simulated and experimental data are presented in Section 10 and 11 respectively. Concluding remarks are given in section 12.

2 Statements of Forward and Inverse Problems

Below all functions are real valued ones. Let b>1b>1 be a known number, xx\in\mathbb{R} be the spatial variable and the function c(x)C3()c(x)\in C^{3}(\mathbb{R}), represents the spatially distributed dielectric constant. We assume that

c(x)\displaystyle c(x) [1,b],x,\displaystyle\in[1,b],x\in\mathbb{R}, (2.1)
c(x)=1\displaystyle c(x)=1 if x(,ε][1,),\displaystyle\mbox{ if }x\in(-\infty,\varepsilon]\cup[1,\infty), (2.2)

where ε(0,1)\varepsilon\in\left(0,1\right) is a small number. Let TT be a positive number. We consider the following Cauchy problem for a 1D hyperbolic PDE with a variable coefficient in the principal part of the operator:

c(x)utt=uxx, x,t(0,T),c(x)u_{tt}=u_{xx},\text{ }x\in\mathbb{R},t\in\left(0,T\right), (2.3)
u(x,0)=0,ut(x,0)=δ(x).u\left(x,0\right)=0,u_{t}\left(x,0\right)=\delta\left(x\right). (2.4)

The problem of finding the function u(x,t)u(x,t) from conditions (2.3), (2.4) is our Forward Problem.

Let τ(x)\tau\left(x\right) be the travel time needed for the wave to travel from the point source at {0}\left\{0\right\} to the point xx,

τ(x)=0xc(s)ds\tau\left(x\right)=\mathop{\displaystyle\int}\limits_{0}^{x}\sqrt{c(s)}ds (2.5)

By (2.5) the following 1D analog of the eikonal equation is valid:

τ(x)=c(x).\tau^{\prime}\left(x\right)=\sqrt{c\left(x\right)}. (2.6)

Let H(z),zH(z),z\in\mathbb{R} be the Heaviside function,

H(z)={1,z>0,0,z<0.H(z)=\left\{\begin{array}[]{c}1,\quad z>0,\\ 0,\quad z<0.\end{array}\right.

Lemma 2.1 [30, Lemma 2.1]. For x0x\geq 0, the function u(x,t)u(x,t) has the form

u(x,t)=H(tτ(x))2c1/4(x)+u^(x,t),u(x,t)=\frac{H(t-\tau\left(x\right))}{2c^{1/4}(x)}+\hat{u}(x,t), (2.7)

where the function u^C2(tτ(x)),u^(x,τ(x))=0.\hat{u}\in C^{2}(t\geq\tau\left(x\right)),\hat{u}(x,\tau\left(x\right))=0. Furthermore,

limtτ÷(x)u(x,t)=12c1/4(x).\lim_{t\rightarrow\tau^{\div}\left(x\right)}u(x,t)=\frac{1}{2c^{1/4}(x)}. (2.8)

Lemma 2.2 [Absorbing boundary condition] [30, 50]. Let b>εb>\varepsilon be the number in (2.1). Let x1bx_{1}\geq b and x2εx_{2}\leq\varepsilon be two arbitrary numbers. Then the solution u(x,t)u(x,t) of Forward Problem (2.3), (2.4) satisfies the absorbing boundary condition at x=x1x=x_{1} and x=x2x=x_{2}, i.e.

ux(x1,t)+ut(x1,t)=0 for t(0,T),u_{x}(x_{1},t)+u_{t}(x_{1},t)=0\mbox{ for }t\in(0,T), (2.9)
ux(x2,t)ut(x2,t)=0 for t(0,T).u_{x}(x_{2},t)-u_{t}(x_{2},t)=0\mbox{ for }t\in(0,T). (2.10)

We are interested in the following inverse problem:

Coefficient Inverse Problem (CIP). Suppose that the following two functions g0(t)g_{0}(t), g1(t)g_{1}(t) are known:

u(ε,t)=g0(t),ux(ε,t)=g1(t),t(0,T).u\left(\varepsilon,t\right)=g_{0}\left(t\right),\quad u_{x}\left(\varepsilon,t\right)=g_{1}\left(t\right),\quad t\in\left(0,T\right). (2.11)

Determine the function c(x)c(x) for x(ε,1)x\in(\varepsilon,1), assuming that the number b>εb>\varepsilon in (2.1) is known.

Remark 2.1. Note that only the function g0(t)g_{0}\left(t\right) can be measured. As to the function g1(t),g_{1}\left(t\right), it follows from (2.10) that

g1(t)=g0(t).g_{1}\left(t\right)=g_{0}^{\prime}\left(t\right). (2.12)

We differentiate noisy function using the Tikhonov regularization method [52]. Since this method is well known, we do not describe it here.

3 A Boundary Value Problem for a Nonlinear PDE With Non-Local Terms

We now introduce a change of variable

q(x,t)=u(x,t+τ(x)).q(x,t)=u(x,t+\tau(x)). (3.1)

We will consider the function q(x,t)q(x,t) only for t0.t\geq 0. Using (2.6) and (3.1), we obtain

qxx2qxtτqtτ′′=0.q_{xx}-2q_{xt}\tau^{\prime}-q_{t}\tau^{\prime\prime}=0. (3.2)

Furthermore, it follows from (2.8)

q(x,0)=12c1/4(x)0.q(x,0)=\frac{1}{2c^{1/4}(x)}\neq 0. (3.3)

By (2.5) and (3.3)

τ′′(x)=qx(x,0)2q3(x,0).\tau^{\prime\prime}(x)=-\frac{q_{x}(x,0)}{2q^{3}(x,0)}. (3.4)

Substituting (2.6), (3.3) and (3.4) in (3.2), we obtain

L(q)=qxxqxt12q2(x,0)+qtqx(x,0)2q3(x,0)=0.L(q)=q_{xx}-q_{xt}\frac{1}{2q^{2}(x,0)}+q_{t}\frac{q_{x}(x,0)}{2q^{3}(x,0)}=0. (3.5)

Equation, (3.5) is a nonlinear PDE with respect to the function q(x,t)q(x,t) with nonlocal terms qx(x,0)q_{x}(x,0) and q(x,0)q(x,0). We now need to obtain boundary conditions for the function q.q. By (2.2) and (2.5) τ(x)=x\tau\left(x\right)=x for x[0,ε].x\in\left[0,\varepsilon\right]. Hence, (2.11), (2.12) and (3.1) lead to

q(ε,t)=g0(t+ε),qx(ε,t)=2g0(t+ε), t(0,T),q\left(\varepsilon,t\right)=g_{0}\left(t+\varepsilon\right),q_{x}\left(\varepsilon,t\right)=2g_{0}^{\prime}\left(t+\varepsilon\right),\text{ }t\in\left(0,T\right), (3.6)

We will solve equation (3.5) in the rectangle

Ω={(x,t):x(ε,b),t(0,T)}.\Omega=\left\{{(x,t):x\in(\varepsilon,b),t\in(0,T)}\right\}. (3.7)

By (3.1)

qx(x,t)=ux(x,t+τ(x))+ut(x,t+τ(x))τ(x)q_{x}(x,t)=u_{x}(x,t+\tau(x))+u_{t}(x,t+\tau(x))\tau^{\prime}(x) (3.8)

By (2.1), (2.2) and (2.5) τ(b)=1\tau^{\prime}(b)=1. Hence, using (2.9) and (3.8), we obtain

qx(b,t)=0.q_{x}(b,t)=0. (3.9)

It follows from (2.1) and (3.3) that

12b1/4q(x,0)12,x[ε,b].\frac{1}{2b^{1/4}}\leq q(x,0)\leq\frac{1}{2},\quad x\in[\varepsilon,b]. (3.10)

In addition, we need qC2(Ω¯)q\in C^{2}(\overline{\Omega}) and we also need to bound the norm qC2(Ω¯)\left\|q\right\|_{C^{2}(\overline{\Omega})} from the above. Let R>0R>0 be an arbitrary number. We define the set B(R,g0)B(R,g_{0}) as

B(R,g0)={qH4(Ω):qH4(Ω)<R,q(ε,t+ε)=g0(t+ε),qx(ε,t)=2g0(t+ε),qx(b,t)=0,12b1/4q(x,0)1/2,x[ε,b].B(R,g_{0})=\left\{\begin{array}[]{c}q\in H^{4}(\Omega):\left\|q\right\|_{H^{4}(\Omega)}<R,\\ q(\varepsilon,t+\varepsilon)=g_{0}(t+\varepsilon),q_{x}(\varepsilon,t)=2g_{0}^{\prime}\left(t+\varepsilon\right),\\ q_{x}(b,t)=0,\\ \frac{1}{2b^{1/4}}\leq q\left(x,0\right)\leq 1/2,x\in[\varepsilon,b].\end{array}\right. (3.11)

We assume below that

B(R,g0).B(R,g_{0})\neq\varnothing. (3.12)

By embedding theorem B(R,g0)¯C2(Ω¯)\overline{B(R,g_{0})}\subset C^{2}(\overline{\Omega}) and

qC2(Ω¯)KR, qB(R,g0)¯.\left\|q\right\|_{C^{2}(\overline{\Omega})}\leq KR,\text{ }\forall q\in\overline{B(R,g_{0})}. (3.13)

where the constant K=K(Ω)>0K=K\left(\Omega\right)>0 depends only on the domain Ω.\Omega. Using (3.10), we define the function q0(x,0)q^{0}\left(x,0\right) as:

q0(x,0)\displaystyle q^{0}\left(x,0\right) =\displaystyle= {q(x,0) if q(x,0)[1/(2b1/4),1/2],1/(2b1/4), if q(x,0)<1/(2b1/4),1/2 if q(x,0)>1/2. qB(R,g0)¯,\displaystyle\left\{\begin{array}[]{c}q\left(x,0\right)\text{ if }q\left(x,0\right)\in\left[1/\left(2b^{1/4}\right),1/2\right],\\ 1/\left(2b^{1/4}\right),\text{ if }q\left(x,0\right)<1/\left(2b^{1/4}\right),\\ 1/2\text{ if }q\left(x,0\right)>1/2.\end{array}\right.\text{ }\forall q\in\overline{B(R,g_{0})}, (3.17)
x\displaystyle\forall x \displaystyle\in [ε,b].\displaystyle\left[\varepsilon,b\right].

Then the function q0(x,0)q^{0}\left(x,0\right) is piecewise continuously differentiable in [ε,b]\left[\varepsilon,b\right] and by (3.13) and (3.17)

q0(x,0)C[ε,b],max[ε,b]|qx0(x,0)|KR, qB(R,g0),\left\|q^{0}\left(x,0\right)\right\|_{C\left[\varepsilon,b\right]},\max_{\left[\varepsilon,b\right]}\left|q_{x}^{0}\left(x,0\right)\right|\leq KR,\text{ }\forall q\in B(R,g_{0}), (3.18)
12b1/4q0(x,0)12,x[ε,b].\frac{1}{2b^{1/4}}\leq q^{0}(x,0)\leq\frac{1}{2},\quad x\in[\varepsilon,b]. (3.19)

Thus, (3.5), (3.6), (3.9), (3.17) and (3.19) result in the following BVP for a nonlinear PDE with non-local terms:

qxx(x,t)qxt(x,t)12(q0(x,0))2+qt(x,t)qx(x,0)2(q0(x,0))3=0 in Ω,q_{xx}(x,t)-q_{xt}(x,t)\frac{1}{2\left(q^{0}(x,0)\right)^{2}}+q_{t}(x,t)\frac{q_{x}(x,0)}{2\left(q^{0}(x,0)\right)^{3}}=0\text{ in }\Omega, (3.20)
q(ε,t)=g0(t+ε), qx(ε,t)=2g0(t+ε), qx(b,t)=0.q(\varepsilon,t)=g_{0}(t+\varepsilon),\text{ }q_{x}(\varepsilon,t)=2g_{0}^{\prime}(t+\varepsilon),\text{ }q_{x}(b,t)=0. (3.21)

Thus, we focus below on the numerical solution of the following problem:

Problem 3.1. Find a function qB(R,g0)q\in B(R,g_{0}) satisfying conditions (3.20), (3.21), where the function q0(x,0)q^{0}(x,0) is defined in (3.17).

Suppose that we have solved Problem 3.1. Then, using (3.3) and (3.17), we set

c(x)=1(2q0(x,0))4.c\left(x\right)=\frac{1}{\left(2q^{0}\left(x,0\right)\right)^{4}}. (3.22)

4 Numerical Method for Problem 3.1

4.1 The function q0(x,t)q_{0}\left(x,t\right)

We now find the first approximation q0(x,t)q_{0}(x,t) for the function q(x,t).q(x,t). Using (2.2), we choose c(x)1c\left(x\right)\equiv 1 as the first guess for the function c(x).c\left(x\right). Hence, by (3.3),

q0(x,0)12.q_{0}(x,0)\equiv\frac{1}{2}. (4.1)

We now need to find the function q0(x,t).q_{0}(x,t). To do this, drop the nonlinear third term in the left hand side of equation (3.20) and, using (4.1) and (3.17), set 1/(2q0(x,0))2:=21/\left(2q^{0}(x,0)\right)^{2}:=2. Then (3.20), (3.21) become:

q0xx(x,t)2q0xt(x,t)=0 in Ω,q_{0xx}(x,t)-2q_{0xt}(x,t)=0\text{ in }\Omega, (4.2)
q0(ε,t)=g0(t+ε), q0x(ε,t)=2g0(t+ε),q0x(b,t)=0.q_{0}(\varepsilon,t)=g_{0}(t+\varepsilon),\text{ }q_{0x}(\varepsilon,t)=2g_{0}^{\prime}(t+\varepsilon),q_{0x}(b,t)=0. (4.3)

BVP (4.2), (4.3) has overdetermined boundary conditions (4.3). Typically, QRM works well for BVPs for PDEs with overdetermined boundary conditions [19, 31]. Therefore, we solve BVP (4.2), (4.3) via CQRM. This means that we consider the following minimization problem:

Minimization Problem Number 0. Assuming (3.12), minimize the functional Jλ,β(0):B(R,g0)¯J_{\lambda,\beta}^{(0)}:\overline{B(R,g_{0})}\rightarrow\mathbb{R} on the set ,

Jλ,β(0)(q0)=Ω(q0xx2q0xt)2e2λφdxdt+βq0H4(Ω)2,J_{\lambda,\beta}^{(0)}(q_{0})=\mathop{\displaystyle\int}\limits_{\Omega}\left(q_{0xx}-2q_{0xt}\right)^{2}e^{2\lambda\varphi}dxdt+\beta\|q_{0}\|_{H^{4}(\Omega)}^{2}, (4.4)

where e2λφe^{2\lambda\varphi} is the Carleman Weight Function for the operator x2t\partial_{x}-2\partial_{t} [50, 51]

e2λφ=e2λ(x+αt),e^{2\lambda\varphi}=e^{-2\lambda\left(x+\alpha t\right)}, (4.5)

 where α(0,1/2)\alpha\in\left(0,1/2\right) is the parameter, and β(0,1)\beta\in\left(0,1\right) is the regularization parameter. Both parameters will be chosen later.

Theorem 6.1 guarantees that, for appropriate values of parameters λ,β,\lambda,\beta, there exists unique minimizer q0,minq_{0,\min} B(R,g0)¯\overline{B(R,g_{0})} of the functional Jλ,β(0)(q0).J_{\lambda,\beta}^{(0)}(q_{0}).

4.2 The function qn(x,t)q_{n}\left(x,t\right) for n1n\geq 1

Assume that functionals Jλ,β(m)(qm):B(R,g0)¯J_{\lambda,\beta}^{(m)}(q_{m}):\overline{B(R,g_{0})}\rightarrow\mathbb{R} are defined for m=0,,n1m=0,...,n-1 and their minimizers functions {qm,min}m=0n1B(R,g0)¯\left\{q_{m,\min}\right\}_{m=0}^{n-1}\subset\overline{B(R,g_{0})} are constructed already, all for the same values of parameters λ,α,β\lambda,\alpha,\beta. Replace in (3.20) q(x,t)q\left(x,t\right) with qn(x,t),q0(x,0)q_{n}\left(x,t\right),q^{0}(x,0) with qn1,min0(x,0),qx(x,t)q_{n-1,\min}^{0}(x,0),q_{x}\left(x,t\right) with xq(n1),min(x,t)\partial_{x}q_{\left(n-1\right),\min}\left(x,t\right) and qt(x,t)q_{t}\left(x,t\right) with tq(n1),min(x,t).\partial_{t}q_{\left(n-1\right),\min}\left(x,t\right). Then problem (3.20), (3.21) becomes a linear one with respect to the function qn(x,t),q_{n}\left(x,t\right),

qnxx(x,t)qnxt(x,t)2(qn1,min0(x,0))2+tq(n1),min(x,t)xq(n1)min(x,0)2(qn1,min0(x,0))3=0 in Ω,q_{nxx}(x,t)-\frac{q_{nxt}(x,t)}{2\left(q_{n-1,\min}^{0}(x,0)\right)^{2}}+\frac{\partial_{t}q_{\left(n-1\right),\min}\left(x,t\right)\partial_{x}q_{\left(n-1\right)\min}(x,0)}{2\left(q_{n-1,\min}^{0}(x,0)\right)^{3}}=0\text{ in }\Omega, (4.6)
qn(ε,t)=g0(t+ε), qnx(ε,t)=2g0(t+ε),qnx(b,t)=0.q_{n}(\varepsilon,t)=g_{0}(t+\varepsilon),\text{ }q_{nx}(\varepsilon,t)=2g_{0}^{\prime}(t+\varepsilon),q_{nx}(b,t)=0. (4.7)

To solve problem (4.6), (4.7), we consider the following minimization problem:

Minimization Problem Number nn. Assuming (3.12), minimize the functional Jλ,β(n):H4(Ω)J_{\lambda,\beta}^{(n)}:H^{4}\left(\Omega\right)\rightarrow\mathbb{R} on the set B(R,g0),B(R,g_{0}),

Jλ,β(n)(qn)=Ω(qnxx(x,t)qnxt(x,t)2(qn1,min0(x,0))2+tq(n1),min(x,t)xq(n1),min(x,0)2(qn1,min0(x,0))3)2e2λφ𝑑x𝑑tJ_{\lambda,\beta}^{(n)}(q_{n})=\int_{\Omega}\Big{(}q_{nxx}(x,t)-\frac{q_{nxt}(x,t)}{2\big{(}q_{n-1,\min}^{0}(x,0)\big{)}^{2}}+\frac{\partial_{t}q_{\left(n-1\right),\min}\left(x,t\right)\partial_{x}q_{\left(n-1\right),\min}(x,0)}{2\big{(}q_{n-1,\min}^{0}(x,0)\big{)}^{3}}\Big{)}^{2}e^{2\lambda\varphi}dxdt
+βqnH4(Ω)2.+\beta\left\|q_{n}\right\|_{H^{4}\left(\Omega\right)}^{2}. (4.8)

Suppose that there exists unique minimizer qn,minB(R,g0)¯q_{n,\min}\in\overline{B(R,g_{0})} of the functional Jλ,β(n)(qn).J_{\lambda,\beta}^{(n)}(q_{n}). Then, following (3.22), (3.17) and (3.22), we set

cn(x)=1(2qn,min0(x,0))4.c_{n}\left(x\right)=\frac{1}{\left(2q_{n,\min}^{0}\left(x,0\right)\right)^{4}}. (4.9)

The rest of the analytical part of this paper is devoted to the convergence analysis of the iterative numerical method presented in this section.

5 The Carleman Estimate

In this section we formulate the Carleman estimate, which is the main tool of our construction. This estimate follows from Theorem 3.1 of [30] as well as from (3.18). Let q(x,t)B(R,g0)¯q\left(x,t\right)\in\overline{B(R,g_{0})} be an arbitrary function and let the function q0(x,0)q^{0}\left(x,0\right) be constructed from the function q(x,t)q\left(x,t\right) as in (3.17). Consider the operator L0:H2(Ω)L2(Ω),L_{0}:H^{2}(\Omega)\rightarrow L_{2}(\Omega),

L0u=uxx(x,t)uxt(x,t)12(q0(x,0))2,for all (x,t)Ω.L_{0}u=u_{xx}(x,t)-u_{xt}(x,t)\frac{1}{2\left(q^{0}(x,0)\right)^{2}},\quad\mbox{for all }(x,t)\in\Omega.

Theorem 5.1 (Carleman estimate [30]). There exists a number α0=α0(R,Ω)>0\alpha_{0}=\alpha_{0}\left(R,\Omega\right)>0 depending only on R,ΩR,\Omega such that for any α(0,α0)\alpha\in(0,\alpha_{0}) there exists a sufficiently large number λ0=λ0(R,Ω,α)>1\lambda_{0}=\lambda_{0}\left(R,\Omega,\alpha\right)>1 depending only on RR,Ω,α\Omega,\alpha such that for all λλ0\lambda\geq\lambda_{0} and for all functions vH2(Ω)v\in H^{2}(\Omega) the following Carleman estimate holds:

Ω(L0v)2e2λφdxdt\mathop{\displaystyle\int}\limits_{\Omega}\left(L_{0}v\right)^{2}e^{2\lambda\varphi}dxdt
CΩ(λ(vx2+vt2)+λ3v2)e2λφdxdt+Cεb(λvx2(x,0)+λ3v2(x,0))e2λxdx\geq C\mathop{\displaystyle\int}\limits_{\Omega}\big{(}\lambda\left(v_{x}^{2}+v_{t}^{2}\right)+\lambda^{3}v^{2}\big{)}e^{2\lambda\varphi}dxdt+C\mathop{\displaystyle\int}\limits_{\varepsilon}^{b}\big{(}\lambda v_{x}^{2}(x,0)+\lambda^{3}v^{2}(x,0)\big{)}e^{-2\lambda x}dx
C0T(λ(vx2(ε,t)+vt2(ε,t))+λ3v2(ε,t))e2λ(ε+αt)dt-C\mathop{\displaystyle\int}\limits_{0}^{T}\left(\lambda\left(v_{x}^{2}(\varepsilon,t)+v_{t}^{2}(\varepsilon,t)\right)+\lambda^{3}v^{2}(\varepsilon,t)\right)e^{-2\lambda(\varepsilon+\alpha t)}dt (5.1)
Cεb(λvx2(x,T)+λ3v2(x,T))e2λ(x+αT)dx.-C\mathop{\displaystyle\int}\limits_{\varepsilon}^{b}\left(\lambda v_{x}^{2}(x,T)+\lambda^{3}v^{2}(x,T)\right)e^{-2\lambda(x+\alpha T)}dx.

 

Remark 5.1. Here and everywhere below C=C(R,Ω,α)>0C=C\left(R,\Omega,\alpha\right)>0 denotes different constants depending only on listed parameters.

6 Strict Convexity of Functional (4.8) on the Set B(R,g0)B(R,g_{0}), Existence and Uniqueness of Its Minimizer

Functional (4.8) is quadratic. We prove in this section that it is strictly convex on the set B(R,g0)¯.\overline{B(R,g_{0})}. In addition, we prove existence and uniqueness of its minimizer on this set. Although similar results were proven in many of the above cited publications on the convexification, see, e.g. [30] for the closest one, there are some peculiarities here, which are important for our convergence analysis, see Remarks 6.1, 6.2 below.

Introduce the subspace H04(Ω)H4(Ω)H_{0}^{4}\left(\Omega\right)\subset H^{4}\left(\Omega\right) as:

H04(Ω)={vH4(Ω):v(ε,t)=vx(ε,t)=vx(b,t)=0}.H_{0}^{4}\left(\Omega\right)=\left\{v\in H^{4}\left(\Omega\right):v\left(\varepsilon,t\right)=v_{x}\left(\varepsilon,t\right)=v_{x}\left(b,t\right)=0\right\}. (6.1)

Denote [,]\left[,\right] the scalar product in the space H4(Ω).H^{4}\left(\Omega\right). Also, denote

An(q)(x,t)=qxx(x,t)qxt(x,t)12(qn1,min0(x,0))2,A_{n}\left(q\right)\left(x,t\right)=q_{xx}(x,t)-q_{xt}(x,t)\frac{1}{2\left(q_{n-1,\min}^{0}(x,0)\right)^{2}}, (6.2)
Bn(x,t)=tq(n1),min(x,t)xq(n1),min(x,0)2(qn1,min0(x,0))3.B_{n}\left(x,t\right)=\partial_{t}q_{\left(n-1\right),\min}\left(x,t\right)\frac{\partial_{x}q_{\left(n-1\right),\min}(x,0)}{2\left(q_{n-1,\min}^{0}(x,0)\right)^{3}}. (6.3)

Theorem 6.1. 1. For any set of parameters λ,β\lambda,\beta and for any qB(R,g0)¯q\in\overline{B(R,g_{0})} the functional Jλ,β(n)J_{\lambda,\beta}^{\left(n\right)} has the Fréchet derivative (Jλ,β(n)(q))H04(Ω).\left(J_{\lambda,\beta}^{\left(n\right)}(q)\right)^{\prime}\in H_{0}^{4}\left(\Omega\right). The formula for (Jλ,β(n)(q))\left(J_{\lambda,\beta}^{\left(n\right)}(q)\right)^{\prime} is:

(Jλ,β(n)(q))(h)=2Ω(An(q)(x,t)+Bn(x,t))An(h)(x,t)e2λφdxdt+2β[q,h],\left(J_{\lambda,\beta}^{\left(n\right)}(q)\right)^{\prime}\left(h\right)=2\mathop{\displaystyle\int}\limits_{\Omega}\left(A_{n}\left(q\right)\left(x,t\right)+B_{n}\left(x,t\right)\right)A_{n}\left(h\right)\left(x,t\right)e^{2\lambda\varphi}dxdt+2\beta\left[q,h\right], (6.4)

for all hH04(Ω).h\in H_{0}^{4}\left(\Omega\right).

2. This derivative is Lipschitz continuous in B(R,g0)¯\overline{B(R,g_{0})}, i.e. there exists a constant D>0D>0 such that

(Jλ,β(n)(q2))(Jλ,β(n)(q1))H4(Ω)Dq2q1H4(Ω),for all q1,q2B(R,g0)¯.\left\|\left(J_{\lambda,\beta}^{\left(n\right)}(q_{2})\right)^{\prime}-\left(J_{\lambda,\beta}^{\left(n\right)}(q_{1})\right)^{\prime}\right\|_{H^{4}\left(\Omega\right)}\leq D\left\|q_{2}-q_{1}\right\|_{H^{4}\left(\Omega\right)},\quad\mbox{for all }q_{1},q_{2}\in\overline{B(R,g_{0})}. (6.5)

3. Let α0=α0(R,Ω)>0,α(0,α0)\alpha_{0}=\alpha_{0}\left(R,\Omega\right)>0,\alpha\in\left(0,\alpha_{0}\right) and λ0=λ0(R,Ω,α)1\lambda_{0}=\lambda_{0}\left(R,\Omega,\alpha\right)\geq 1 be the numbers of Theorem 5.1. Then there exists a sufficiently large constant

λ1=λ1(R,Ω,α)λ0>1\lambda_{1}=\lambda_{1}\left(R,\Omega,\alpha\right)\geq\lambda_{0}>1 (6.6)

depending only on listed parameters such that for all λλ1\lambda\geq\lambda_{1} and for all β[2eλαT,1)\beta\in\left[2e^{-\lambda\alpha T},1\right) the functional Jλ,β(n)(q)J_{\lambda,\beta}^{\left(n\right)}(q) is strictly convex on the set B(R,g)¯.\overline{B\left(R,g\right)}. More precisely, let qB(R,g)¯q\in\overline{B(R,g)} be an arbitrary function and also let the function q+hB(R,g)¯.q+h\in\overline{B(R,g)}. Then the following inequality holds:

Jλ,β(n)(q+h)Jλ,β(n)(q)(Jλ,β(n)(q))(h)CΩ[λ(hx2+ht2)+λ3h2]e2λφdxdtJ_{\lambda,\beta}^{\left(n\right)}(q+h)-J_{\lambda,\beta}^{\left(n\right)}(q)-\left(J_{\lambda,\beta}^{\left(n\right)}(q)\right)^{\prime}(h)\geq C\mathop{\displaystyle\int}\limits_{\Omega}\Big{[}\lambda\left(h_{x}^{2}+h_{t}^{2}\right)+\lambda^{3}h^{2}\Big{]}e^{2\lambda\varphi}dxdt
+Cεb(λhx2(x,0)+λ3h2(x,0))e2λxdx+β2hH4(Ω)2, λλ1.+C\mathop{\displaystyle\int}\limits_{\varepsilon}^{b}\big{(}\lambda h_{x}^{2}(x,0)+\lambda^{3}h^{2}(x,0)\big{)}e^{-2\lambda x}dx+\frac{\beta}{2}\left\|h\right\|_{H^{4}\left(\Omega\right)}^{2},\text{ }\forall\lambda\geq\lambda_{1}. (6.7)

4. For any λλ1\lambda\geq\lambda_{1} there exists unique minimizer

qn,minB(R,g0)¯q_{n,\min}\in\overline{B(R,g_{0})} (6.8)

 of the functional Jλ,β(n)(q)J_{\lambda,\beta}^{\left(n\right)}(q) on the set B(R,g0)¯\overline{B(R,g_{0})} and the following inequality holds:

[(Jλ,β(n)(q)),qqn,min]0, qB(R,g0)¯,\left[\left(J_{\lambda,\beta}^{\left(n\right)}(q)\right)^{\prime},q-q_{n,\min}\right]\geq 0,\text{ }\forall q\in\overline{B(R,g_{0})}, (6.9)

where [,]\left[,\right] denotes the scalar product in H4(Ω).H^{4}\left(\Omega\right).

Remark 6.1. Even though the expression in the right hand side of (6.4) is linear with respect to the function q,q, we cannot use Riesz theorem here to prove existence and uniqueness of the minimizer qn,min,q_{n,\min}, at which (Jλ,β(n)(qn,min))=0.\left(J_{\lambda,\beta}^{\left(n\right)}(q_{n,\min})\right)^{\prime}=0. Rather, all what we can prove is (6.9). This is because we need to ensure that the function qn,minB(R,g0)¯q_{n,\min}\in\overline{B(R,g_{0})}.

Proof of Theorem 6.1. Since both function q,q+hB(R,g)¯q,q+h\in\overline{B(R,g)} satisfy the same boundary conditions, then

hH04(Ω).h\in H_{0}^{4}\left(\Omega\right). (6.10)

By (4.8) and (6.10)

Jλ,β(n)(q+h)Jλ,β(n)(q)=2Ω(An(q)(x,t)+Bn(x,t))An(h)(x,t)e2λφdxdt+2β[q,h]J_{\lambda,\beta}^{(n)}(q+h)-J_{\lambda,\beta}^{(n)}(q)=2\mathop{\displaystyle\int}\limits_{\Omega}\left(A_{n}\left(q\right)\left(x,t\right)+B_{n}\left(x,t\right)\right)A_{n}\left(h\right)\left(x,t\right)e^{2\lambda\varphi}dxdt+2\beta\left[q,h\right]
+Ω[An(h)(x,t)]2e2λφdxdt+βhH4(Ω)2, hH04(Ω).+\mathop{\displaystyle\int}\limits_{\Omega}\left[A_{n}\left(h\right)\left(x,t\right)\right]^{2}e^{2\lambda\varphi}dxdt+\beta\left\|h\right\|_{H^{4}\left(\Omega\right)}^{2},\text{ }\forall h\in H_{0}^{4}\left(\Omega\right). (6.11)

The expression in the first line of (6.11) coincides with expression in the right hand side of (6.4). In fact, this is a bounded linear functional mapping H04(Ω)H_{0}^{4}\left(\Omega\right) in \mathbb{R}. Therefore, by Riesz theorem, there exists unique function J~n(q)H04(Ω)\widetilde{J}_{n}\left(q\right)\in H_{0}^{4}\left(\Omega\right) such that

[J~n(q),h]=2Ω(An(q)(x,t)+Bn(x,t))An(h)(x,t)e2λφdxdt+2β[q,h], hH04(Ω).\left[\widetilde{J}_{n}\left(q\right),h\right]=2\mathop{\displaystyle\int}\limits_{\Omega}\left(A_{n}\left(q\right)\left(x,t\right)+B_{n}\left(x,t\right)\right)A_{n}\left(h\right)\left(x,t\right)e^{2\lambda\varphi}dxdt+2\beta\left[q,h\right],\text{ }\forall h\in H_{0}^{4}\left(\Omega\right).

In addition, it is clear from (6.11) and (LABEL:5.9) that

limhH4(Ω)0{1hH4(Ω)[Jλ,β(n)(q+h)Jλ,β(n)(q)[J~n(q),h]]}=0.\lim_{\left\|h\right\|_{H^{4}\left(\Omega\right)}\rightarrow 0}\left\{\frac{1}{\left\|h\right\|_{H^{4}\left(\Omega\right)}}\left[J_{\lambda,\beta}^{(n)}(q+h)-J_{\lambda,\beta}^{(n)}(q)-\left[\widetilde{J}_{n}\left(q\right),h\right]\right]\right\}=0.

Therefore,

J~n(q)=(Jλ,β(n)(q))H04(Ω)\widetilde{J}_{n}\left(q\right)=\left(J_{\lambda,\beta}^{\left(n\right)}(q)\right)^{\prime}\in H_{0}^{4}\left(\Omega\right) (6.12)

is the Fréchet derivative of the functional Jλ,β(n)(q):B(R,g0)¯J_{\lambda,\beta}^{\left(n\right)}(q):\overline{B(R,g_{0})}\rightarrow\mathbb{R} at the point qB(R,g0)¯,q\in\overline{B(R,g_{0})}, and the right hand side of (6.4) indeed represents (Jλ,β(n)(q))(h).\left(J_{\lambda,\beta}^{\left(n\right)}(q)\right)^{\prime}\left(h\right). Estimate (6.5) obviously follows from (6.4).

We now prove strict convexity estimate (6.7). To do this, we apply Carleman estimate (5.1) to the third line of (6.11). We obtain

Jλ,β(n)(q+h)Jλ,β(n)(q)(Jλ,β(n)(q))(h)CΩ(λ(hx2+ht2)+λ3h2)e2λφdxdt+Cεb(λhx2(x,0)+λ3h2(x,0))e2λxdxCεb(λhx2(x,T)+λ3h2(x,T))e2λ(x+αT)dx+βhH4(Ω)2, hH04(Ω).J_{\lambda,\beta}^{(n)}(q+h)-J_{\lambda,\beta}^{(n)}(q)-\left(J_{\lambda,\beta}^{\left(n\right)}(q)\right)^{\prime}\left(h\right)\\ \geq C\mathop{\displaystyle\int}\limits_{\Omega}\big{(}\lambda\left(h_{x}^{2}+h_{t}^{2}\right)+\lambda^{3}h^{2}\big{)}e^{2\lambda\varphi}dxdt+C\mathop{\displaystyle\int}\limits_{\varepsilon}^{b}\big{(}\lambda h_{x}^{2}(x,0)+\lambda^{3}h^{2}(x,0)\big{)}e^{-2\lambda x}dx\\ -C\mathop{\displaystyle\int}\limits_{\varepsilon}^{b}\left(\lambda h_{x}^{2}(x,T)+\lambda^{3}h^{2}(x,T)\right)e^{-2\lambda(x+\alpha T)}dx+\beta\left\|h\right\|_{H^{4}\left(\Omega\right)}^{2},\text{ }\forall h\in H_{0}^{4}\left(\Omega\right). (6.13)

By trace theorem there exists a constant C1=C1(Ω)>0C_{1}=C_{1}\left(\Omega\right)>0 depending only on the domain Ω\Omega such that

vH4(Ω)2C1v(x,T)H1(Ω)2, vH4(Ω).\left\|v\right\|_{H^{4}\left(\Omega\right)}^{2}\geq C_{1}\left\|v\left(x,T\right)\right\|_{H^{1}\left(\Omega\right)}^{2},\text{ }\forall v\in H^{4}\left(\Omega\right).

Since the regularization parameter β[2eλαT,1),\beta\in\left[2e^{-\lambda\alpha T},1\right), then we can choose λ1\lambda_{1} so large that

β2C1C1eλαT>Ce2λ(x+αT), λλ1,x[ε,b].\frac{\beta}{2}C_{1}\geq C_{1}e^{-\lambda\alpha T}>Ce^{-2\lambda(x+\alpha T)},\text{ }\forall\lambda\geq\lambda_{1},\forall x\in\left[\varepsilon,b\right].

Hence, for these values of λ,\lambda, the expression in the last line of (6.13) can be estimated from the below as:

Cεb(λhx2(x,T)+λ3h2(x,T))e2λ(x+αT)dx+βhH4(Ω)2β2hH4(Ω)2, hH04(Ω).-C\mathop{\displaystyle\int}\limits_{\varepsilon}^{b}\left(\lambda h_{x}^{2}(x,T)+\lambda^{3}h^{2}(x,T)\right)e^{-2\lambda(x+\alpha T)}dx+\beta\left\|h\right\|_{H^{4}\left(\Omega\right)}^{2}\geq\frac{\beta}{2}\left\|h\right\|_{H^{4}\left(\Omega\right)}^{2},\text{ }\forall h\in H_{0}^{4}\left(\Omega\right). (6.14)

Hence, (6.13) and (6.14) imply

Jλ,β(n)(q+h)Jλ,β(n)(q)(Jλ,β(n)(q))(h)CΩ(λ(hx2+ht2)+λ3h2)e2λφdxdt+Cεb(λhx2(x,0)+λ3h2(x,0))e2λxdx+β2hH4(Ω)2,λλ1.J_{\lambda,\beta}^{(n)}(q+h)-J_{\lambda,\beta}^{(n)}(q)-\left(J_{\lambda,\beta}^{\left(n\right)}(q)\right)^{\prime}\left(h\right)\geq C\mathop{\displaystyle\int}\limits_{\Omega}\big{(}\lambda\left(h_{x}^{2}+h_{t}^{2}\right)+\lambda^{3}h^{2}\big{)}e^{2\lambda\varphi}dxdt\\ +C\mathop{\displaystyle\int}\limits_{\varepsilon}^{b}\big{(}\lambda h_{x}^{2}(x,0)+\lambda^{3}h^{2}(x,0)\big{)}e^{-2\lambda x}dx+\frac{\beta}{2}\left\|h\right\|_{H^{4}\left(\Omega\right)}^{2},\forall\lambda\geq\lambda_{1}.

This proves (6.7). The existence and uniqueness of the minimizer qmin,nB(R,g0)¯q_{\min,n}\in\overline{B(R,g_{0})} and inequality (6.9) follow from (6.7) as well as from a combination of Lemma 2.1 and Theorem 2.1 of [1], also see [42, Chapter 10, section 3]. \square

Remark 6.2. Since the functional Jλ,β(n)(q)J_{\lambda,\beta}^{\left(n\right)}(q) is quadratic, then its strict convexity on the whole space H4(Ω)H^{4}\left(\Omega\right) follows immediately from the presence of the regularization term βqnH4(Ω)2\beta\left\|q_{n}\right\|_{H^{4}\left(\Omega\right)}^{2} in it. However, in addition to the claim of its strict convexity, we actually need in our convergence analysis below those terms in the right hand side of the strict convexity estimate (6.7), which are different from the term βhH4(Ω)2/2.\beta\left\|h\right\|_{H^{4}\left(\Omega\right)}^{2}/2. These terms are provided by Carleman estimate (6.7). The condition β[2eλαT,1)\beta\in\left[2e^{-\lambda\alpha T},1\right) of Theorem 6.1 is imposed to dominate the negative term in the last line of (6.7).

7 How to Find the Minimizer

Since we search for the minimizer qn,minq_{n,\min} of functional (4.8) on the bounded set B(R,g0)¯\overline{B(R,g_{0})} rather than on the whole space H4(Ω),H^{4}\left(\Omega\right), then we cannot just use Riesz theorem to find this minimizer, also see [1] and [42, Chapter 10, section 3] for the case of finding a minimizer of a strictly convex functional on a bounded set. Two ways of finding the minimizer qn,minB(R,g0)¯q_{n,\min}\in\overline{B(R,g_{0})} are described in this section.

7.1 Gradient descent method

Keeping in mind (3.12), choose an arbitrary function q0,nB(R,g0).q_{0,n}\in B(R,g_{0}). We arrange the gradient descent method of the minimization of functional (4.8) as follows:

qk,n=q(k1),nη(Jλ,β(n)(q(k1),n)), k=1,2,,q_{k,n}=q_{\left(k-1\right),n}-\eta\left(J_{\lambda,\beta}^{\left(n\right)}(q_{\left(k-1\right),n})\right)^{\prime},\text{ }k=1,2,..., (7.1)

where η(0,1)\eta\in\left(0,1\right) is a small number, which is chosen later. It is important to note that since functions Jλ,β(n)(q(k1),n)H04(Ω),J_{\lambda,\beta}^{\left(n\right)}(q_{\left(k-1\right),n})\in H_{0}^{4}\left(\Omega\right), then boundary conditions (4.7) are kept the same for all functions qk,n(x,t).q_{k,n}\left(x,t\right). Also, using (3.17) and (4.9), we set

ck,n(x)=1(2qk,n0(x,0))4, x[ε,b],c_{k,n}\left(x\right)=\frac{1}{\left(2q_{k,n}^{0}\left(x,0\right)\right)^{4}},\text{ }x\in\left[\varepsilon,b\right], (7.2)
cn,min(x)=1(2qn,min0(x,0))4, x[ε,b].c_{n,\min}\left(x\right)=\frac{1}{\left(2q_{n,\min}^{0}\left(x,0\right)\right)^{4}},\text{ }x\in\left[\varepsilon,b\right]. (7.3)

Theorem 7.1 claims the global convergence of the gradient descent method (7.1) in the case when qmin,nB(R/3,g0).q_{\min,n}\in B(R/3,g_{0}).

Theorem 7.1. Let the number λ1=λ1(R,Ω,α)λ0>1\lambda_{1}=\lambda_{1}\left(R,\Omega,\alpha\right)\geq\lambda_{0}>1 be the one defined in (6.6). Let λλ1\lambda\geq\lambda_{1}. For this value of λ,\lambda, let qn,minB(R,g0)¯q_{n,\min}\in\overline{B\left(R,g_{0}\right)} be the unique minimizer of the functional Jλ,β(n)(qn)J_{\lambda,\beta}^{\left(n\right)}(q_{n}) on the set B(R,g0)¯\overline{B\left(R,g_{0}\right)} with the regularization parameter β[2eλαT,1)\beta\in\left[2e^{-\lambda\alpha T},1\right) (Theorem 6.1). Assume that the function qn,minB(R/3,g0).q_{n,\min}\in B(R/3,g_{0}). For each n,n, choose the starting point of the gradient descent method (7.1) as q0,nB(R/3,g0).q_{0,n}\in B(R/3,g_{0}). Then, there exists a number η0(0,1)\eta_{0}\in\left(0,1\right) such that for any η(0,η0)\eta\in(0,\eta_{0}) all functions qk,nB(R,g0).q_{k,n}\in B(R,g_{0}). Furthermore, there exists a number θ=θ(η)(0,1)\theta=\theta\left(\eta\right)\in(0,1) such that the following convergence estimates are valid:

qk,nqn,minH4(Ω)θkq0,nqn,minH4(Ω),k=1,,\left\|q_{k,n}-q_{n,\min}\right\|_{H^{4}\left(\Omega\right)}\leq\theta^{k}\left\|q_{0,n}-q_{n,\min}\right\|_{H^{4}\left(\Omega\right)},k=1,..., (7.4)
ck,ncn,minH3(ε,b)Cθkq0,nqn,minH4(Ω), k=1,\left\|c_{k,n}-c_{n,\min}\right\|_{H^{3}\left(\varepsilon,b\right)}\leq C\theta^{k}\left\|q_{0,n}-q_{n,\min}\right\|_{H^{4}\left(\Omega\right)},\text{ }k=1,... (7.5)

Proof. Estimate (7.4) follows immediately from [33, Theorem 4.6] combined with “corrections” of functions qk,n(x,0),qn,min(x,0)q_{k,n}\left(x,0\right),q_{n,\min}\left(x,0\right) via (3.17). Estimate (7.5) follows immediately from trace theorem, (3.17) and (7.2)-(7.4). \square

7.2 Gradient projection method

Suppose now that there is no information on whether or not the function  qn,minB(R/3,g0).q_{n,\min}\in B(R/3,g_{0}). In this case we construct the gradient projection method. We introduce the function F(x,t)F\left(x,t\right) since it is easy to construct the projection operator on a ball with the center at {0}.\left\{0\right\}.

Suppose that there exists a function FH4(Ω)F\in H^{4}\left(\Omega\right) such that

F(ε,t)=g0(t+ε), Fx(ε,t)=2g0(t+ε), Fx(b,t)=0.F(\varepsilon,t)=g_{0}(t+\varepsilon),\text{ }F_{x}(\varepsilon,t)=2g_{0}^{\prime}(t+\varepsilon),\text{ }F_{x}(b,t)=0. (7.6)

This function exists, for example, in the case when the function g0(t)H5(0,T+ε).g_{0}(t)\in H^{5}\left(0,T+\varepsilon\right). Indeed, as one of examples of this function (among many), one can take, e.g.

F(x,t)=χ(x)(g0(t+ε)+2xg0(t+ε)),F\left(x,t\right)=\chi\left(x\right)\left(g_{0}(t+\varepsilon)+2xg_{0}^{\prime}(t+\varepsilon)\right),

where the function χ(x)\chi\left(x\right) is such that

χ(x)C4[ε,b],χ(x)={1,x[ε,b/4],0,x[b/2,b],between 0 and 1 for x(b/4,b/2).\chi\left(x\right)\in C^{4}\left[\varepsilon,b\right],\chi\left(x\right)=\left\{\begin{array}[]{c}1,x\in\left[\varepsilon,b/4\right],\\ 0,x\in\left[b/2,b\right],\\ \text{between }0\text{ and }1\text{ for }x\in\left(b/4,b/2\right).\end{array}\right.

The existence of such functions χ(x)\chi\left(x\right) is well known from the Analysis course. Denote

pn(x,t)=qn(x,t)F(x,t).p_{n}\left(x,t\right)=q_{n}\left(x,t\right)-F(x,t). (7.7)

Then (4.6), (4.7) become:

pnxx(x,t)pnxt(x,t)12(qn1,min0(x,0))2p_{nxx}(x,t)-p_{nxt}(x,t)\frac{1}{2\left(q_{n-1,\min}^{0}(x,0)\right)^{2}} (7.8)
+Fxx(x,t)Fxt(x,t)12(qn1,min0(x,0))2+tq(n1),min(x,t)xq(n1),min(x,0)2(qn1,min0(x,0))3=0 in Ω,+F_{xx}(x,t)-F_{xt}(x,t)\frac{1}{2\left(q_{n-1,\min}^{0}(x,0)\right)^{2}}+\partial_{t}q_{\left(n-1\right),\min}\left(x,t\right)\frac{\partial_{x}q_{\left(n-1\right),\min}(x,0)}{2\left(q_{n-1,\min}^{0}(x,0)\right)^{3}}=0\text{ in }\Omega,
pn(ε,t)=0, pnx(ε,t)=0,pnx(b,t)=0.p_{n}(\varepsilon,t)=0,\text{ }p_{nx}(\varepsilon,t)=0,p_{nx}(b,t)=0. (7.9)

Assume that

FH4(Ω)<R.\left\|F\right\|_{H^{4}\left(\Omega\right)}<R. (7.10)

By (7.7), (7.9), (7.10) and triangle inequality

pnB0(2R)={pH04(Ω):pH04(Ω)<2R}.p_{n}\in B_{0}\left(2R\right)=\left\{p\in H_{0}^{4}\left(\Omega\right):\left\|p\right\|_{H_{0}^{4}\left(\Omega\right)}<2R\right\}. (7.11)

To find the function pnB0(2R)p_{n}\in B_{0}\left(2R\right) satisfying conditions (7.8), (7.9), we minimize the following functional Iλ,β(n)(pn):B0(2R)¯I_{\lambda,\beta}^{\left(n\right)}(p_{n}):\overline{B_{0}\left(2R\right)}\rightarrow\mathbb{R}

Iλ,β(n)(pn)=Jλ,β(n)(pn+F), pnB0(2R).I_{\lambda,\beta}^{\left(n\right)}(p_{n})=J_{\lambda,\beta}^{\left(n\right)}(p_{n}+F),\text{ }p_{n}\in B_{0}\left(2R\right). (7.12)

Remark 7.1. An obvious analog of Theorem 6.1 is valid of course for functional (7.12). But in this case we should have instead of (6.6) λλ~1=λ1(2R,Ω,α)λ0>1.\lambda\geq\widetilde{\lambda}_{1}=\lambda_{1}\left(2R,\Omega,\alpha\right)\geq\lambda_{0}>1. In particular, there exists unique minimizer pn,minB0(2R)¯p_{n,\min}\in\overline{B_{0}\left(2R\right)} of this functional on the closed ball B0(2R)¯.\overline{B_{0}\left(2R\right)}. We omit the formulation of this theorem, since it is an obvious reformulation of Theorem 6.1.

Let PB0(2R)¯:H04(Ω)B0(2R)¯P_{\overline{B_{0}\left(2R\right)}}:H_{0}^{4}\left(\Omega\right)\rightarrow\overline{B_{0}\left(2R\right)} be the projection operator of the space H04(Ω)H_{0}^{4}\left(\Omega\right) on the closed ball B0(2R)¯H04(Ω).\overline{B_{0}\left(2R\right)}\subset H_{0}^{4}\left(\Omega\right). Then this operator can be easily constructed:

PB0(2R)¯(p)={p if pB0(2R)¯,2RpH4(Ω)p if pB0(2R)¯.P_{\overline{B_{0}\left(2R\right)}}\left(p\right)=\left\{\begin{array}[]{c}p\text{ if }p\in\overline{B_{0}\left(2R\right)},\\ \frac{2R}{\left\|p\right\|_{H^{4}\left(\Omega\right)}}p\text{ if }p\notin\overline{B_{0}\left(2R\right)}.\end{array}\right.

We now construct the gradient projection method of the minimization of the functional Iλ,β(n)(pn)I_{\lambda,\beta}^{\left(n\right)}(p_{n}) on the set B0(2R)¯.\overline{B_{0}\left(2R\right)}. Let p0,nB0(2R)p_{0,n}\in B_{0}\left(2R\right) be an arbitrary function. Then the sequence of the gradient projection method is:

pk,n=PB0(2R)¯(p(k1),nη(Iλ,β(n)(p(k1),n))), k=1,2,,p_{k,n}=P_{\overline{B_{0}\left(2R\right)}}\left(p_{\left(k-1\right),n}-\eta\left(I_{\lambda,\beta}^{\left(n\right)}(p_{\left(k-1\right),n})\right)^{\prime}\right),\text{ }k=1,2,..., (7.13)

where η(0,1)\eta\in\left(0,1\right) is a small number, which is chosen later. Using (7.2), (7.3) and (7.7), we set

c~k,n(x)=1(2q~k,n0(x,0))4, x[ε,b],\widetilde{c}_{k,n}\left(x\right)=\frac{1}{\left(2\widetilde{q}_{k,n}^{0}\left(x,0\right)\right)^{4}},\text{ }x\in\left[\varepsilon,b\right], (7.14)
c~n,min(x)=1(2q~n,min0(x,0))4, x[ε,b].\widetilde{c}_{n,\min}\left(x\right)=\frac{1}{\left(2\widetilde{q}_{n,\min}^{0}\left(x,0\right)\right)^{4}},\text{ }x\in\left[\varepsilon,b\right]. (7.15)

Here the function q~n,k0(x,0)\widetilde{q}_{n,k}^{0}\left(x,0\right) is obtained as follows: First, we consider the function (pn,k+F)(x,0).\left(p_{n,k}+F\right)\left(x,0\right). Next, we apply procedure (3.17) to this function. Similarly for q~n,min0(x,0).\widetilde{q}_{n,\min}^{0}\left(x,0\right).

Denote pn,minB0(2R)¯p_{n,\min}\in\overline{B_{0}\left(2R\right)} the unique minimizer of functional (7.12) on the set B0(2R)¯\overline{B_{0}\left(2R\right)} (Remark 7.1). Following (7.7), denote

q~n,min=pn,min+F, q~k,n=pk,n+F.\widetilde{q}_{n,\min}=p_{n,\min}+F,\text{ }\widetilde{q}_{k,n}=p_{k,n}+F. (7.16)

We omit the proof of Theorem 7.2 since it is very similar with the proof of Theorem 7.1. The only difference is that instead of Theorem 4.6 of [33] one should use Theorem 4.1 of [30].

Theorem 7.2. Let (7.6), (7.10) hold. Let the number λ1=λ1(R,Ω,α)λ0>1\lambda_{1}=\lambda_{1}\left(R,\Omega,\alpha\right)\geq\lambda_{0}>1 be the one defined in (6.6) and let

λλ~1=λ1(2R,Ω,α)λ1(R,Ω,α).\lambda\geq\widetilde{\lambda}_{1}=\lambda_{1}\left(2R,\Omega,\alpha\right)\geq\lambda_{1}\left(R,\Omega,\alpha\right).

For this value of λ\lambda and for β[2eλαT,1)\beta\in\left[2e^{-\lambda\alpha T},1\right) let pn,minB0(2R)¯p_{n,\min}\in\overline{B_{0}\left(2R\right)} be the unique minimizer of functional (7.12) on the set B0(2R)¯\overline{B_{0}\left(2R\right)} (Remark 7.1). Let notations (7.14) hold. For each n,n, choose the starting point of the gradient projection method (7.1) as p0,nB0(2R).p_{0,n}\in B_{0}(2R). Then there exists a number η0(0,1)\eta_{0}\in\left(0,1\right) such that for any η(0,η0)\eta\in(0,\eta_{0}) there exists a number θ=θ(η)(0,1)\theta=\theta\left(\eta\right)\in(0,1) such that the following convergence estimates are valid for the iterative process (7.13):

q~k,nq~n,minH4(Ω)θkq~0,nq~n,minH4(Ω),\left\|\widetilde{q}_{k,n}-\widetilde{q}_{n,\min}\right\|_{H^{4}\left(\Omega\right)}\leq\theta^{k}\left\|\widetilde{q}_{0,n}-\widetilde{q}_{n,\min}\right\|_{H^{4}\left(\Omega\right)},
c~k,nc~n,minH4(Ω)Cθkq~0,nq~n,minH4(Ω),\left\|\widetilde{c}_{k,n}-\widetilde{c}_{n,\min}\right\|_{H^{4}\left(\Omega\right)}\leq C\theta^{k}\left\|\widetilde{q}_{0,n}-\widetilde{q}_{n,\min}\right\|_{H^{4}\left(\Omega\right)},

where functions c~k,n,c~n,min,q~k,n,q~n,min\widetilde{c}_{k,n},\widetilde{c}_{n,\min},\widetilde{q}_{k,n},\widetilde{q}_{n,\min} are defined in (7.14)-(7.16).

Remark 7.2. Both theorems 7.1 and 7.2 claim the global convergence of corresponding minimization procedures. This is because the starting function in both cases is an arbitrary one either in B(R/3,g0)B\left(R/3,g_{0}\right) or in B0(2R)B_{0}\left(2R\right) and a smallness condition is not imposed on the number RR.

8 Contraction Mapping and Global Convergence

In this section, we prove the global convergence of the numerical method of section 4 for solving Problem 3.1. To do this, we introduce first the exact solution of our CIP. Recall that the concept of the existence of the exact solution is one of the main concepts of the theory of ill-posed problems [4, 52]. In particular, an estimate in our global convergence theorem is very similar to the one in contraction mapping.

Suppose that there exists a function c(x)c^{\ast}\left(x\right) satisfying conditions (2.1), (2.2). Let u(x,t)u^{\ast}\left(x,t\right) be the solution of problem (2.3), (2.4) with c:=c.c:=c^{\ast}. We assume that the corresponding data g0(t),tg0(t)g_{0}^{\ast}\left(t\right),\partial_{t}g_{0}^{\ast}\left(t\right) for the CIP are noiseless, see (2.11), (2.12). Let q(x,t)q^{\ast}\left(x,t\right) be the function q(x,t)q^{\ast}\left(x,t\right) which is constructed from the function u(x,t)u^{\ast}\left(x,t\right) as in (3.1). Following (3.11), we assume that

qB(R,g0),q^{\ast}\in B(R,g_{0}^{\ast}), (8.1)
B(R,g0)={qH4(Ω):qH4(Ω)<R,q(ε,t+ε)=g0(t+ε),qx(ε,t)=2(g0)(t+ε),qx(b,t)=0,12b1/4q(x,0)1/2,x[ε,b]}.B(R,g_{0}^{\ast})=\left\{\begin{array}[]{c}q\in H^{4}(\Omega):\left\|q\right\|_{H^{4}(\Omega)}<R,\\ q(\varepsilon,t+\varepsilon)=g_{0}^{\ast}(t+\varepsilon),q_{x}(\varepsilon,t)=2\left(g_{0}^{\ast}\right)^{\prime}\left(t+\varepsilon\right),\\ q_{x}(b,t)=0,\\ \frac{1}{2b^{1/4}}\leq q\left(x,0\right)\leq 1/2,x\in[\varepsilon,b]\end{array}\right\}. (8.2)

By (3.20)-(3.21)

qxx(x,t)qxt(x,t)12(q(x,0))2+qt(x,t)qx(x,0)2(q(x,0))3=0 in Ω,q_{xx}^{\ast}(x,t)-q_{xt}^{\ast}(x,t)\frac{1}{2\left(q^{\ast}(x,0)\right)^{2}}+q_{t}^{\ast}(x,t)\frac{q_{x}^{\ast}(x,0)}{2\left(q^{\ast}(x,0)\right)^{3}}=0\text{ in }\Omega, (8.3)
q(ε,t)=g0(t+ε), qx(ε,t)=2tg0(t+ε),qx(b,t)=0.q^{\ast}(\varepsilon,t)=g_{0}^{\ast}(t+\varepsilon),\text{ }q_{x}^{\ast}(\varepsilon,t)=2\partial_{t}g_{0}^{\ast}\left(t+\varepsilon\right),q_{x}^{\ast}(b,t)=0. (8.4)

By (3.3)

c(x)=1(2q(x,0))4.c^{\ast}\left(x\right)=\frac{1}{\left(2q^{\ast}\left(x,0\right)\right)^{4}}. (8.5)

It is important in the formulation of Theorem 6.1 that both functions qq and q+hq+h should have the same boundary conditions as prescribed in B(R,g0).B(R,g_{0}). However, boundary conditions for functions qnq_{n} and qq^{\ast} are different. Hence, similarly to subsection 7.2, we consider a function FH4(Ω)F^{\ast}\in H^{4}\left(\Omega\right) such that (see (7.6))

F(ε,t)=g0(t+ε), Fx(ε,t)=2tg0(t+ε), Fx(b,t)=0.F^{\ast}(\varepsilon,t)=g_{0}^{\ast}(t+\varepsilon),\text{ }F_{x}^{\ast}(\varepsilon,t)=2\partial_{t}g_{0}^{\ast}(t+\varepsilon),\text{ }F_{x}^{\ast}(b,t)=0. (8.6)

We assume similarly to (7.10) that

FH4(Ω)<R.\left\|F^{\ast}\right\|_{H^{4}\left(\Omega\right)}<R. (8.7)

Also, similarly to (7.7), we introduce the function p(x,t)p^{\ast}\left(x,t\right) as:

p(x,t)=q(x,t)F(x,t).p^{\ast}\left(x,t\right)=q^{\ast}\left(x,t\right)-F^{\ast}(x,t). (8.8)

Let λ1\lambda_{1} be the number defined in (6.6), let λλ1\lambda\geq\lambda_{1} and the function qn,minB(R,g0)¯q_{n,\min}\in\overline{B\left(R,g_{0}\right)} (see (6.8)) be the unique minimizer of the functional Jλ,β(n)(qn)J_{\lambda,\beta}^{\left(n\right)}(q_{n}) on the set B(R,g0)¯,\overline{B\left(R,g_{0}\right)}, the existence of which is established in Theorem 6.1. Following (7.7), denote

pn,min(x,t)=qn,min(x,t)F(x,t).p_{n,\min}\left(x,t\right)=q_{n,\min}\left(x,t\right)-F\left(x,t\right). (8.9)

By (6.8), (7.11), (8.1), (8.2) and (8.7)-(8.9)

pn,min,pB0(2R)¯.p_{n,\min},p^{\ast}\in\overline{B_{0}\left(2R\right)}. (8.10)

Also, it follows from embedding theorem, (3.13), (6.8), (7.10), (8.7) and (8.10) that

qC2(Ω¯),pC2(Ω¯),qn,minC2(Ω¯),pn,minC2(Ω¯)C.\left\|q^{\ast}\right\|_{C^{2}\left(\overline{\Omega}\right)},\left\|p^{\ast}\right\|_{C^{2}\left(\overline{\Omega}\right)},\left\|q_{n,\min}\right\|_{{}_{C^{2}\left(\overline{\Omega}\right)}},\left\|p_{n,\min}\right\|_{{}_{C^{2}\left(\overline{\Omega}\right)}}\leq C. (8.11)

We assume that the data g0,g0g_{0},g_{0}^{\prime} for our CIP are given with noise of the level δ,\delta, where the number δ>0\delta>0 is sufficiently small. More precisely, we assume that

FFH4(Ω)<δ.\left\|F-F^{\ast}\right\|_{H^{4}\left(\Omega\right)}<\delta. (8.12)

Observe that (3.17), (8.1) and (8.2) imply that

|qn1,min0(x,0)q(x,0)||qn1,min(x,0)q(x,0)|, x[ε,b].\left|q_{n-1,\min}^{0}\left(x,0\right)-q^{\ast}\left(x,0\right)\right|\leq\left|q_{n-1,\min}\left(x,0\right)-q^{\ast}\left(x,0\right)\right|,\text{ }x\in\left[\varepsilon,b\right]. (8.13)

By (8.3), (8.4), (8.6) and (8.8)

pxx(x,t)pxt(x,t)12(q(x,0))2+qt(x,t)qx(x,0)2(q(x,0))3 +FxxFxt(x,t)12(q(x,0))2=0p_{xx}^{\ast}(x,t)-p_{xt}^{\ast}(x,t)\frac{1}{2\left(q^{\ast}(x,0)\right)^{2}}+q_{t}^{\ast}(x,t)\frac{q_{x}^{\ast}(x,0)}{2\left(q^{\ast}(x,0)\right)^{3}}\text{ }+F_{xx}^{\ast}-F_{xt}^{\ast}(x,t)\frac{1}{2\left(q^{\ast}(x,0)\right)^{2}}=0 (8.14)

in Ω×[0,T]\Omega\times[0,T] and

p(ε,t)=px(ε,t)=px(b,t)=0.p^{\ast}(\varepsilon,t)=p_{x}^{\ast}(\varepsilon,t)=p_{x}^{\ast}(b,t)=0. (8.15)

Theorem 8.1 (contraction mapping and global convergence of the method of section 3). Let functions FF,FH4(Ω)F^{\ast}\in H^{4}\left(\Omega\right) satisfy conditions (7.6), (7.10), (8.6), (8.7) and (8.12). Let a sufficiently large number λ1=λ1(R,Ω,α)λ0>1\lambda_{1}=\lambda_{1}\left(R,\Omega,\alpha\right)\geq\lambda_{0}>1 be the one defined in (6.6). Let

λλ~1=λ1(2R,Ω,α)λ1(R,Ω,α).\lambda\geq\widetilde{\lambda}_{1}=\lambda_{1}\left(2R,\Omega,\alpha\right)\geq\lambda_{1}\left(R,\Omega,\alpha\right). (8.16)

For this value of λ,\lambda, let qn,minB(R,g0)¯q_{n,\min}\in\overline{B\left(R,g_{0}\right)} be the unique minimizer of the functional Jλ,β(n)(qn)J_{\lambda,\beta}^{\left(n\right)}(q_{n}) on the set B(R,g0)¯\overline{B\left(R,g_{0}\right)} with the regularization parameter β[2eλαT,1)\beta\in\left[2e^{-\lambda\alpha T},1\right) (Theorem 6.1). Let

q¯n=qn,minq, c¯n=cn,minc, \overline{q}_{n}=q_{n,\min}-q^{\ast},\text{ }\overline{c}_{n}=c_{n,\min}-c^{\ast},\emph{\ } (8.17)

where cn,minc_{n,\min} is defined in (7.3). Then the following convergence estimate holds

Ω(q¯nx2+q¯nt2+q¯n2)(x,t)e2λφdxdt+εb(q¯nx2+q¯n2)(x,0)e2λxdx\mathop{\displaystyle\int}\limits_{\Omega}\left(\overline{q}_{nx}^{2}+\overline{q}_{nt}^{2}+\overline{q}_{n}^{2}\right)\left(x,t\right)e^{2\lambda\varphi}dxdt+\mathop{\displaystyle\int}\limits_{\varepsilon}^{b}\left(\overline{q}_{nx}^{2}+\overline{q}_{n}^{2}\right)\left(x,0\right)e^{-2\lambda x}dx
CλΩ(q¯(n1)x2+q¯(n1)t2+q¯n12)(x,t)e2λφdxdt\leq\frac{C}{\lambda}\mathop{\displaystyle\int}\limits_{\Omega}\left(\overline{q}_{\left(n-1\right)x}^{2}+\overline{q}_{\left(n-1\right)t}^{2}+\overline{q}_{n-1}^{2}\right)\left(x,t\right)e^{2\lambda\varphi}dxdt (8.18)
+Cλεb(q¯(n1)x2(x,0)+q¯n12)(x,0)e2λxdx+C(δ2+β),+\frac{C}{\lambda}\mathop{\displaystyle\int}\limits_{\varepsilon}^{b}\left(\overline{q}_{\left(n-1\right)x}^{2}\left(x,0\right)+\overline{q}_{n-1}^{2}\right)\left(x,0\right)e^{-2\lambda x}dx+C\left(\delta^{2}+\beta\right),

which leads to:

Ω(q¯nx2+q¯nt2+q¯n2)(x,t)e2λφdxdt\mathop{\displaystyle\int}\limits_{\Omega}\left(\overline{q}_{nx}^{2}+\overline{q}_{nt}^{2}+\overline{q}_{n}^{2}\right)\left(x,t\right)e^{2\lambda\varphi}dxdt (8.19)
CnλnΩ(q¯0x2+q¯0t2+q¯02)(x,t)e2λφdxdt+C(δ2+β).\leq\frac{C^{n}}{\lambda^{n}}\mathop{\displaystyle\int}\limits_{\Omega}\left(\overline{q}_{0x}^{2}+\overline{q}_{0t}^{2}+\overline{q}_{0}^{2}\right)\left(x,t\right)e^{2\lambda\varphi}dxdt+C\left(\delta^{2}+\beta\right).

In addition,

c¯nH1(ε,b)2CnλnΩ(q¯0x2+q¯0t2+q¯02)(x,t)e2λφdxdt+C(δ2+β).\left\|\overline{c}_{n}\right\|_{H^{1}\left(\varepsilon,b\right)}^{2}\leq\frac{C^{n}}{\lambda^{n}}\mathop{\displaystyle\int}\limits_{\Omega}\left(\overline{q}_{0x}^{2}+\overline{q}_{0t}^{2}+\overline{q}_{0}^{2}\right)\left(x,t\right)e^{2\lambda\varphi}dxdt+C\left(\delta^{2}+\beta\right). (8.20)

Remarks 8.1:

1. Due to the presence of the term C/λC/\lambda with a sufficiently large λ\lambda, estimate (8.18) is similar to the one in the classic contraction mapping principle, although we do not claim here the existence of the fixed point. This explains the title of our paper.

2. When computing the unique minimizer q0,minq_{0,\min} of functional (4.4) on the set B(R,g0)¯,\overline{B(R,g_{0})}, we do not impose a smallness condition on the number RR. Therefore, Theorem 8.1 claims the global convergence of the method of section 4, see Introduction for our definition of the global convergence. The same is true for Theorems 9.1 and 9.2 in the next section.

Proof Theorem 8.1. Denote hn=ppn,min.h_{n}=p^{\ast}-p_{n,\min}. By (8.9) hn=q¯n+(FF).h_{n}=-\overline{q}_{n}+\left(F-F^{\ast}\right). Hence, (8.12) and embedding theorem imply:

hn212q¯n2Cδ2,hnx212q¯nx2Cδ2,hnt212q¯nt2Cδ2 in Ω¯,h_{n}^{2}\geq\frac{1}{2}\overline{q}_{n}^{2}-C\delta^{2},h_{nx}^{2}\geq\frac{1}{2}\overline{q}_{nx}^{2}-C\delta^{2},h_{nt}^{2}\geq\frac{1}{2}\overline{q}_{nt}^{2}-C\delta^{2}\text{ in }\overline{\Omega}, (8.21)
hn2+hnx2+hnt2C(q¯n2+q¯nx2+q¯nt2+δ2) in Ω¯.h_{n}^{2}+h_{nx}^{2}+h_{nt}^{2}\leq C\left(\overline{q}_{n}^{2}+\overline{q}_{nx}^{2}+\overline{q}_{nt}^{2}+\delta^{2}\right)\text{ in }\overline{\Omega}. (8.22)

Consider the functional Iλ,β(n)(p)=Jλ,β(n)(p+F).I_{\lambda,\beta}^{\left(n\right)}(p^{\ast})=J_{\lambda,\beta}^{\left(n\right)}(p^{\ast}+F). Since both functions pn,minp_{n,\min} and pp^{\ast} have the same zero boundary conditions (7.9) and since by (8.10) both of them belong to the set B0(2R)¯\overline{B_{0}\left(2R\right)}, then the analog of Theorem 6.1, which is mentioned in Remark 7.1, implies (see (6.7))

Iλ,β(n)(p)Iλ,β(n)(pn,min)(Iλ,β(n)(pn,min))(hn)I_{\lambda,\beta}^{\left(n\right)}(p^{\ast})-I_{\lambda,\beta}^{\left(n\right)}(p_{n,\min})-\left(I_{\lambda,\beta}^{\left(n\right)}(p_{n,\min})\right)^{\prime}(h_{n})
CΩ[λ(hnx2+hnt2)+λ3hn2]e2λφdxdt\geq C\mathop{\displaystyle\int}\limits_{\Omega}\Big{[}\lambda\left(h_{nx}^{2}+h_{nt}^{2}\right)+\lambda^{3}h_{n}^{2}\Big{]}e^{2\lambda\varphi}dxdt (8.23)
+Cεb(λhnx2(x,0)+λ3hn2(x,0))e2λxdx+β2hnH4(Ω)2, λλ~1.+C\mathop{\displaystyle\int}\limits_{\varepsilon}^{b}\big{(}\lambda h_{nx}^{2}(x,0)+\lambda^{3}h_{n}^{2}(x,0)\big{)}e^{-2\lambda x}dx+\frac{\beta}{2}\left\|h_{n}\right\|_{H^{4}\left(\Omega\right)}^{2},\text{ }\forall\lambda\geq\widetilde{\lambda}_{1}.

By (6.9)

(Iλ,β(n)(pn,min))(hn)0.-\left(I_{\lambda,\beta}^{\left(n\right)}(p_{n,\min})\right)^{\prime}(h_{n})\leq 0.

Hence, the left hand side of (8.23) can be estimated as:

Iλ,β(n)(p)Iλ,β(n)(pn,min)(Iλ,β(n)(pn,min))(h)Iλ,β(n)(p).I_{\lambda,\beta}^{\left(n\right)}(p^{\ast})-I_{\lambda,\beta}^{\left(n\right)}(p_{n,\min})-\left(I_{\lambda,\beta}^{\left(n\right)}(p_{n,\min})\right)^{\prime}(h)\leq I_{\lambda,\beta}^{\left(n\right)}(p^{\ast}). (8.24)

We now estimate the right hand side of (8.24) from the above. It follows from (4.8), (7.12) and (8.3) that

Iλ,β(n)(p)=ΩGn2e2λφdxdt+βp+FH4(Ω)2,I_{\lambda,\beta}^{(n)}(p^{\ast})=\mathop{\displaystyle\int}\limits_{\Omega}G_{n}^{2}e^{2\lambda\varphi}dxdt+\beta\left\|p^{\ast}+F\right\|_{H^{4}\left(\Omega\right)}^{2}, (8.25)

where

Gn=pxx(x,t)pxt(x,t)12(q(n1),min0(x,0))2+tq(n1),min(x,t)xq(n1),min(x,0)2(q(n1),min0(x,0))3G_{n}=p_{xx}^{\ast}(x,t)-p_{xt}^{\ast}(x,t)\frac{1}{2\left(q_{\left(n-1\right),\min}^{0}(x,0)\right)^{2}}+\partial_{t}q_{\left(n-1\right),\min}\left(x,t\right)\frac{\partial_{x}q_{\left(n-1\right),\min}(x,0)}{2\left(q_{\left(n-1\right),\min}^{0}(x,0)\right)^{3}}
+FxxFxt(x,t)12(q(n1),min0(x,0))2+F_{xx}-F_{xt}(x,t)\frac{1}{2\left(q_{\left(n-1\right),\min}^{0}(x,0)\right)^{2}}
=qxx(x,t)qxt(x,t)12(q(x,0))2+qt(x,t)qx(x,0)2(q(x,0))3=q_{xx}^{\ast}(x,t)-q_{xt}^{\ast}(x,t)\frac{1}{2\left(q^{\ast}(x,0)\right)^{2}}+q_{t}^{\ast}(x,t)\frac{q_{x}^{\ast}(x,0)}{2\left(q^{\ast}(x,0)\right)^{3}} (8.26)
+(FF)xx(FF)xt12(q(n1),min0(x,0))2+\left(F-F^{\ast}\right)_{xx}-\left(F-F^{\ast}\right)_{xt}\frac{1}{2\left(q_{\left(n-1\right),\min}^{0}(x,0)\right)^{2}}
qxt(12(q(n1),min0(x,0))212(q(x,0))2)-q_{xt}^{\ast}\left(\frac{1}{2\left(q_{\left(n-1\right),\min}^{0}(x,0)\right)^{2}}-\frac{1}{2\left(q^{\ast}(x,0)\right)^{2}}\right)
+(tq(n1),min(x,t)xq(n1),min(x,0)2(q(n1),min0(x,0))3qt(x,t)qx(x,0)2(q(x,0))3).+\left(\partial_{t}q_{\left(n-1\right),\min}\left(x,t\right)\frac{\partial_{x}q_{\left(n-1\right),\min}(x,0)}{2\left(q_{\left(n-1\right),\min}^{0}(x,0)\right)^{3}}-q_{t}^{\ast}(x,t)\frac{q_{x}^{\ast}(x,0)}{2\left(q^{\ast}(x,0)\right)^{3}}\right).

By (8.3), the third line of (8.26) equals zero. Hence, (8.26) becomes

Gn=(FF)xx+(FF)xt12(q(n1),min0(x,0))2G_{n}=\left(F-F^{\ast}\right)_{xx}+\left(F-F^{\ast}\right)_{xt}\frac{1}{2\left(q_{\left(n-1\right),\min}^{0}(x,0)\right)^{2}}
qxt(12(q(n1),min0(x,0))212(q(x,0))2)-q_{xt}^{\ast}\left(\frac{1}{2\left(q_{\left(n-1\right),\min}^{0}(x,0)\right)^{2}}-\frac{1}{2\left(q^{\ast}(x,0)\right)^{2}}\right)
+(tq(n1),min(x,t)xq(n1),min(x,0)2(q(n1),min0(x,0))3qt(x,t)qx(x,0)2(q(x,0))3).+\left(\partial_{t}q_{\left(n-1\right),\min}\left(x,t\right)\frac{\partial_{x}q_{\left(n-1\right),\min}(x,0)}{2\left(q_{\left(n-1\right),\min}^{0}(x,0)\right)^{3}}-q_{t}^{\ast}(x,t)\frac{q_{x}^{\ast}(x,0)}{2\left(q^{\ast}(x,0)\right)^{3}}\right).

Hence, by (3.17), (8.12) and embedding theorem

|Gn|Cδ+C|12(q(n1),min0(x,0))212(q(x,0))2|\left|G_{n}\right|\leq C\delta+C\left|\frac{1}{2\left(q_{\left(n-1\right),\min}^{0}(x,0)\right)^{2}}-\frac{1}{2\left(q^{\ast}(x,0)\right)^{2}}\right|
+|tq(n1),min(x,t)xq(n1),min(x,0)2(q(n1),min0(x,0))3qt(x,t)qx(x,0)2(q(x,0))3|.+\left|\partial_{t}q_{\left(n-1\right),\min}\left(x,t\right)\frac{\partial_{x}q_{\left(n-1\right),\min}(x,0)}{2\left(q_{\left(n-1\right),\min}^{0}(x,0)\right)^{3}}-q_{t}^{\ast}(x,t)\frac{q_{x}^{\ast}(x,0)}{2\left(q^{\ast}(x,0)\right)^{3}}\right|. (8.27)

Using (8.12) and (8.13), we obtain

|12(q(n1),min0(x,0))212(q(x,0))2|\left|\frac{1}{2\left(q_{\left(n-1\right),\min}^{0}(x,0)\right)^{2}}-\frac{1}{2\left(q^{\ast}(x,0)\right)^{2}}\right|
=|(q(n1),min0(x,0)q(x,0))(FF)(x,0)||q(n1),min0(x,0)+q(x,0)|2(q(n1),min0(x,0))2(q(x,0))2.=\frac{\left|\left(q_{\left(n-1\right),\min}^{0}\left(x,0\right)-q^{\ast}(x,0)\right)-\left(F-F^{\ast}\right)\left(x,0\right)\right|\left|q_{\left(n-1\right),\min}^{0}(x,0)+q^{\ast}(x,0)\right|}{2\left(q_{\left(n-1\right),\min}^{0}(x,0)\right)^{2}\left(q^{\ast}(x,0)\right)^{2}.}
Cδ+C|hn1(x,0)|.\leq C\delta+C\left|h_{n-1}\left(x,0\right)\right|.

Combining this with (8.27), we obtain

|Gn|Cδ+C|hn1(x,0)|\left|G_{n}\right|\leq C\delta+C\left|h_{n-1}\left(x,0\right)\right|
+|tq(n1),min(x,t)xq(n1),min(x,0)2(q(n1),min0(x,0))3qt(x,t)qx(x,0)2(q(x,0))3|.+\left|\partial_{t}q_{\left(n-1\right),\min}\left(x,t\right)\frac{\partial_{x}q_{\left(n-1\right),\min}(x,0)}{2\left(q_{\left(n-1\right),\min}^{0}(x,0)\right)^{3}}-q_{t}^{\ast}(x,t)\frac{q_{x}^{\ast}(x,0)}{2\left(q^{\ast}(x,0)\right)^{3}}\right|. (8.28)

Next,

12(q(n1),min0(x,0))3=12(q(x,0))3\frac{1}{2\left(q_{\left(n-1\right),\min}^{0}(x,0)\right)^{3}}=\frac{1}{2\left(q^{\ast}(x,0)\right)^{3}}
+(12(q(n1),min0(x,0))312(q(x,0))3)+\left(\frac{1}{2\left(q_{\left(n-1\right),\min}^{0}\left(x,0\right)\right)^{3}}-\frac{1}{2\left(q^{\ast}(x,0)\right)^{3}}\right)
=12(q(x,0))3+Sn1(x,0)[(q(n1),min0(x,0)q(x,0))(FF)(x,0)],=\frac{1}{2\left(q^{\ast}(x,0)\right)^{3}}+S_{n-1}\left(x,0\right)\left[\left(q_{\left(n-1\right),\min}^{0}\left(x,0\right)-q^{\ast}(x,0)\right)-\left(F-F^{\ast}\right)\left(x,0\right)\right],

where the function Sn1(x,0)S_{n-1}\left(x,0\right) can be estimated as

|Sn1(x,0)|C.\left|S_{n-1}\left(x,0\right)\right|\leq C. (8.29)

Hence,

tq(n1),min(x,t)xq(n1),min(x,0)2(q(n1),min0(x,0))3=\partial_{t}q_{\left(n-1\right),\min}\left(x,t\right)\frac{\partial_{x}q_{\left(n-1\right),\min}(x,0)}{2\left(q_{\left(n-1\right),\min}^{0}(x,0)\right)^{3}}=
tq(n1),min(x,t)xq(n1),min(x,0)2(q(x,0))3+\frac{\partial_{t}q_{\left(n-1\right),\min}\left(x,t\right)\partial_{x}q_{\left(n-1\right),\min}(x,0)}{2\left(q^{\ast}(x,0)\right)^{3}}+
+[tq(n1),min(x,t)xq(n1),min(x,0)]×+\left[\partial_{t}q_{\left(n-1\right),\min}\left(x,t\right)\partial_{x}q_{\left(n-1\right),\min}(x,0)\right]\times
×Sn1(x,0)[(q(n1),min0(x,0)q(x,0))(FF)(x,0)].\times S_{n-1}\left(x,0\right)\left[\left(q_{\left(n-1\right),\min}^{0}\left(x,0\right)-q^{\ast}(x,0)\right)-\left(F-F^{\ast}\right)\left(x,0\right)\right].

Hence, using (3.11), (3.17), (8.1), (8.13) and (8.29), we obtain

|tq(n1),min(x,t)xq(n1),min(x,0)2(q(n1),min0(x,0))3qt(x,t)qx(x,0)2(q(x,0))3|\left|\partial_{t}q_{\left(n-1\right),\min}\left(x,t\right)\frac{\partial_{x}q_{\left(n-1\right),\min}(x,0)}{2\left(q_{\left(n-1\right),\min}^{0}(x,0)\right)^{3}}-q_{t}^{\ast}(x,t)\frac{q_{x}^{\ast}(x,0)}{2\left(q^{\ast}(x,0)\right)^{3}}\right|
12(q(x,0))3|tq(n1),min(x,t)xq(n1),min(x,0)qt(x,t)qx(x,0)|\leq\frac{1}{2\left(q^{\ast}(x,0)\right)^{3}}\left|\partial_{t}q_{\left(n-1\right),\min}\left(x,t\right)\partial_{x}q_{\left(n-1\right),\min}(x,0)-q_{t}^{\ast}(x,t)q_{x}^{\ast}(x,0)\right| (8.30)
+C(δ+|hn1(x,0)|)+C\left(\delta+\left|h_{n-1}\left(x,0\right)\right|\right)
C(δ+|hn1(x,0)|+|xhn1,x(x,0)|+|hn1,t(x,t)|).\leq C\left(\delta+\left|h_{n-1}\left(x,0\right)\right|+\left|\partial_{x}h_{n-1,x}\left(x,0\right)\right|+\left|h_{n-1,t}\left(x,t\right)\right|\right).

Combining (8.28) with (8.30), we obtain

|Gn(x,t)|C(δ+|hn1(x,0)|+|hn1,x(x,0)|+|hn1,t(x,t)|).\left|G_{n}\left(x,t\right)\right|\leq C\left(\delta+\left|h_{n-1}\left(x,0\right)\right|+\left|h_{n-1,x}\left(x,0\right)\right|+\left|h_{n-1,t}\left(x,t\right)\right|\right).

Hence, by (8.25)

Iλ,β(n)(p)CΩ(δ2+hn12(x,0)+hn1,x2(x,0)+hn1,t2(x,t))e2λφdxdt+Cβ.I_{\lambda,\beta}^{(n)}(p^{\ast})\leq C\mathop{\displaystyle\int}\limits_{\Omega}\left(\delta^{2}+h_{n-1}^{2}\left(x,0\right)+h_{n-1,x}^{2}\left(x,0\right)+h_{n-1,t}^{2}\left(x,t\right)\right)e^{2\lambda\varphi}dxdt+C\beta.

Substituting this in (8.24) and then using (8.23), we obtain

Ω(hnx2+hnt2+hn2)e2λφdxdt+εb(hnx2(x,0)+hn2(x,0))e2λxdx\mathop{\displaystyle\int}\limits_{\Omega}\left(h_{nx}^{2}+h_{nt}^{2}+h_{n}^{2}\right)e^{2\lambda\varphi}dxdt+\mathop{\displaystyle\int}\limits_{\varepsilon}^{b}\left(h_{nx}^{2}(x,0)+h_{n}^{2}(x,0)\right)e^{-2\lambda x}dx
CλΩ(δ2+hn12(x,0)+hn1,x2(x,0)+hn1,t2(x,t))e2λφdxdt+Cβ.\leq\frac{C}{\lambda}\mathop{\displaystyle\int}\limits_{\Omega}\left(\delta^{2}+h_{n-1}^{2}\left(x,0\right)+h_{n-1,x}^{2}\left(x,0\right)+h_{n-1,t}^{2}\left(x,t\right)\right)e^{2\lambda\varphi}dxdt+C\beta. (8.31)

Obviously

Ω(hn12(x,0)+hn1,x2(x,0))e2λφdxdt12λαεb(h(n1)x2(x,0)+hn12(x,0))e2λxdx,\mathop{\displaystyle\int}\limits_{\Omega}\left(h_{n-1}^{2}\left(x,0\right)+h_{n-1,x}^{2}\left(x,0\right)\right)e^{2\lambda\varphi}dxdt\leq\frac{1}{2\lambda\alpha}\mathop{\displaystyle\int}\limits_{\varepsilon}^{b}\left(h_{\left(n-1\right)x}^{2}(x,0)+h_{n-1}^{2}(x,0)\right)e^{-2\lambda x}dx, (8.32)
Ωδ2e2λφdxdtCδ2λ2.\mathop{\displaystyle\int}\limits_{\Omega}\delta^{2}e^{2\lambda\varphi}dxdt\leq C\frac{\delta^{2}}{\lambda^{2}}. (8.33)

Denote

yn=Ω(hnx2+hnt2+hn2)e2λφdxdt+εb(hnx2(x,0)+hn2(x,0))e2λxdx.y_{n}=\mathop{\displaystyle\int}\limits_{\Omega}\left(h_{nx}^{2}+h_{nt}^{2}+h_{n}^{2}\right)e^{2\lambda\varphi}dxdt+\mathop{\displaystyle\int}\limits_{\varepsilon}^{b}\left(h_{nx}^{2}(x,0)+h_{n}^{2}(x,0)\right)e^{-2\lambda x}dx. (8.34)

Then (8.31)-(8.33) imply

ynCλyn1+C(δ2λ2+β).y_{n}\leq\frac{C}{\lambda}y_{n-1}+C\left(\frac{\delta^{2}}{\lambda^{2}}+\beta\right). (8.35)

Iterating (8.35) with respect to n,n, we obtain

Ω(hnx2+hnt2+hn2)e2λφdxdt+εb(hnx2(x,0)+hn2(x,0))e2λxdx\mathop{\displaystyle\int}\limits_{\Omega}\left(h_{nx}^{2}+h_{nt}^{2}+h_{n}^{2}\right)e^{2\lambda\varphi}dxdt+\mathop{\displaystyle\int}\limits_{\varepsilon}^{b}\left(h_{nx}^{2}(x,0)+h_{n}^{2}(x,0)\right)e^{-2\lambda x}dx
Cλn[Ω(h0x2+h0t2+h02)e2λφdxdt+εb(h0x2(x,0)+h02(x,0))e2λxdx]\leq\frac{C}{\lambda^{n}}\left[\mathop{\displaystyle\int}\limits_{\Omega}\left(h_{0x}^{2}+h_{0t}^{2}+h_{0}^{2}\right)e^{2\lambda\varphi}dxdt+\mathop{\displaystyle\int}\limits_{\varepsilon}^{b}\left(h_{0x}^{2}(x,0)+h_{0}^{2}(x,0)\right)e^{-2\lambda x}dx\right] (8.36)
+C(δ2λ2+β).+C\left(\frac{\delta^{2}}{\lambda^{2}}+\beta\right).

Apply (8.21) to the left hand side of estimate (8.36). Also, apply (8.22) at n=0n=0 to the right hand side of (8.36). We obtain (8.19). Estimate (8.20) follows from an obvious combination of (8.19) with (7.3), (8.5), (8.13) and (LABEL:8.150). Finally, estimate (8.18) follows immediately from (8.21), (8.22) and (8.31). \square

9 Global Convergence of the Gradient and Gradient Projection Methods to the Exact Solution

First, we consider the gradient method of the minimization of functionals Jλ,β(n)(q(k1),n)J_{\lambda,\beta}^{\left(n\right)}(q_{\left(k-1\right),n}) on the set B(R,g0)¯,\overline{B(R,g_{0})}, see (7.1). The proof of Theorem 9.1 follows immediately from the triangle inequality combined with Theorems 7.1 and 8.1.

Theorem 9.1. Let α0\alpha_{0} and λ0\lambda_{0} be the numbers of Theorem 5.1. Let the sufficiently large number λ1=λ1(R,Ω,α)λ0>1\lambda_{1}=\lambda_{1}\left(R,\Omega,\alpha\right)\geq\lambda_{0}>1 be the one defined in (6.6). Let the number λ~1\widetilde{\lambda}_{1} be the same as in (8.16),

λ~1=λ1(2R,Ω,α)λ1(R,Ω,α).\widetilde{\lambda}_{1}=\lambda_{1}\left(2R,\Omega,\alpha\right)\geq\lambda_{1}\left(R,\Omega,\alpha\right).

Let λλ1\lambda\geq\lambda_{1} and let the regularization parameter β[2eλαT,1).\beta\in\left[2e^{-\lambda\alpha T},1\right). Assume that the functions qmin,nB(R/3,g0)q_{\min,n}\in B(R/3,g_{0}) for all n.n. For each nn, choose the starting point of the gradient method (7.1) as q0,nB(R/3,g0).q_{0,n}\in B(R/3,g_{0}). Then there exists a number η0(0,1)\eta_{0}\in\left(0,1\right) such that for any η(0,η0)\eta\in(0,\eta_{0}) all functions qk,nB(R,g0),k=1,;q_{k,n}\in B(R,g_{0}),k=1,...; n=1,.n=1,.... Furthermore, there exists a number θ=θ(η)(0,1)\theta=\theta\left(\eta\right)\in(0,1) such that the following convergence estimate is valid:

ck,ncH1(ε,b)Cθkq0,nqn,minH4(Ω)\left\|c_{k,n}-c^{\ast}\right\|_{H^{1}\left(\varepsilon,b\right)}\leq C\theta^{k}\left\|q_{0,n}-q_{n,\min}\right\|_{H^{4}\left(\Omega\right)}
+Cn/2λn/2(Ω(q¯0x2+q¯0t2+q¯02)(x,t)e2λφdxdt)1/2+C(δ+β),+\frac{C^{n/2}}{\lambda^{n/2}}\left(\mathop{\displaystyle\int}\limits_{\Omega}\left(\overline{q}_{0x}^{2}+\overline{q}_{0t}^{2}+\overline{q}_{0}^{2}\right)\left(x,t\right)e^{2\lambda\varphi}dxdt\right)^{1/2}+C\left(\delta+\sqrt{\beta}\right),

where the function q¯0\overline{q}_{0} is defined in (8.17).

Consider now the gradient projection method of the minimization of the functionals Iλ,β(n)(pn)=Jλ,β(n)(pn+F)I_{\lambda,\beta}^{\left(n\right)}(p_{n})=J_{\lambda,\beta}^{\left(n\right)}(p_{n}+F) in (7.12) on the set B0(2R)¯\overline{B_{0}\left(2R\right)}, see (7.13). We use notations (7.14). Theorem 9.2 follows immediately from the triangle inequality combined with Theorems 7.2 and 8.1.

Theorem 9.2. Let the number λ1=λ1(R,Ω,α)λ0>1\lambda_{1}=\lambda_{1}\left(R,\Omega,\alpha\right)\geq\lambda_{0}>1 be the one defined in (6.6). Let the number λ~1\widetilde{\lambda}_{1} be the same as in (8.16),

λ~1=λ1(2R,Ω,α)λ1(R,Ω,α).\widetilde{\lambda}_{1}=\lambda_{1}\left(2R,\Omega,\alpha\right)\geq\lambda_{1}\left(R,\Omega,\alpha\right).

Let λλ1\lambda\geq\lambda_{1} and let the regularization parameter β[2eλαT,1).\beta\in\left[2e^{-\lambda\alpha T},1\right). Consider the gradient projection method (7.13). For each nn, choose the starting point p0,np_{0,n} of this method as an arbitrary point of the ball B0(2R).B_{0}(2R). Then there exists a number η0(0,1)\eta_{0}\in\left(0,1\right) such that for any η(0,η0)\eta\in(0,\eta_{0}) there exists a number θ=θ(η)(0,1)\theta=\theta\left(\eta\right)\in(0,1) such that the following convergence estimate is valid:

c~k,ncH1(ε,b)Cθkq~0,nq~n,minH4(Ω)\left\|\widetilde{c}_{k,n}-c^{\ast}\right\|_{H^{1}\left(\varepsilon,b\right)}\leq C\theta^{k}\left\|\widetilde{q}_{0,n}-\widetilde{q}_{n,\min}\right\|_{H^{4}\left(\Omega\right)}
+Cn/2λn/2(Ω(q¯0x2+q¯0t2+q¯02)(x,t)e2λφdxdt)1/2+C(δ+β),+\frac{C^{n/2}}{\lambda^{n/2}}\left(\mathop{\displaystyle\int}\limits_{\Omega}\left(\overline{q}_{0x}^{2}+\overline{q}_{0t}^{2}+\overline{q}_{0}^{2}\right)\left(x,t\right)e^{2\lambda\varphi}dxdt\right)^{1/2}+C\left(\delta+\sqrt{\beta}\right),

where functions c~k,n,q~0,n,q~n,min\widetilde{c}_{k,n},\widetilde{q}_{0,n},\widetilde{q}_{n,\min} and q¯0\overline{q}_{0} are defined in (7.14), (7.16) and (8.17) respectively.

10 Numerical Studies

10.1 Numerical implementation

To generate the simulated data, by Lemma 2.2, we solve problem (2.3), (2.4) for the case when the whole real line is replaced by a large interval (a,a)(-a,a) with the absorbing boundary conditions (2.9)–(2.10). More precisely, just as in Section 6.1 of [30], we use the implicit scheme to numerically solve

{c(x)utt(x,t)=uxx(x,t)(x,t)(a,a)×(0,T),u(a,t)ux(a,t)=0t(0,T),u(a,t)+ux(a,t)=0t(0,T),u(x,0)=0x,ut(x,0)=δ~(x)x,\left\{\begin{array}[]{rcll}c(x)u_{tt}(x,t)&=&u_{xx}(x,t)&(x,t)\in(-a,a)\times(0,T),\\ u(-a,t)-u_{x}(-a,t)&=&0&t\in(0,T),\\ u(a,t)+u_{x}(a,t)&=&0&t\in(0,T),\\ u(x,0)&=&0&x\in\mathbb{R},\\ u_{t}(x,0)&=&\widetilde{\delta}(x)&x\in\mathbb{R},\end{array}\right. (10.1)

where a=5a=5, T=6T=6 and

δ~(x)=302πe(30x)22\widetilde{\delta}(x)=\frac{30}{\sqrt{2\pi}}e^{-\frac{(30x)^{2}}{2}}

is a smooth approximation of the Dirac function δ(x)\delta\left(x\right). We solve problem (10.1) by the implicit finite difference method. In the finite difference scheme, we arrange a uniform partition for the interval [a,a][-a,a] as {y0=a,y1,,yN=a}[a,a]\{y_{0}=-a,y_{1},\dots,y_{N}=a\}\subset[-a,a] with yi=a+2ia/Nxy_{i}=a+2ia/N_{x}, i=0,,Nxi=0,\dots,N_{x}, where NxN_{x} is a large number. In the time domain, we split the interval [0,T][0,T] into Nt+1N_{t}+1 uniform sub-intervals [tj,tj+1][t_{j},t_{j+1}], j=0,,Ntj=0,\dots,N_{t}, with tj=jT/Nt,t_{j}=jT/N_{t}, where NtN_{t} is a large number. In our computational setting, Nx=3001N_{x}=3001 and Nt=301N_{t}=301. These numbers are the same as in [30].

We observe a computational error for the function uu near (x=0,t=0)(x=0,t=0). This is due to the fact that the function δ~(x)\widetilde{\delta}\left(x\right) is not exactly equal to the Dirac function. We correct the error as follows. It follows from (2.2), (2.7) and (2.8) that u(x,t)=1/2u(x,t)=1/2 in a neighborhood of the point (x,t)=(0,0).\left(x,t\right)=\left(0,0\right). We, therefore, replace the data u(ϵ,t)u(\epsilon,t) by 1/21/2 when |t||t| is small. In our computation, we set u(x,t)=1/2u(x,t)=1/2 for (x,t)[0,0.0067]×[0,0.26](x,t)\in[0,0.0067]\times[0,0.26]. This data correction step is exactly the same as in Section 6.1 of [30] and is illustrated by Figure 2 in that publication. Then, we can extract the noiseless data g0g_{0}^{\ast} easily. We next add the noise into the data via the formula

g0=g0(1+δrand)g_{0}=g_{0}^{\ast}(1+\delta\mbox{rand}) (10.2)

where δ\delta is the noise level and rand is the function that generates uniformly distributed random numbers in the range [1,1].[-1,1]. In all numerical tests with simulated data below, the noise level δ=0.05,\delta=0.05, i.e. 5%.5\%. Due to (2.12), the function g1=g0.g_{1}=g_{0}^{\prime}. Due to the presence of noise, see (10.2), we cannot compute g1=g0g_{1}=g_{0}^{\prime} by the finite difference method. Hence, the function g0g_{0}^{\prime} is computed by the Tikhonov regularization method. The version of the Tikhonov regularization method for this problem is well-known. Hence, we do not describe this step here.

Having the data for the function qq in hand, we proceed as in Algorithm 1.

Algorithm 1 A numerical method to solve Problem 3.1
1:   Choose a set of parameters λ\lambda, α\alpha and β\beta.
2:   Compute the function q0q_{0} by minimizing the functional Jλ,β(0)J^{(0)}_{\lambda,\beta} defined in (4.4). Due to (3.3), the initial reconstruction is given by
cinit(x)=1(2q0(x,0))4for all x[ϵ,b].c_{\mathrm{i}nit}(x)=\frac{1}{(2q_{0}(x,0))^{4}}\quad\mbox{for all }x\in[\epsilon,b].
3:   Assume that the function qn1q_{n-1} is known. We compute the function qnq_{n} by minimizing the function Jλ,β(n)J^{(n)}_{\lambda,\beta} defined in (4.8).
4:  Set qcomp=qnq_{\mathrm{c}omp}=q_{n} when n=nn=n^{*} is large enough.
5:   Due to (3.3), the function ccompc_{\mathrm{c}omp} is set to be
ccomp(x)=1(2qcomp(x,0))4for all x[ϵ,b].c_{\mathrm{c}omp}(x)=\frac{1}{(2q_{\mathrm{c}omp}(x,0))^{4}}\quad\mbox{for all }x\in[\epsilon,b].

In step 1 of Algorithm 1, we choose λ=2\lambda=2, α=0.3\alpha=0.3 and β=1011.\beta=10^{-11}. These parameters were chosen by a trial-error process that is similar to the one in [50]. Just as in [50], we choose a reference numerical test in which we know the true solution. In fact, test 1 of section 10.2 was our reference test. We tried several values of λ,\lambda, α,\alpha, and β\beta until the numerical solution to that reference test became acceptable. Then, we have used the same values of these parameters for all other tests, including all five (5) available cases of experimental data.

We next implement Steps 2 and 3 of Algorithm 1. We write differential operators in the functionals Jλ,β(0)J_{\lambda,\beta}^{(0)} and Jλ,β(n)J_{\lambda,\beta}^{(n)} in the finite differences with the step size in space Δx=0.0033\Delta x=0.0033 and the step size in time Δt=0.02\Delta t=0.02 and minimize resulting functionals with respect to values of corresponding functions at grid points. Since the integrand in the definitions of the functional Jλ,β(n),n=0,1,J_{\lambda,\beta}^{(n)},n=0,1,... is the square of linear functions, then its minimizer is its critical point. In finite differences, we can write a linear system whose solution is the desired critical point. We solve this system by the command “lsqlin” of Matlab. The details in implementation by the finite difference method including the discretization, the derivation of the linear system for the critical point, and the use of “lsqlin” are very similar to the scheme in [43, 47]. Recall that in our theory, in the definition of the functional Jλ,β(n)J_{\lambda,\beta}^{(n)} acting on qnq_{n}, see (4.8), we replaced qn1q_{n-1} with its analog qn1,minq_{n-1,\mathrm{min}} which belongs to the bounded set B(R,g0)¯,\overline{B(R,g_{0})}, and also replaced qn1,min(x,0)q_{n-1,\mathrm{min}}(x,0) with qn1,min0(x,0)q_{n-1,\mathrm{min}}^{0}(x,0). These replacements are only for the theoretical purpose to avoid the case when qn1q_{n-1} blows up. However, in the numerical studies, these steps can be relaxed. This means that in Step 3, we have minimized the finite difference analog of the functional

qnΩ(qnxx(x,t)qnxt(x,t)2(qn1(x,0))2+tqn1(x,t)xqn1(x,0)2(qn1(x,0))3)2e2λφ𝑑x𝑑t+βqnH2(Ω)2,q_{n}\mapsto\int_{\Omega}\Big{(}q_{nxx}(x,t)-\frac{q_{nxt}(x,t)}{2\big{(}q_{n-1}(x,0)\big{)}^{2}}+\frac{\partial_{t}q_{n-1}\left(x,t\right)\partial_{x}q_{n-1}(x,0)}{2\big{(}q_{n-1}(x,0)\big{)}^{3}}\Big{)}^{2}e^{2\lambda\varphi}dxdt+\beta\left\|q_{n}\right\|_{H^{2}\left(\Omega\right)}^{2}, (10.3)

subject to the boundary conditions in lines 2 and 3 of (3.11). Another numerical simplification is that rather than using the H4H^{4}-norm in the regularization term, we use the H2H^{2}- norm in (10.3). Although the theoretical analysis supporting the above simplifications is missing, we did not experience any difficulties in computing the numerical solutions to Problem 3.1. All of our numerical results are satisfactory.

10.2 Numerical results for computationally simulated data

To test Algorithm 1, we present four (4) numerical examples.

Test 1 (the reference test). We first test the case of one inclusion with a high inclusion/background contrast. The true dielectric constant function c(x)c(x) has the following form

ctrue(x)={1+14e(x1)2(x1)20.22if |x1|<0.2,1otherwise.c_{\mathrm{true}}(x)=\left\{\begin{array}[]{ll}1+14e^{\frac{(x-1)^{2}}{(x-1)^{2}-0.2^{2}}}&\mbox{if }|x-1|<0.2,\\ 1&\mbox{otherwise}.\end{array}\right. (10.4)
Refer to caption
(a)
Refer to caption
(b)
Figure 1: Test 1. The true and reconstructed functions c(x),c(x), where ctruec_{\mathrm{true}} is given in (10.4). (a) The functions cinitc_{\mathrm{init}} and ccompc_{\mathrm{comp}} are obtained by Step 2 and Step 5 of Algorithm 1 respectively. (b) The consecutive relative error cncn1L(ϵ,M)/cnL(ϵ,M)\|c_{n}-c_{n-1}\|_{L^{\infty}(\epsilon,M)}/\|c_{n}\|_{L^{\infty}(\epsilon,M)}, n=1,,10.n=1,\dots,10. The data is with δ=5%\delta=5\% noise.

Thus, the inclusion/background contrast as 15:1. It is evident from Figure 1 that we can successfully detect an object. The diameter of this object is 0.4 and the distance between this object and the source is 1. The true inclusion/background contrast is 15/1. The maximal value of the computed dielectric constant is 15.28. The relative error in this maximal value is 1.89% while the noise level in the data is 5%. Although the contrast is high, our method provides good numerical solution without requiring any knowledge of the true solution. Our method converges fast. Although the initial reconstruction cinitc_{\mathrm{init}} computed in Step 2 of Algorithm 1 is not very good, see Figure 1a, one can see in Figure 1b, that the convergence occurs after only five (5) iterations. This fact verifies numerically the “contraction property” of Theorem 8.1 including the key estimate (8.20).

Test 2. The true function cc in this test has two “inclusions”,

ctrue(x)={1+5e(x0.6)2(x0.6)20.22 if |x0.6|<0.2,1+8e(x1.4)2(x1.4)20.32 if |x1.4|<0.3,1 otherwise. c_{\mathrm{true}}(x)=\left\{\begin{array}[]{ll}1+5e^{\frac{(x-0.6)^{2}}{(x-0.6)^{2}-0.2^{2}}}&\mbox{ if }\left|x-0.6\right|<0.2,\\ 1+8e^{\frac{(x-1.4)^{2}}{(x-1.4)^{2}-0.3^{2}}}&\mbox{ if }\left|x-1.4\right|<0.3,\\ 1&\mbox{ otherwise. }\end{array}\right. (10.5)

Numerical results of this test are displayed in Figure 2.

Refer to caption
(a)
Refer to caption
(b)
Figure 2: Test 2. The true and reconstructed functions c(x),c(x), where ctruec_{\mathrm{true}} is given in (10.5). (a) The functions cinitc_{\mathrm{init}} and ccompc_{\mathrm{comp}} are obtained by Step 2 and Step 5 of Algorithm 1 respectively. (b) The consecutive relative error cncn1L(ϵ,M)/cnL(ϵ,M)\|c_{n}-c_{n-1}\|_{L^{\infty}(\epsilon,M)}/\|c_{n}\|_{L^{\infty}(\epsilon,M)}, n=1,,10.n=1,\dots,10. The data is with δ=5%\delta=5\% noise.

Test 2 is more complicated than Test 1. However, we still obtain good numerical results. It is evident from Figure 2a that we can precisely detect the two inclusions without any initial guess. The true maximal values of the dielectric constant of the inclusions in the left and the right are 6 and 9 respectively. The reconstructed maximal values in those inclusions are acceptable. They are 5.31 (relative error 11.5%) and 7.8 (relative error 13.3%). As in Test 1, the initial reconstruction cinitc_{\mathrm{init}} computed in Step 2 of Algorithm 1 is far from ctrue(x).c_{\mathrm{true}}(x). Still, our iterative procedure converges after 7 iterations, see Figure 2b.

Test 3 We test the case when the function ctrue(x)c_{\mathrm{true}}(x) is discontinuous. It is a piecewise constant function given by

ctrue(x)={10 if |x1|<0.15,1 otherwise. c_{\mathrm{true}}(x)=\left\{\begin{array}[]{ll}10&\mbox{ if }\left|x-1\right|<0.15,\\ 1&\mbox{ otherwise. }\end{array}\right. (10.6)

The numerical solution for this test is presented in Figure 3.

Refer to caption
(a)
Refer to caption
(b)
Figure 3: Test 3. The true and reconstructed functions c(x),c(x), where ctruec_{\mathrm{true}} is given in (10.6). (a) The functions cinitc_{\mathrm{init}} and ccompc_{\mathrm{comp}} are obtained by Step 2 and Step 5 of Algorithm 1 respectively. (b) The consecutive relative error cncn1L(ϵ,M)/cnL(ϵ,M)\|c_{n}-c_{n-1}\|_{L^{\infty}(\epsilon,M)}/\|c_{n}\|_{L^{\infty}(\epsilon,M)}, n=1,,10.n=1,\dots,10. The data is with δ=5%\delta=5\% noise.

Although the function ctruec_{\text{true}} is not smooth and actually not even continuous, Algorithm 1 works and provides a reliable numerical solution. The computed maximal value of the dielectric constant of the object is 9.25 (relative error 7.5%), which is acceptable. The location of the object is detected precisely, see Figure 3a. As in the previous two tests, Algorithm 1 converges fast. After the fifth iteration, the reconstructed function cnc_{n} becomes unchanged. Again, this fact numerically confirms both Theorem 8.1 and the robustness of our method.

Test 4 In this test, the function ctrue(x)c_{\mathrm{true}}(x) has the following form:

ctrue(x)={3.5+0.3sin(π(x1.35)) if |x0.9|<0.5,8 if |x2|<0.37,1 otherwise. c_{\mathrm{true}}(x)=\left\{\begin{array}[]{ll}3.5+0.3\ast\sin(\pi(x-1.35))&\mbox{ if }\left|x-0.9\right|<0.5,\\ 8&\mbox{ if }\left|x-2\right|<0.37,\\ 1&\mbox{ otherwise. }\end{array}\right. (10.7)

This test is interesting since the graph of the function (10.7) consists of a curve and an inclusion. The numerical solution for this case is presented in Figure 4.

Refer to caption
(a)
Refer to caption
(b)
Figure 4: Test 4. The true and reconstructed functions c(x),c(x), where ctruec_{\mathrm{true}} is given in (10.7). (a) The functions cinitc_{\mathrm{init}} and ccompc_{\mathrm{comp}} are obtained by Step 2 and Step 5 of Algorithm 1 respectively. (b) The consecutive relative error cncn1L(ϵ,M)/cnL(ϵ,M)\|c_{n}-c_{n-1}\|_{L^{\infty}(\epsilon,M)}/\|c_{n}\|_{L^{\infty}(\epsilon,M)}, n=1,,10.n=1,\dots,10. The data is with δ=5%\delta=5\% noise.

One can observe from Figure 4a that our method to compute the initial reconstruction in Step 2 of Algorithm 1 is not very effective. However, after only 6 iterations, good numerical results are obtained. The curve in the first inclusion locally coincides with the true one and the maximal value of the computed dielectric constant within inclusion is quite accurate: it is 7.83 (relative error 2.12%). Our method converges at the iteration number 6.

Remark 10.1.

It follows from all above tests that Algorithm 1 is robust in solving a highly nonlinear and severely ill-posed Problem 3.1. It provides satisfactory numerical results with a fast convergence. For each test, the computational time to compute the numerical solution is about 29 seconds on a MacBook Pro 2019 with 2.6 GHz processor and 6 Intel i7 cores. This is almost a real time computation.

11 Numerical Results for Experimental Data

The experimental data we use to test Algorithm 1 were collected by the Forward Looking Radar built in the US Army Research Laboratory [46]. The data were initially collected to detect and identify targets mimicking shallow anti-personnel land mines and IEDs. Five mimicking devices were: a bush, a wood stake, a metal box, a metal cylinder, and a plastic cylinder. The bush and the wood stake were placed in the air, while the other three objects were buried at a few centimeters depth in the ground. We refer the reader to [46] for details about the experimental set up and the data collection. This was recently repeated in [30]. We, therefore, do not describe here the device and the data collection procedure again. Since the location of the targets can be detected by the Ground Position System (GPS), we are only interested in computing the values of the dielectric constants of those targets. We do so by our new method written in Algorithm 1.

As in the earlier works using these data [13, 30, 23, 24, 25, 35, 36, 50], we compute the function crel(x),c_{\mathrm{rel}}(x), where

crel(x)={ctargetcbckgr(x) if maxctargetcbckgr(x)>1and xD,1otherwise,c_{\mathrm{rel}}(x)=\left\{\begin{array}[]{ll}\frac{c_{\mathrm{target}}}{c_{\mathrm{bckgr}}}\left(x\right)\text{ if }\max\frac{c_{\mathrm{target}}}{c_{\mathrm{bckgr}}}\left(x\right)>1&\text{and }x\in D,\\ 1&\text{otherwise},\end{array}\right. (11.1)
crel(x)={minctargetcbckgr(x) if maxctargetcbckgr(x)1and xD,1otherwise,c_{\mathrm{rel}}(x)=\left\{\begin{array}[]{ll}\min\frac{c_{\mathrm{target}}}{c_{\mathrm{bckgr}}}\left(x\right)\text{ if }\max\frac{c_{\mathrm{target}}}{c_{\mathrm{bckgr}}}\left(x\right)\leq 1&\text{and }x\in D,\\ 1&\text{otherwise},\end{array}\right. (11.2)

where DD is a sub interval of the interval [ϵ,M],[\epsilon,M], which is occupied by the target. Next, we define the computed value of ctargetc_{\mathrm{target}} as [23]:

ccomp=cbckgr×{maxcrel(x) if maxcrel(x)>1,mincrel(x) if maxcrel(x)<1.c_{\mathrm{comp}}=c_{\mathrm{bckgr}}\times\left\{\begin{array}[]{c}\max c_{\mathrm{rel}}(x)\text{ if }\max c_{\mathrm{rel}}(x)>1,\\ \min c_{\mathrm{rel}}(x)\text{ if }\max c_{\mathrm{rel}}(x)<1.\end{array}\right. (11.3)

As in the above cited publications, we have to preprocess the raw data of [46] before importing them into our solver. The data preprocessing is exactly the same as in [30, Section 7.1]. First, we observe that the LL_{\infty}-norm of the experimental data far exceeds the LL_{\infty}-norm of the computationally simulated data. Therefore, we need to scale the experimental data by dividing it by a calibration factor μ>0,\mu>0, i.e. we replace the raw experimental data fraw(t)f_{\text{raw}}\left(t\right) with fscale(t)=fraw(t)/μf_{\text{scale}}\left(t\right)=f_{\text{raw}}\left(t\right)/\mu. Then we work only with fscale(t).f_{\text{scale}}\left(t\right). To find the calibration factor, we use a trial-and-error process. First, we select a reference target. We then try many values of μ\mu such that the reconstruction of the reference target is satisfactory, i.e. the computed dielectric constant is in its published range, see below in this section about the published range. Then this calibration factor is used is all other tests. When the object is in the air, our reference target is bush. In this case, the calibration factor μair=534,592.\mu_{\mathrm{air}}=534,592. When the object is buried under the ground, our reference target is the metal box and the calibration factor was μground=265,223\mu_{\mathrm{ground}}=265,223.

Next, we preprocess the data fscale(t)f_{\text{scale}}\left(t\right) as follows. First, we first use a lower envelop (built in Matlab) to bound the data. We then truncate the data that is not in a small interval centered at the global minimizer of the data, see [30, Section 7.1] for the choice of this small interval. But in the case of the plastic cylinder we use the upper envelop. The choice of the upper or lower envelopes is as follows. We look at the function fscale(t)f_{\text{scale}}\left(t\right) and find the three extrema with largest absolute values. If the middle extremal value among these three is a minimum, then we bound the data by a lower envelop. Otherwise, we use an upper envelop. See [30, Section 7.1] for more details and the reason of this choice. In particular, Figures 4a, 4b, 4d, 5a, 5b, 5d, 5e, 5g, 5h of [30] provide illustrations. Likewise, our Figures 5 displays computed functions ctarget(x)c_{\text{target}}\left(x\right) for our five targets. The computed dielectric constants ccompc_{\mathrm{comp}} defined in (11.3) by Algorithm 1 is listed in Table 1.

Refer to caption
(a) ctargetc_{\mathrm{t}arget} for bush
Refer to caption
(b) ctargetc_{\mathrm{t}arget} for wood stake
Refer to caption
(c) ctargetc_{\mathrm{t}arget} for metal box
Refer to caption
(d) ctargetc_{\mathrm{t}arget} for metal cylinder
Refer to caption
(e) ctargetc_{\mathrm{t}arget} for plastic cylinder
Figure 5: Computed functions ctarget(x,y)c_{\mathrm{target}}(x,y) for our five targets, also see (11.1)- (11.3) and Table 1.
Target cbckgrc_{\mathrm{bckgr}} computed crelc_{\mathrm{rel}} cbckgrc_{\mathrm{bckgr}} computed ctargetc_{\text{target}} True ctargetc_{\text{target}}
Bush 1 7.62 1 10.99 [3,20][3,20]
Wood stake 1 2.01 1 2.26 [2,6][2,6]
Metal box 4 4.00 [3,5][3,5] [12.00,20.00][12.00,20.00] [10,30][10,30]
Metal cylinder 4 4.01 [3,5][3,5] [12.3,20.5][12.3,20.5] [10,30][10,30]
Plastic cylinder 4 0.59 [3,5][3,5] [1.6,2.95][1.6,2.95] [1.1,3.2]\left[1.1,3.2\right]
Table 1: Computed dielectric constants of five targets

The true values of dielectric constants of our targets were not measured in the experiment. Therefore, we compare our computed values with the published ones. The published values of the dielectric constants of our targets are listed in the last column of Table 1. They can be found on the website of Honeywell (Table of dielectric constants, https://goo.gl/kAxtzB). Also, see [9] for the experimentally measured range of the dielectric constants of vegetation samples, which we assume have the same range as the dielectric constant of bush. In the table of dielectric constants of Honeywell as well as in [9], any dielectric constant is not a number. Rather, each dielectric constant of these references is given within a certain interval. As to the metallic targets, it was established in [35] that they have the so-called “apparent” dielectric constant whose values are in the interval [10,30].\left[10,30\right].

Conclusion. It is clear from Table 1 that our computed dielectric constants are consistent with the intervals of published ones. Therefore, our results for experimental data are satisfactory.

Remark 11.1 (Speed of computations).

Our experimental data are sparse. The size of the data in time is Nt=80N_{t}=80, which is a lot smaller than that in the dense simulated data (Nt=300N_{t}=300). Therefore, the speed of computations is much faster than for the case of simulated data of section 10. Thus, all results of this section were computed in real time on the same computer (MacBook Pro 2019 with 2.6 GHz processor and 6 Intel i7 cores).

12 Concluding Remarks

We have developed a new globally convergent numerical method for a 1-D Coefficient Inverse Problem with backscattering data for the wave-like PDE (2.3). This is the second generation of the above cited convexification method of our research group. The main novelty here is that, rather than minimizing a globally strictly convex weighted cost functional arising in the convexification, we solve on each iterative step a linear boundary value problem. This is done using the so-called Carleman Quasi-Reversibility Method. Just like in the convexification, the key element of the technique of this paper, on which our convergence analysis is based, is the presence of the Carleman Weight Function in each quadratic functional which we minimize. The convergence estimate is similar to the key estimate of the classical contraction mapping principle. The latter explains the title of this paper. We have proven a global convergence theorem of our method. Our numerical results for computationally simulated data demonstrate high reconstruction accuracies in the presence of 5% random noise in the data.

Furthermore, our numerical results for experimentally collected data have satisfactory accuracy via providing values of computed dielectric constants of explosive-like targets within their published ranges.

A practically important observation here is that our computations for experimental data were performed in real time.

Acknowledgement

The effort of TTL, MVK and LHN was supported by US Army Research Laboratory and US Army Research Office grant W911NF-19-1-0044.

References

  • [1] A. B. Bakushinskii, M. V. Klibanov, and N. A. Koshev. Carleman weight functions for a globally convergent numerical method for ill-posed Cauchy problems for some quasilinear PDEs. Nonlinear Anal. Real World Appl., 34:201–224, 2017.
  • [2] L. Baudouin, M. de Buhan, and S. Ervedoza. Convergent algorithm based on Carleman estimates for the recovery of a potential in the wave equation. SIAM J. Nummer. Anal., 55:1578–1613, 2017.
  • [3] L. Baudouin, M. de Buhan, S. Ervedoza, and A. Osses. Carleman-based reconstruction algorithm for the waves. SIAM Journal on Numerical Analysis, 59(2):998–1039, 2021.
  • [4] L. Beilina and M. V. Klibanov. Approximate Global Convergence and Adaptivity for Coefficient Inverse Problems. Springer, New York, 2012.
  • [5] M. Bellassoued and M. Yamamoto. Carleman Estimates and Applications to Inverse Problems for Hyperbolic Systems. Springer, Japan, 2017.
  • [6] A. L. Bukhgeim and M. V. Klibanov. Uniqueness in the large of a class of multidimensional inverse problems. Soviet Math. Doklady, 17:244–247, 1981.
  • [7] T. Carleman. Sur un probleme d’unicit e pur les systeé mes d’equations aux dériveés partielles à deux varibales indépendantes. Ark. Mat. Astr. Fys., 26B:1–9, 1939.
  • [8] G. Chavent. Nonlinear Least Squares for Inverse Problems: Theoretical Foundations and Step-by-Step Guide for Applications, Scientic Computation. Springer, New York, 2009.
  • [9] H. T. Chuah, K. Y. Lee, and T. W. Lau. Dielectric constants of rubber and oil palm leaf samples at X-band. IEEE Trans. Geosci. Remote Sens, 33, 221–223, 1995.
  • [10] M. de Hoop, P. Kepley, and L. Oksanen. Recovery of a smooth metric via wave field and coordinate transformation reconstruction. SIAM J. Appl. Math., 78:1931–1953, 2018.
  • [11] A. V. Goncharsky and S. Y. Romanov. Iterative methods for solving coefficient inverse problems of wave tomography in models with attenuation. Inverse Problems, 33:025003, 2017.
  • [12] V. Goncharsky and S. Y. Romanov. A method of solving the coefficient inverse problems of wave tomography. Comput. Math. Appl., 77:967–980, 2019.
  • [13] A.L. Karchevsky, M.V. Klibanov, L. Nguyen, N. Pantong, and A. Sullivan. The Krein method and the globally convergent method for experimental data. Applied Numerical Mathematics, 74:111–127, 2013.
  • [14] V. A. Khoa, G. W. Bidney, M. V. Klibanov, L. H. Nguyen, L. Nguyen, A. Sullivan, and V. N. Astratov. Convexification and experimental data for a 3D inverse scattering problem with the moving point source. Inverse Problems, 36:085007, 2020.
  • [15] V. A. Khoa, G. W. Bidney, M. V. Klibanov, L. H. Nguyen, L. Nguyen, A. Sullivan, and V. N. Astratov. An inverse problem of a simultaneous reconstruction of the dielectric constant and conductivity from experimental backscattering data. Inverse Problems in Science and Engineering, 29(5):712–735, 2021.
  • [16] V. A. Khoa, M. V. Klibanov, and L. H. Nguyen. Convexification for a 3D inverse scattering problem with the moving point source. SIAM J. Imaging Sci., 13(2):871–904, 2020.
  • [17] M. V. Klibanov. Inverse problems and Carleman estimates. Inverse Problems, 8:575–596, 1992.
  • [18] M. V. Klibanov. Carleman estimates for global uniqueness, stability and numerical methods for coefficient inverse problems. J. Inverse and Ill-Posed Problems, 21:477–560, 2013.
  • [19] M. V. Klibanov. Carleman estimates for the regularization of ill-posed Cauchy problems. Applied Numerical Mathematics, 94:46–74, 2015.
  • [20] M. V. Klibanov and O. V. Ioussoupova. Uniform strict convexity of a cost functional for three-dimensional inverse scattering problem. SIAM J. Math. Anal., 26:147–179, 1995.
  • [21] M. V. Klibanov. Global convexity in a three-dimensional inverse acoustic problem. SIAM J. Math. Anal., 28:1371–1388, 1997.
  • [22] M. V. Klibanov and A. Timonov. Carleman Estimates for Coefficient Inverse Problems and Numerical Applications. Inverse and Ill-Posed Problems Series. VSP, Utrecht, 2004.
  • [23] M. V. Klibanov, L. H. Nguyen, A. Sullivan, and L. Nguyen. A globally convergent numerical method for a 1-d inverse medium problem with experimental data. Inverse Problems and Imaging, 10:1057–1085, 2016.
  • [24] M. V. Klibanov, A. E. Kolesov, L. Nguyen, and A. Sullivan. Globally strictly convex cost functional for a 1-D inverse medium scattering problem with experimental data. SIAM J. Appl. Math., 77:1733–1755, 2017.
  • [25] M. V. Klibanov, A. E. Kolesov, L. Nguyen, and A. Sullivan. A new version of the convexification method for a 1-D coefficient inverse problem with experimental data. Inverse Problems, 34:35005, 2018.
  • [26] M. V. Klibanov, A. E. Kolesov, and D-L. Nguyen. Convexification method for an inverse scattering problem and its performance for experimental backscatter data for buried targets. SIAM J. Imaging Sci., 12:576–603, 2019.
  • [27] M. V. Klibanov, J. Li, and W. Zhang. Convexification of electrical impedance tomography with restricted Dirichlet-to-Neumann map data. Inverse Problems, 35:035005, 2019.
  • [28] M. V. Klibanov, Z. Li, and W. Zhang. Convexification for the inversion of a time dependent wave front in a heterogeneous medium. SIAM J. Appl. Math., 79:1722–1747, 2019.
  • [29] M. V. Klibanov, J. Li, and W. Zhang. Convexification for an inverse parabolic problem. Inverse Problems, 36:085008, 2020.
  • [30] M. V. Klibanov, T. T. Le, L. H. Nguyen, A. Sullivan, and L. Nguyen. Convexification-based globally convergent numerical method for a 1D coefficient inverse problem with experimental data. preprint, Arxiv:2104.11392, 2021.
  • [31] M. V. Klibanov and J. Li. Inverse Problems and Carleman Estimates: Global Uniqueness, Global Convergence and Experimental Data. , De Gruyter, 2021.
  • [32] M. V. Klibanov, A. V. Smirnov, V. A. Khoa, A. Sullivan, and L. Nguyen. Through-the-wall nonlinear SAR imaging. IEEE Transactions on Geoscience and Remote Sensing, 59: 7474-7486, 2021.
  • [33] M. V. Klibanov, V. A. Khoa, A. V. Smirnov, L. H. Nguyen, G. W. Bidney, L. Nguyen, A. Sullivan, and V. N. Astratov. Convexification inversion method for nonlinear SAR imaging with experimentally collected data. to appear in J. Applied and Industrial Mathematics, see also Arxiv:2103.10431, 2021.
  • [34] J. Korpela, M. Lassas, and L. Oksanen. Regularization strategy for an inverse problem for a 1+11+1 dimensional wave equation. Inverse Problems, 32:065001, 2016.
  • [35] A. Kuzhuget, L. Beilina, M.V. Klibanov, A. Sullivan, Lam Nguyen, and M. A. Fiddy. Blind backscattering experimental data collected in the field and an approximately globally convergent inverse algorithm. Inverse Problems, 28:095007, 2012.
  • [36] A. V. Kuzhuget, L. Beilina, M. V. Klibanov, A. Sullivan, L. Nguyen, and M. A. Fiddy. Quantitative image recovery from measured blind backscattered data using a globally convergent inverse method. IEEE Trans. Geosci. Remote Sens, 51:2937–2948, 2013.
  • [37] R. Lattès and J. L. Lions. The Method of Quasireversibility: Applications to Partial Differential Equations. Elsevier, New York, 1969.
  • [38] M. M. Lavrent’ev, V. G. Romanov, and S. P. Shishat\cdotskiĭ. Ill-Posed Problems of Mathematical Physics and Analysis. Translations of Mathematical Monographs. AMS, Providence: RI, 1986.
  • [39] T. T. Le and L. H. Nguyen. A convergent numerical method to recover the initial condition of nonlinear parabolic equations from lateral Cauchy data. Journal of Inverse and Ill-posed Problems, DOI: https://doi.org/10.1515/jiip-2020-0028, 2020.
  • [40] T. T. Le and L. H. Nguyen. The gradient descent method for the convexification to solve boundary value problems of quasi-linear PDEs and a coefficient inverse problem. preprint Arxiv:2103.04159, 2021.
  • [41] B. M. Levitan. Inverse Sturm–Liouville Problems. VSP, Utrecht, 1987.
  • [42] M. Minoux. Mathematical Programming: Theory and Algorithms. John Wiley & Sons, New York, 1986.
  • [43] L. H. Nguyen. A new algorithm to determine the creation or depletion term of parabolic equations from boundary measurements. Computers and Mathematics with Applications, 80:2135–2149, 2020.
  • [44] L. H. Nguyen and M. V. Klibanov. Carleman estimates and the contraction principle for an inverse source problem for nonlinear hyperbolic equations. preprint arXiv:2108.03500, 2021.
  • [45] L. H. Nguyen and M. V. Klibanov. Carleman estimates and the contraction principle for an inversesource problem for nonlinear hyperbolic equations. preprint arXix:2108.03500v3, 2021.
  • [46] N. Nguyen, D. Wong, M. Ressler, F. Koenig, Stanton B., G. Smith, J. Sichina, and K. Kappra. Obstacle avolidance and concealed target detection using the Army Research Lab ultra-wideband synchronous impulse Reconstruction (UWB SIRE) forward imaging radar. Proc. SPIE, 6553:65530H (1)–65530H (8), 2007.
  • [47] P. M. Nguyen and L. H. Nguyen. A numerical method for an inverse source problem for parabolic equations and its application to a coefficient inverse problem. Journal of Inverse and Ill-posed Problems, 38:232–339, 2020.
  • [48] V.G. Romanov. Inverse Problems of Mathematical Physics, VNU Press, 1986.
  • [49] J. A. Scales, M. L. Smith, and T. L. Fischer. Global optimization methods for multimodal inverse problems. J. Computational Physics, 103:258–268, 1992.
  • [50] A. V. Smirnov, M. V. Klibanov, and L. H. Nguyen. Convexification for a 1D hyperbolic coefficient inverse problem with single measurement data. Inverse Probl. Imaging, 14(5):913–938, 2020.
  • [51] A. V. Smirnov, M. V. Klibanov, A. Sullivan, and L. Nguyen. Convexifcation for an inverse problem for a 1D wave equation with experimental data. Inverse Problems, 36:095008, 2020.
  • [52] A. N. Tikhonov, A. Goncharsky, V. V. Stepanov, and A. G. Yagola. Numerical Methods for the Solution of Ill-Posed Problems. Kluwer Academic Publishers Group, Dordrecht, 1995.
  • [53] M. Yamamoto. Carleman estimates for parabolic equations. Topical Review. Inverse Problems, 25:123013, 2009.