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

Constructing a variational quasi-reversibility method for a Cauchy problem for elliptic equations

Vo Anh Khoa11footnotemark: 1 Department of Mathematics and Statistics, University of North Carolina at Charlotte, Charlotte, North Carolina 28223, USA [email protected] Pham Truong Hoang Nhan22footnotemark: 2 Department of Mathematics and Computer Science, VNUHCM-University of Science, 227 Nguyen Van Cu Str., Dist. 5, Ho Chi Minh City, Vietnam [email protected]
Abstract

In the recent developments of regularization theory for inverse and ill-posed problems, a variational quasi-reversibility (QR) method has been designed to solve a class of time-reversed quasi-linear parabolic problems. Known as a PDE-based approach, this method relies on adding a suitable perturbing operator to the original problem and consequently, on gaining the corresponding fine stabilized operator, which leads us to a forward-like problem. In this work, we establish new conditional estimates for such operators to solve a prototypical Cauchy problem for elliptic equations. This problem is based on the stationary case of the inverse heat conduction problem, where one wants to identify the heat distribution in a certain medium, given the partial boundary data. Using the new QR method, we obtain a second-order initial value problem for a wave-type equation, whose weak solvability can be deduced using a priori estimates and compactness arguments. Weighted by a Carleman-like function, a new type of energy estimates is explored in a variational setting when we investigate the Hölder convergence rate of the proposed scheme. Besides, a linearized version of this scheme is analyzed. Numerical examples are provided to corroborate our theoretical analysis.

keywords:
Inverse and ill-posed problems, quasi-reversibility method , convergence rates , energy estimates, Carleman weight
MSC:
65J05 , 65J20 , 35K92
journal: TBA

1 Introduction

1.1 Statement of the inverse problem

Solving boundary value determination problems is one of the classical research topics in the field of inverse and ill-posed problems; cf. the survey [1] for the background of some classical inverse problems for partial differential equations (PDEs). Physically, this type of problems stems from the stationary case of the inverse heat conduction problem. In this regard, we seek the unknown temperature distribution in a certain medium when informative data are given on some parts of the boundary. In terms of PDEs, this finding is governed by the so-called Cauchy problem for elliptic equations, which is highly ill-posed in the sense of Hadamard. In this work, we look for the real-valued function u(x,y)u\left(x,y\right) in a unit rectangle [0,1]×[0,1][0,1]\times[0,1] from the data u0H1(0,1)u_{0}\in H^{1}(0,1) at x=0x=0, when uu obeys the following Laplace system:

{uxx+uyy=0in (0,1)×(0,1),u(x,0)=u(x,1)=0for x[0,1],u(0,y)=u0(y),ux(0,y)=0for y[0,1].\displaystyle\begin{cases}u_{xx}+u_{yy}=0&\text{in }\left(0,1\right)\times\left(0,1\right),\\ u\left(x,0\right)=u\left(x,1\right)=0&\text{for }x\in[0,1],\\ u\left(0,y\right)=u_{0}\left(y\right),u_{x}\left(0,y\right)=0&\text{for }y\in[0,1].\end{cases} (1)

Here, the zero Neumann boundary condition at x=0x=0 means that there is no heat entering or escaping at this boundary. Meanwhile, we assume the zero Dirichlet conditions at y=0,1y=0,1. Since in real-world applications the data u0u_{0} can only be measured, we suppose to have the measured data u0εH1(0,1)u_{0}^{\varepsilon}\in H^{1}(0,1) associated with some noise level ε(0,1)\varepsilon\in(0,1), which satisfies

u0εu0H1(0,1)ε.\displaystyle\left\|u_{0}^{\varepsilon}-u_{0}\right\|_{H^{1}\left(0,1\right)}\leq\varepsilon. (2)

Even though the model (1) is the simplest case of the Cauchy problem for elliptic equations, its extensions to more general scenarios were already mentioned in the previous works for regularization of Laplace equations. In fact, some particular generalizations have the analytical capability of reducing to the system (1); cf. subsection 5.3 for the revisited. The existing literature on the Cauchy problems for elliptic equations is vast from theoretical and numerical viewpoints. In principle, such problems are solved for decades by distinctive approaches and thus, it is pertinent to address some fundamental researches. For instance, we would like to mention here the spectral regularization method for the sustainable development in the field of inverse and ill-posed problems; cf. e.g. [2, 3, 4, 5]. This method and its variants are based on stabilizing the unbounded kernels appearing in the explicit representation of solution. The Tikhonov-type regularization with convex and strictly convex functionals in [6, 7, 8, 9] has also received much attention in this field, throughout the minimization procedure. This method is essentially related to the variational logarithmic convexity in [10, 11], where it “convexifies” the energy functional logarithmically using a Carleman-like weight. Since this notion further concerns massive Carleman estimates, details of the convexification can be referred to the survey [12] and references cited therein. Some other approaches should be addressed here include the moment method in [13] and the integral equation approach in [14, 15].

1.2 Goals and novelty

This paper is aimed at enhancing our understanding of the application of the novel quasi-reversibility (QR) method which has been designed so far in our pioneering work by [16] for regularization of time-reversed parabolic systems. Based upon the original QR method by Lattès and Lions in the textbook [17], this is a PDE-based approach that consists in the establishment of two operators along with their conditional estimates. We call those the perturbing and stabilized operators. In this setting, the perturbing operator is added to the original PDE for the sake of “absorbing” the unbounded operator (i.e. the aimed operator to be stabilized in the PDE). Then we obtain a stable approximate problem with the corresponding stabilized operator, which is usually referred to as the regularized problem. The key idea of our new QR method lies in the fact that this addition turns the inverse problem into the forward-like problem by just acquiring the boundedness of the leading coefficient of the unbounded operator. Thereby, it reveals the perfect connection between the inverse and forward problems right in the governing equations. It is obvious that there are certainly hundreds of numerical methods to solve the forward problem and thus, our inversion method would be implemented easily.

Due to the aforementioned major feature, it is a vast potential that our QR method can be extended to many different concerns in the field of inverse problems. Since this is the inception stage of this approach, we deliberately apply it to design a regularized problem for the severely ill-posed problem (1) using the measurement u0εu_{0}^{\varepsilon} observed in (2). When doing so, we explore that the regularized problem for (1) is essentially expressed as an acoustic wave equation, which singles out one of the prominent aspects of the method we are studying. It is worth mentioning that the conditional estimates for our perturbing and stabilized operators are crucial for convergence analysis of the QR scheme, and they vary for different types of PDEs. Henceforth, our first novelty here is devoted to deriving these new conditional estimates for the Cauchy problem of elliptic equations. Especially, such estimates are established with relevance to the practical and computational aspects.

As in [16], to prove the convergence, we rely on a Carleman weight function to deduce a noise-scaled difference problem. In the standard variational setting, this weight plays a vital role in getting rid of large quantities involved in the difference system, which yields our second novelty (compared to many other works for inverse boundary value problems). Then our proof is centered around a weighted energy method that is coupled with careful conditional estimates for the established operators. The obtained error estimate is of the Hölder rate under some certain assumptions on the true solution. In some sense, the true solution can be assumed as a unique strong solution in the standard elliptic boundary value problem. We also analyze the convergence of a linearized version of the proposed QR scheme.

1.3 Outline of this article

The rest of the paper is organized as follows. Section 2 is devoted to the concrete setting of the QR framework we would like to study in this work. In this part, we define a regularized system for our Laplace problem (1). In section 3, we prove the weak solvability of the regularized system using a priori estimates with compactness arguments. Convergence analysis is then conducted in section 4, where we exploit a Carleman-like weight to prove the Hölder rate of convergence of the scheme. In section 5, we discuss a choice of the operators we are constructing in the QR framework. Furthermore, we design a linearized version of the QR scheme and show its convergence by making a condition between the discretization in xx and the noise level ε\varepsilon. We also recall some generalizations of (1) and address some concerns about large noise levels, where our method can be modified to use. We provide two numerical examples in section 6 to see how our method works. Essentially, it performs very well when ε=10%,1%\varepsilon=10\%,1\%; these types of noise are relevant in physical applications. We close the paper by some conclusions in 7 for some works in the near future.

2 A modified quasi-reversibility framework

In the sequel, ,\left\langle\cdot,\cdot\right\rangle indicates either the scalar product in L2(0,1)L^{2}(0,1) or the dual pairing of a continuous linear functional and an element of a function space. We mean \left\|\cdot\right\| the norm in the Hilbert space L2(0,1)L^{2}(0,1). Different inner products and norms should be written as ,X\left\langle\cdot,\cdot\right\rangle_{X} and X\left\|\cdot\right\|_{X}, respectively, where XX is a certain Banach space. Throughout this paper, CC denotes a non-negative constant which is not dependent on the noise level ε\varepsilon. This constant may vary from line to line, but we usually indicate its dependences if necessary.

From now on, we introduce the so-called regularization parameter β(ε)(0,1)\beta\left(\varepsilon\right)\in(0,1) such that β0\beta\to 0 as ε0\varepsilon\to 0. In doing so, we consider an auxiliary function γ:(0,1)\gamma:(0,1)\to\mathbb{R} such that for β(0,1)\beta\in(0,1) there holds

γ(β)1,limβ0γ(β)=.\gamma(\beta)\geq 1,\quad\lim_{\beta\to 0}\gamma(\beta)=\infty.
Definition 2.1 (perturbing operator).

The linear mapping 𝐐εβ:L2(0,1)L2(0,1)\mathbf{Q}_{\varepsilon}^{\beta}:L^{2}(0,1)\to L^{2}(0,1) is said to be a perturbing operator if there exist a function space 𝕎L2(0,1)\mathbb{W}\subset L^{2}(0,1) and a noise-independent constant C0>0C_{0}>0 such that

𝐐εβuC0u𝕎/γ(β)for any u𝕎.\displaystyle\left\|\mathbf{Q}_{\varepsilon}^{\beta}u\right\|\leq C_{0}\left\|u\right\|_{\mathbb{W}}/\gamma(\beta)\quad\text{for any }u\in\mathbb{W}. (3)
Definition 2.2 (stabilized operator).

The linear mapping 𝐏εβ:H1(0,1)L2(0,1)\mathbf{P}_{\varepsilon}^{\beta}:H^{1}(0,1)\to L^{2}(0,1) is said to be a stabilized operator if there exists a noise-independent constant C1>0C_{1}>0 such that

𝐏εβuC1log(γ(β))uH1(0,1)for any uH1(0,1).\displaystyle\left\|\mathbf{P}_{\varepsilon}^{\beta}u\right\|\leq C_{1}\log\left(\gamma\left(\beta\right)\right)\left\|u\right\|_{H^{1}(0,1)}\quad\text{for any }u\in H^{1}\left(0,1\right). (4)

For each noise level, our quasi-reversibility scheme is constructed in the following manner. We perturb the PDE in (1) by a linear mapping 𝐐εβ\mathbf{Q}_{\varepsilon}^{\beta} and take 𝐏εβ=𝐐εβ+22/y2\mathbf{P}_{\varepsilon}^{\beta}=\mathbf{Q}_{\varepsilon}^{\beta}+2\partial^{2}/\partial y^{2}. Thus, we have

2x2uβε2y2uβε+𝐏εβuβε=0in (0,1)×(0,1),\displaystyle\frac{\partial^{2}}{\partial x^{2}}u_{\beta}^{\varepsilon}-\frac{\partial^{2}}{\partial y^{2}}u_{\beta}^{\varepsilon}+\mathbf{P}_{\varepsilon}^{\beta}u_{\beta}^{\varepsilon}=0\quad\text{in }(0,1)\times(0,1), (5)

associated with the Dirichlet boundary condition and the initial conditions:

{uβε(x,0)=uβε(x,1)=0for x[0,1],uβε(0,y)=u0ε(y),xuβε(0,y)=0for y[0,1].\displaystyle\begin{cases}u_{\beta}^{\varepsilon}\left(x,0\right)=u_{\beta}^{\varepsilon}\left(x,1\right)=0&\text{for }x\in[0,1],\\ u_{\beta}^{\varepsilon}\left(0,y\right)=u_{0}^{\varepsilon}\left(y\right),\partial_{x}u_{\beta}^{\varepsilon}\left(0,y\right)=0&\text{for }y\in[0,1].\end{cases} (6)

Essentially, (5)–(6) form our regularized problem.

Remark 2.3.

The variable xx in (5)–(6) can be understood as a parametric time. Therefore, system (5)–(6) resembles a Dirichlet–Cauchy wave equation in a unit rectangle controlled by the noise level ε\varepsilon. Since energy of the stabilized term 𝐏εβuβε\mathbf{P}_{\varepsilon}^{\beta}u_{\beta}^{\varepsilon} is large with respect to the noise argument, a careful adaption of fundamental energy techniques that we usually enjoy in forward problems for wave equations is really needed.

According to the standard result for the Dirichlet eigenvalue problem, there exists an orthonormal basis of L2(0,1)L^{2}(0,1), denoted by {ϕj}j\left\{\phi_{j}\right\}_{j\in\mathbb{N}}, such that ϕjH01(0,1)C[0,1]\phi_{j}\in H^{1}_{0}(0,1)\cap C^{\infty}[0,1] and y22ϕj(y)=μjϕj(y)-\partial^{2}_{y^{2}}\phi_{j}(y)=\mu_{j}\phi_{j}(y) for y(0,1)y\in(0,1). The Dirichlet eigenvalues {μj}j\left\{\mu_{j}\right\}_{j\in\mathbb{N}} form an infinite sequence which goes to infinity in the following sense

0μ0<μ1<μ2<, and limjμj=.0\leq\mu_{0}<\mu_{1}<\mu_{2}<\ldots,\text{ and }\lim_{j\to\infty}\mu_{j}=\infty.
Remark 2.4.

The conditional estimate (4) is weaker than the one we have proposed in [16] because of the following three reasons. First, the regularized problem in this work is expressed as a hyperbolic equation; compared to the parabolic one in [16], which turns out that we eventually need the information in H1(0,1)H^{1}(0,1) during the energy estimates. Second, we mimic the Fourier truncation method to design a computable stabilized operator. It is due to the fact that the stabilized operators we have introduced in [16] are formed by a Fourier series with modified kernels, and it is hard to approximate an infinite series without truncating high frequencies in a suitable manner. Finally, the space 𝕎\mathbb{W} for the true solution we need for our convergence analysis is usually a Gevrey-type space, which is not natural to be assumed according to the forward problem for linear elliptic equations. This is completely different from the Gevrey assumption on the true solution we have made in [16] because it is well-known that for some analytical parabolic equations they possess themselves a local-in-time weak solution in Gevrey spaces; cf. [18]. In the present work, we rely on a special property of solution using the Fourier representation to show a C1C^{1} regularity bound for a particular Gevrey criterion we need. This bound is a perfect match for the estimate (4). The second and third reasons can be manifested in subsection 5.1.

3 Weak solvability of the regularized system (5)–(6)

To show the weak solvability of the regularized problem, we introduce its weak formulation in the following manner.

Definition 3.1.

For each ε>0\varepsilon>0, a function uβε:[0,1]H01(0,1)u_{\beta}^{\varepsilon}:[0,1]\to H_{0}^{1}(0,1) is said to be a weak solution to system (5)–(6) if

  • 1.

    uβεC([0,1];H01(0,1))u_{\beta}^{\varepsilon}\in C([0,1];H_{0}^{1}(0,1)), xuβεC([0,1];L2(0,1))\partial_{x}u_{\beta}^{\varepsilon}\in C([0,1];L^{2}(0,1)), x22uβεL2(0,1;H1(0,1))\partial_{x^{2}}^{2}u_{\beta}^{\varepsilon}\in L^{2}(0,1;H^{-1}(0,1));

  • 2.

    For every test function ψH01(0,1)\psi\in H_{0}^{1}(0,1), it holds that

    2x2uβε,ψ+yuβε,yψ+𝐏εβuβε,ψ=0for a.e. in (0,1);\displaystyle\left\langle\frac{\partial^{2}}{\partial x^{2}}u_{\beta}^{\varepsilon},\psi\right\rangle+\left\langle\partial_{y}u_{\beta}^{\varepsilon},\partial_{y}\psi\right\rangle+\left\langle\mathbf{P}_{\varepsilon}^{\beta}u_{\beta}^{\varepsilon},\psi\right\rangle=0\quad\text{for a.e. in }(0,1); (7)
  • 3.

    uβε(0)=u0εu_{\beta}^{\varepsilon}(0)=u_{0}^{\varepsilon} and xuβε(0)=0\partial_{x}u_{\beta}^{\varepsilon}(0)=0.

Consider the nn-dimensional subspace 𝕊n\mathbb{S}_{n} of H01(0,1)H_{0}^{1}(0,1) generated by ϕ0,ϕ1,,ϕn\phi_{0},\phi_{1},\ldots,\phi_{n}. Using the Galerkin projection, we construct approximate solutions to (5)–(6) in the form

unε(x,y)=j=0nUjnε(x)ϕj(y).\displaystyle u_{n}^{\varepsilon}(x,y)=\sum_{j=0}^{n}U_{jn}^{\varepsilon}(x)\phi_{j}(y). (8)

Note that for ease of presentation, we neglect the presence of β\beta in this section. The solution unεu_{n}^{\varepsilon} is the solution of the following approximate problem:

2x2unε,ψ+yunε,yψ+𝐏εβunε,ψ=0for ψ𝕊nand a.e. in (0,1),\displaystyle\left\langle\frac{\partial^{2}}{\partial x^{2}}u_{n}^{\varepsilon},\psi\right\rangle+\left\langle\partial_{y}u_{n}^{\varepsilon},\partial_{y}\psi\right\rangle+\left\langle\mathbf{P}_{\varepsilon}^{\beta}u_{n}^{\varepsilon},\psi\right\rangle=0\quad\text{for }\psi\in\mathbb{S}_{n}\;\text{and a.e. in }(0,1), (9)

with xunε(0)=0\partial_{x}u_{n}^{\varepsilon}\left(0\right)=0 and

unε(0)=j=0n(U0ε)jnϕju0εstrongly in H1(0,1)as n.\displaystyle u_{n}^{\varepsilon}\left(0\right)=\sum_{j=0}^{n}\left(U_{0}^{\varepsilon}\right)_{jn}\phi_{j}\to u_{0}^{\varepsilon}\quad\text{strongly in }H^{1}\left(0,1\right)\;\text{as }n\to\infty. (10)

By the choice ψ=ϕj\psi=\phi_{j}, the functions UjnεU_{jn}^{\varepsilon} in (8) are to be found as solutions to the Cauchy problem for the system of nn ordinary differential equations:

d2dx2Ujnε+μjUjnε+i=0nUinε𝐏εβϕi,ϕj=0,\displaystyle\frac{d^{2}}{dx^{2}}U_{jn}^{\varepsilon}+\mu_{j}U_{jn}^{\varepsilon}+\sum_{i=0}^{n}U_{in}^{\varepsilon}\left\langle\mathbf{P}_{\varepsilon}^{\beta}\phi_{i},\phi_{j}\right\rangle=0, (11)
Ujnε(0)=(U0ε)jn,ddxUjnε(0)=0.\displaystyle U_{jn}^{\varepsilon}\left(0\right)=\left(U_{0}^{\varepsilon}\right)_{jn},\frac{d}{dx}U_{jn}^{\varepsilon}\left(0\right)=0. (12)
Lemma 3.2.

For any fixed nn\in\mathbb{N} and for each ε>0\varepsilon>0, system (11)–(12) has a unique solution UjnεC1([0,1])U_{jn}^{\varepsilon}\in C^{1}([0,1]).

Proof.

Let Zjnε=ddxUjnεZ_{jn}^{\varepsilon}=\frac{d}{dx}U_{jn}^{\varepsilon}. It follows from system (11)–(12) that

ddx[UjnεZjnε]=[01μj0][UjnεZjnε]+[0i=0nUinε𝐏εβϕi,ϕj],\displaystyle\frac{d}{dx}\begin{bmatrix}U_{jn}^{\varepsilon}\\ Z_{jn}^{\varepsilon}\end{bmatrix}=\begin{bmatrix}0&1\\ -\mu_{j}&0\end{bmatrix}\begin{bmatrix}U_{jn}^{\varepsilon}\\ Z_{jn}^{\varepsilon}\end{bmatrix}+\begin{bmatrix}0\\ -\sum_{i=0}^{n}U_{in}^{\varepsilon}\left\langle\mathbf{P}_{\varepsilon}^{\beta}\phi_{i},\phi_{j}\right\rangle\end{bmatrix},
[Ujnε(0)Zjnε(0)]=[(U0ε)jn0].\displaystyle\begin{bmatrix}U_{jn}^{\varepsilon}\left(0\right)\\ Z_{jn}^{\varepsilon}\left(0\right)\end{bmatrix}=\begin{bmatrix}\left(U_{0}^{\varepsilon}\right)_{jn}\\ 0\end{bmatrix}.

Consider zjnε=[Ujnε,Zjnε]Tz_{jn}^{\varepsilon}=\left[U_{jn}^{\varepsilon},Z_{jn}^{\varepsilon}\right]^{\text{T}}. We thus obtain the following integral equation:

zjnε(x)=zjnε(0)+Aj0xzjnε(s)𝑑s+0xFj(zε)(s)𝑑s,\displaystyle z_{jn}^{\varepsilon}\left(x\right)=z_{jn}^{\varepsilon}\left(0\right)+A_{j}\int_{0}^{x}z_{jn}^{\varepsilon}\left(s\right)ds+\int_{0}^{x}F_{j}\left(z^{\varepsilon}\right)\left(s\right)ds, (13)

where

Aj=[01μj0],Fj(zε)=[0i=0nUinε𝐏εβϕi,ϕj].A_{j}=\begin{bmatrix}0&1\\ -\mu_{j}&0\end{bmatrix},\quad F_{j}\left(z^{\varepsilon}\right)=\begin{bmatrix}0\\ -\sum_{i=0}^{n}U_{in}^{\varepsilon}\left\langle\mathbf{P}_{\varepsilon}^{\beta}\phi_{i},\phi_{j}\right\rangle\end{bmatrix}.

Here, we mean zε=[z0nε,z1nε,znnε]2(n+1)z^{\varepsilon}=\left[z_{0n}^{\varepsilon},z_{1n}^{\varepsilon}\ldots,z_{nn}^{\varepsilon}\right]\in\mathbb{R}^{2(n+1)} and for simplicity, we, from now on, neglect the presence of nn in UjnεU_{jn}^{\varepsilon} and zjnεz_{jn}^{\varepsilon} in this proof. The integral equation (13) can be rewritten as zε(x)=H[zε](x)z^{\varepsilon}(x)=H[z^{\varepsilon}](x), where the same notation as zεz^{\varepsilon} is applied to HH with HjH_{j} being the right-hand side of (13).

Define the norm in Y=C([0,1];2(n+1))Y=C([0,1];\mathbb{R}^{2(n+1)}) as follows:

cY:=supx[0,1]j=0n|cj(x)|with c=[cj]2(n+1).\left\|c\right\|_{Y}:=\sup_{x\in\left[0,1\right]}\sum_{j=0}^{n}\left|c_{j}\left(x\right)\right|\quad\text{with }c=\left[c_{j}\right]\in\mathbb{R}^{2(n+1)}.

We then want to prove that there exists n0n_{0}\in\mathbb{N}^{*} such that the operator Hn0:=H[Hn01]:YYH^{n_{0}}:=H[H^{n_{0}-1}]:Y\to Y is a contraction mapping. In other words, we find K[0,1)K\in[0,1) such that

Hn0[z1ε]Hn0[z2ε]YKz1εz2εYfor any z1ε,z2εY.\left\|H^{n_{0}}\left[z_{1}^{\varepsilon}\right]-H^{n_{0}}\left[z_{2}^{\varepsilon}\right]\right\|_{Y}\leq K\left\|z_{1}^{\varepsilon}-z_{2}^{\varepsilon}\right\|_{Y}\quad\text{for any }z_{1}^{\varepsilon},z_{2}^{\varepsilon}\in Y.

This can be done by induction. Indeed, observe that

|Hj[z1ε](x)Hj[z2ε](x)|\displaystyle\left|H_{j}\left[z_{1}^{\varepsilon}\right]\left(x\right)-H_{j}\left[z_{2}^{\varepsilon}\right]\left(x\right)\right|
0x(μj2+1|z1jε(s)z2jε(s)|+C1log(γ(β))i=0n|U1iε(s)U2iε(s)|)𝑑s\displaystyle\leq\int_{0}^{x}\left(\sqrt{\mu_{j}^{2}+1}\left|z_{1j}^{\varepsilon}\left(s\right)-z_{2j}^{\varepsilon}\left(s\right)\right|+C_{1}\log\left(\gamma\left(\beta\right)\right)\sum_{i=0}^{n}\left|U_{1i}^{\varepsilon}\left(s\right)-U_{2i}^{\varepsilon}\left(s\right)\right|\right)ds
(μj2+1n+1+CC1log(γ(β)))xz1εz2εY,\displaystyle\leq\left(\frac{\sqrt{\mu_{j}^{2}+1}}{n+1}+CC_{1}\log\left(\gamma\left(\beta\right)\right)\right)x\left\|z_{1}^{\varepsilon}-z_{2}^{\varepsilon}\right\|_{Y},

aided by the conditional estimate (4). Here, we indicate C=C(ϕnH01(0,1))>0C=C\left(\left\|\phi_{n}\right\|_{H_{0}^{1}(0,1)}\right)>0. Therefore, it is immediate to prove that

|Hjn[z1ε](x)Hjn[z2ε](x)|[μj2+1n+1+CC1log(γ(β))]nxnn!z1εz2εY,\left|H_{j}^{n}\left[z_{1}^{\varepsilon}\right]\left(x\right)-H_{j}^{n}\left[z_{2}^{\varepsilon}\right]\left(x\right)\right|\leq\left[\frac{\sqrt{\mu_{j}^{2}+1}}{n+1}+CC_{1}\log\left(\gamma\left(\beta\right)\right)\right]^{n}\frac{x^{n}}{n!}\left\|z_{1}^{\varepsilon}-z_{2}^{\varepsilon}\right\|_{Y},

which leads to the fact that

Hn[z1ε]Hn[z2ε]Yz1εz2εYn!j=0n[μj2+1n+1+CC1log(γ(β))]n.\left\|H^{n}\left[z_{1}^{\varepsilon}\right]-H^{n}\left[z_{2}^{\varepsilon}\right]\right\|_{Y}\leq\frac{\left\|z_{1}^{\varepsilon}-z_{2}^{\varepsilon}\right\|_{Y}}{n!}\sum_{j=0}^{n}\left[\frac{\sqrt{\mu_{j}^{2}+1}}{n+1}+CC_{1}\log\left(\gamma\left(\beta\right)\right)\right]^{n}.

In view of the fact that

limn1n!j=0n[μj2+1n+1+CC1log(γ(β))]n=0,\lim_{n\to\infty}\frac{1}{n!}\sum_{j=0}^{n}\left[\frac{\sqrt{\mu_{j}^{2}+1}}{n+1}+CC_{1}\log\left(\gamma\left(\beta\right)\right)\right]^{n}=0,

we can find a sufficiently large n0n_{0} such that

1n0!j=0n0[μj2+1n0+1+CC1log(γ(β))]n0<1.\frac{1}{n_{0}!}\sum_{j=0}^{n_{0}}\left[\frac{\sqrt{\mu_{j}^{2}+1}}{n_{0}+1}+CC_{1}\log\left(\gamma\left(\beta\right)\right)\right]^{n_{0}}<1.

This indicates the existence of K[0,1)K\in[0,1) and that Hn0H^{n_{0}} is a contraction mapping from YY onto itself. By the Banach fixed-point argument, there exists a unique solution zεYz^{\varepsilon}\in Y such that Hn0[zε]=zεH^{n_{0}}[z^{\varepsilon}]=z^{\varepsilon}. Since Hn0[H[zε]]=H[Hn0[zε]]=H[zε]H^{n_{0}}[H[z^{\varepsilon}]]=H[H^{n_{0}}[z^{\varepsilon}]]=H[z^{\varepsilon}], then the integral equation H[zε]=zεH[z^{\varepsilon}]=z^{\varepsilon} admits a unique solution in YY. Hence, we complete the proof of the lemma. ∎

Now we can state the existence result for (5)–(6) in the following theorem.

Theorem 3.3.

Assume (2) holds. For each ε>0\varepsilon>0, the regularized system (5)–(6) admits a weak solution uβεu_{\beta}^{\varepsilon} in the sense of Definition 3.1. Moreover, it holds that uβεC([0,1];H01(0,1))u_{\beta}^{\varepsilon}\in C([0,1];H_{0}^{1}(0,1)) and xuβεC([0,1];L2(0,1))\partial_{x}u_{\beta}^{\varepsilon}\in C([0,1];L^{2}(0,1)).

Proof.

To prove this theorem, we need to derive some energy estimates for approximate solutions unεu_{n}^{\varepsilon}. Thanks to Lemma 3.2, we can prove that xunεC([0,1];𝕊n)\partial_{x}u_{n}^{\varepsilon}\in C([0,1];\mathbb{S}_{n}). Taking in (9) ψ=xunε\psi=\partial_{x}u_{n}^{\varepsilon}, we find that

ddx[xunε2+yunε2]\displaystyle\frac{d}{dx}\left[\left\|\partial_{x}u_{n}^{\varepsilon}\right\|^{2}+\left\|\partial_{y}u_{n}^{\varepsilon}\right\|^{2}\right] =2𝐏εβunε,xunε\displaystyle=-2\left\langle\mathbf{P}_{\varepsilon}^{\beta}u_{n}^{\varepsilon},\partial_{x}u_{n}^{\varepsilon}\right\rangle
C1log(γ(β))unεH1(0,1)2+C1log(γ(β))xunε2.\displaystyle\leq C_{1}\log\left(\gamma\left(\beta\right)\right)\left\|u_{n}^{\varepsilon}\right\|_{H^{1}\left(0,1\right)}^{2}+C_{1}\log\left(\gamma\left(\beta\right)\right)\left\|\partial_{x}u_{n}^{\varepsilon}\right\|^{2}. (14)

The norms unεH01(0,1)=yunε\left\|u_{n}^{\varepsilon}\right\|_{H_{0}^{1}(0,1)}=\left\|\partial_{y}u_{n}^{\varepsilon}\right\| and unεH1(0,1)\left\|u_{n}^{\varepsilon}\right\|_{H^{1}(0,1)} are equivalent333By the Poincaré inequality, we particularly have 12unεH1(0,1)unεH01(0,1)unεH1(0,1)\frac{1}{\sqrt{2}}\left\|u_{n}^{\varepsilon}\right\|_{H^{1}(0,1)}\leq\left\|u_{n}^{\varepsilon}\right\|_{H_{0}^{1}(0,1)}\leq\left\|u_{n}^{\varepsilon}\right\|_{H^{1}(0,1)}. based upon the zero trace. Thereupon, by integrating the estimate (14) with respect to xx, we arrive at

xunε(x,)2+unε(s,)H01(0,1)2\displaystyle\left\|\partial_{x}u_{n}^{\varepsilon}\left(x,\cdot\right)\right\|^{2}+\left\|u_{n}^{\varepsilon}\left(s,\cdot\right)\right\|_{H_{0}^{1}\left(0,1\right)}^{2}
yunε(0,)2+2C1log(γ(β))0x(unε(s,)H01(0,1)2+xunε(s,)2)𝑑s.\displaystyle\leq\left\|\partial_{y}u_{n}^{\varepsilon}\left(0,\cdot\right)\right\|^{2}+2C_{1}\log\left(\gamma\left(\beta\right)\right)\int_{0}^{x}\left(\left\|u_{n}^{\varepsilon}\left(s,\cdot\right)\right\|_{H_{0}^{1}\left(0,1\right)}^{2}+\left\|\partial_{x}u_{n}^{\varepsilon}\left(s,\cdot\right)\right\|^{2}\right)ds.

Using Gronwall’s inequality, we get

xunε(x,)2+unε(x,)H01(0,1)2yunε(0,)2γ2C1x(β).\displaystyle\left\|\partial_{x}u_{n}^{\varepsilon}\left(x,\cdot\right)\right\|^{2}+\left\|u_{n}^{\varepsilon}\left(x,\cdot\right)\right\|_{H_{0}^{1}\left(0,1\right)}^{2}\leq\left\|\partial_{y}u_{n}^{\varepsilon}\left(0,\cdot\right)\right\|^{2}\gamma^{2C_{1}x}(\beta). (15)

In view of (10), it is straightforward to see that yunε(0,)C\left\|\partial_{y}u_{n}^{\varepsilon}\left(0,\cdot\right)\right\|\leq C for any ε>0\varepsilon>0. Thus, for any nn\in\mathbb{N} we have

γC1(β)unεis bounded in L(0,1;H01(0,1)),\displaystyle\gamma^{-C_{1}}\left(\beta\right)u_{n}^{\varepsilon}\;\text{is bounded in }L^{\infty}\left(0,1;H_{0}^{1}\left(0,1\right)\right),
γC1(β)xunεis bounded in L(0,1;L2(0,1)).\displaystyle\gamma^{-C_{1}}\left(\beta\right)\partial_{x}u_{n}^{\varepsilon}\;\text{is bounded in }L^{\infty}\left(0,1;L^{2}\left(0,1\right)\right).

It follows from the Banach–Alaoglu theorem, and the argument that a weak limit of derivatives is the derivative of the weak limit, that we can extract a subsequence of scaled approximate solutions γC1(β)unε\gamma^{-C_{1}}\left(\beta\right)u_{n}^{\varepsilon}, which we still denote by {γC1(β)unε}n\left\{\gamma^{-C_{1}}\left(\beta\right)u_{n}^{\varepsilon}\right\}_{n\in\mathbb{N}}, such that for each ε>0\varepsilon>0,

γC1(β)unεγC1(β)uεweaklyin L(0,1;H01(0,1)),\displaystyle\gamma^{-C_{1}}\left(\beta\right)u_{n}^{\varepsilon}\to\gamma^{-C_{1}}\left(\beta\right)u^{\varepsilon}\;\text{weakly}-*\;\text{in }L^{\infty}\left(0,1;H_{0}^{1}\left(0,1\right)\right), (16)
γC1(β)xunεγC1(β)xuεweaklyin L(0,1;L2(0,1)).\displaystyle\gamma^{-C_{1}}\left(\beta\right)\partial_{x}u_{n}^{\varepsilon}\to\gamma^{-C_{1}}\left(\beta\right)\partial_{x}u^{\varepsilon}\;\text{weakly}-*\;\text{in }L^{\infty}\left(0,1;L^{2}\left(0,1\right)\right). (17)

We remark that (16) holds if and only if

01γC1(β)unε(x,),w𝑑x01γC1(β)uε(x,),w𝑑xfor wL1(0,1;H1(0,1)),\int_{0}^{1}\left\langle\gamma^{-C_{1}}\left(\beta\right)u_{n}^{\varepsilon}\left(x,\cdot\right),w\right\rangle dx\to\int_{0}^{1}\left\langle\gamma^{-C_{1}}\left(\beta\right)u^{\varepsilon}\left(x,\cdot\right),w\right\rangle dx\quad\text{for }w\in L^{1}\left(0,1;H^{-1}\left(0,1\right)\right),

which implies that

unεuεweaklyin L(0,1;H01(0,1)).\displaystyle u_{n}^{\varepsilon}\to u^{\varepsilon}\;\text{weakly}-*\;\text{in }L^{\infty}\left(0,1;H_{0}^{1}\left(0,1\right)\right). (18)

The same result is obtained for the weak-star convergence in (17), viz.

xunεxuεweaklyin L(0,1;L2(0,1)).\displaystyle\partial_{x}u_{n}^{\varepsilon}\to\partial_{x}u^{\varepsilon}\;\text{weakly}-*\;\text{in }L^{\infty}\left(0,1;L^{2}\left(0,1\right)\right). (19)

Now, using the Galerkin equation (9) we can show that x22unεL2(0,1;𝕊n)\partial_{x^{2}}^{2}u_{n}^{\varepsilon}\in L^{2}(0,1;\mathbb{S}_{n}). Thus, we obtain the uniform bound with respect to nn of x22unε\partial_{x^{2}}^{2}u_{n}^{\varepsilon} as follows:

γC1(β)x22unε(x,)H1(0,1)=supψH1(0,1)\{0}γC1(β)x22unε(x,),ψψH01(0,1)\displaystyle\gamma^{-C_{1}}\left(\beta\right)\left\|\partial_{x^{2}}^{2}u_{n}^{\varepsilon}\left(x,\cdot\right)\right\|_{H^{-1}\left(0,1\right)}=\sup_{\psi\in H^{1}\left(0,1\right)\backslash\left\{0\right\}}\frac{\gamma^{-C_{1}}\left(\beta\right)\left\langle\partial_{x^{2}}^{2}u_{n}^{\varepsilon}\left(x,\cdot\right),\psi\right\rangle}{\left\|\psi\right\|_{H_{0}^{1}\left(0,1\right)}}
=supψH1(0,1)\{0}γC1(β)yunε(x,),yψγC1(β)𝐏εβunε(x,),yψψH01(0,1)\displaystyle=\sup_{\psi\in H^{1}\left(0,1\right)\backslash\left\{0\right\}}\frac{-\gamma^{-C_{1}}\left(\beta\right)\left\langle\partial_{y}u_{n}^{\varepsilon}\left(x,\cdot\right),\partial_{y}\psi\right\rangle-\gamma^{-C_{1}}\left(\beta\right)\left\langle\mathbf{P}_{\varepsilon}^{\beta}u_{n}^{\varepsilon}\left(x,\cdot\right),\partial_{y}\psi\right\rangle}{\left\|\psi\right\|_{H_{0}^{1}\left(0,1\right)}}
CγC1(β)unε(x,)H01(0,1).\displaystyle\leq C\gamma^{-C_{1}}\left(\beta\right)\left\|u_{n}^{\varepsilon}\left(x,\cdot\right)\right\|_{H_{0}^{1}\left(0,1\right)}.

Henceforth, by squaring this estimate, integrating the resulting with respect to xx and then using the Banach–Alaoglu theorem we can choose the subsequence of γC1(β)unε\gamma^{-C_{1}}\left(\beta\right)u_{n}^{\varepsilon} so that

γC1(β)x22unεγC1(β)x22uεweakly in L2(0,1;H1(0,1)),\gamma^{-C_{1}}\left(\beta\right)\partial_{x^{2}}^{2}u_{n}^{\varepsilon}\to\gamma^{-C_{1}}\left(\beta\right)\partial_{x^{2}}^{2}u^{\varepsilon}\;\text{weakly in }L^{2}\left(0,1;H^{-1}\left(0,1\right)\right),

which leads to

x22unεx22uεweakly in L2(0,1;H1(0,1)).\partial_{x^{2}}^{2}u_{n}^{\varepsilon}\to\partial_{x^{2}}^{2}u^{\varepsilon}\;\text{weakly in }L^{2}\left(0,1;H^{-1}\left(0,1\right)\right).

Combining the above weak-star and weak limits, the function uεu^{\varepsilon} satisfies

uεL(0,1;H01(0,1)),xuεL(0,1;L2(0,1)),x22uεL2(0,1;H1(0,1)).\displaystyle u^{\varepsilon}\in L^{\infty}\left(0,1;H_{0}^{1}\left(0,1\right)\right),\quad\partial_{x}u^{\varepsilon}\in L^{\infty}\left(0,1;L^{2}\left(0,1\right)\right),\quad\partial_{x^{2}}^{2}u^{\varepsilon}\in L^{2}\left(0,1;H^{-1}\left(0,1\right)\right). (20)

Furthermore, using the Aubin–Lions lemma in combination with the Rellich–Kondrachov theorem H01(0,1)L2(0,1)H_{0}^{1}(0,1)\subset L^{2}(0,1) for (18) and (19), we get

unεuεstronglyin L2(0,1;H01(0,1)).\displaystyle u_{n}^{\varepsilon}\to u^{\varepsilon}\;\text{strongly}\;\text{in }L^{2}\left(0,1;H_{0}^{1}\left(0,1\right)\right). (21)

Now we multiply both sides of the Galerkin equation (9) by an xx-dependent test function w~Cc(0,1)\tilde{w}\in C_{c}^{\infty}(0,1), then by integrating the resulting equation with respect to xx we arrive at

012x2unε,υ𝑑x+01yunε,yυ𝑑x+01𝐏εβunε,υ𝑑x=0,\int_{0}^{1}\left\langle\frac{\partial^{2}}{\partial x^{2}}u_{n}^{\varepsilon},\upsilon\right\rangle dx+\int_{0}^{1}\left\langle\partial_{y}u_{n}^{\varepsilon},\partial_{y}\upsilon\right\rangle dx+\int_{0}^{1}\left\langle\mathbf{P}_{\varepsilon}^{\beta}u_{n}^{\varepsilon},\upsilon\right\rangle dx=0,

where we have denoted by υ=υ(x,y)=w~(x)ψ(y)\upsilon=\upsilon(x,y)=\tilde{w}(x)\psi(y) for ψ𝕊n\psi\in\mathbb{S}_{n}. Hence, passing the limit of this equation as nn\to\infty we obtain

012x2uε,υ𝑑x+01yuε,yυ𝑑x+01𝐏εβuε,υ𝑑x=0,\displaystyle\int_{0}^{1}\left\langle\frac{\partial^{2}}{\partial x^{2}}u^{\varepsilon},\upsilon\right\rangle dx+\int_{0}^{1}\left\langle\partial_{y}u^{\varepsilon},\partial_{y}\upsilon\right\rangle dx+\int_{0}^{1}\left\langle\mathbf{P}_{\varepsilon}^{\beta}u^{\varepsilon},\upsilon\right\rangle dx=0, (22)

where convergence of the second and third terms is guaranteed by (21). Equation (22) holds for v=w~ψv=\tilde{w}\psi with ψH01(0,1)\psi\in H_{0}^{1}(0,1) and since w~Cc(0,1)\tilde{w}\in C_{c}^{\infty}(0,1) is arbitrary, we deduce that the function uεu^{\varepsilon} obtained from approximate solutions unεu_{n}^{\varepsilon} satisfies the weak formulation (7) for every test function ψH01(0,1)\psi\in H_{0}^{1}(0,1). In addition, the arguments in (20) enable us to show that

uεC([0,1];H01(0,1)),xuεC([0,1];L2(0,1)),\displaystyle u^{\varepsilon}\in C([0,1];H_{0}^{1}(0,1)),\quad\partial_{x}u^{\varepsilon}\in C([0,1];L^{2}(0,1)), (23)

where we have applied the Aubin–Lions lemma. Note that for the latter argument in (23), we rely on the Gelfand triple H01(0,1)L2(0,1)H1(0,1)H_{0}^{1}(0,1)\subset L^{2}(0,1)\subset H^{-1}(0,1).

It now remains to verify the initial data. Take κC1([0,1])\kappa\in C^{1}([0,1]) satisfying κ(0)=1\kappa(0)=1 and κ(1)=0\kappa(1)=0. It follows from (19) that

01xunε,ψκ(x)𝑑x01xuε,ψκ(x)𝑑xfor ψL2(0,1).\int_{0}^{1}\left\langle\partial_{x}u_{n}^{\varepsilon},\psi\right\rangle\kappa\left(x\right)dx\to\int_{0}^{1}\left\langle\partial_{x}u^{\varepsilon},\psi\right\rangle\kappa\left(x\right)dx\quad\text{for }\psi\in L^{2}\left(0,1\right).

Then by integration by parts, we have

01unε,ψκx𝑑xunε(0),ψκ(0)01uε,ψκx𝑑xuε(0),ψκ(0)-\int_{0}^{1}\left\langle u_{n}^{\varepsilon},\psi\right\rangle\kappa_{x}dx-\left\langle u_{n}^{\varepsilon}\left(0\right),\psi\right\rangle\kappa\left(0\right)\to-\int_{0}^{1}\left\langle u^{\varepsilon},\psi\right\rangle\kappa_{x}dx-\left\langle u^{\varepsilon}\left(0\right),\psi\right\rangle\kappa\left(0\right)

and thereupon, we get unε(0),ψuε(0),ψ\left\langle u_{n}^{\varepsilon}\left(0\right),\psi\right\rangle\to\left\langle u^{\varepsilon}\left(0\right),\psi\right\rangle for all ψH01(0,1)\psi\in H_{0}^{1}(0,1) by virtue of (18). Cf. (10) for the strong convergence of unε(0)u_{n}^{\varepsilon}(0) in H1(0,1)H^{1}(0,1), we obtain unε(0),ψu0ε,ψ\left\langle u_{n}^{\varepsilon}\left(0\right),\psi\right\rangle\to\left\langle u_{0}^{\varepsilon},\psi\right\rangle for all ψH1(0,1)\psi\in H^{1}(0,1). Hence, uε(0),ψ=u0ε,ψ\left\langle u^{\varepsilon}\left(0\right),\psi\right\rangle=\left\langle u_{0}^{\varepsilon},\psi\right\rangle for all ψH1(0,1)\psi\in H^{1}(0,1), which implies that uε(0)=u0εu^{\varepsilon}\left(0\right)=u_{0}^{\varepsilon} for a.e. in (0,1)(0,1). We complete the proof of the theorem. ∎

Now we close this section by uniqueness of the weak solution of (5)–(6). Proof of this result is still ended up with the use of some energy estimates and with exploiting the Gronwall inequality, eventually.

Theorem 3.4.

Assume (2) holds. For each ε>0\varepsilon>0, the regularized system (5)–(6) admits a unique weak solution uβεu_{\beta}^{\varepsilon} in the sense of Definition 3.1.

Proof.

We sketch out some important steps because this proof is standard. Indeed, let u1εu_{1}^{\varepsilon} and u2εu_{2}^{\varepsilon} be two weak solutions that we have obtained in Theorem 3.3. We can see that the function dε=u1εu2εd^{\varepsilon}=u_{1}^{\varepsilon}-u_{2}^{\varepsilon} satisfies a linear wave equation with zero data (dε=xdε=0d^{\varepsilon}=\partial_{x}d^{\varepsilon}=0) by virtue of the linearity of 𝐏βε\mathbf{P}_{\beta}^{\varepsilon}. Similar to (7), the equation for dεC([0,1];H01(0,1))d^{\varepsilon}\in C([0,1];H_{0}^{1}(0,1)) reads as

2x2dε,ψ+ydε,yψ+𝐏εβdε,ψ=0.\left\langle\frac{\partial^{2}}{\partial x^{2}}d^{\varepsilon},\psi\right\rangle+\left\langle\partial_{y}d^{\varepsilon},\partial_{y}\psi\right\rangle+\left\langle\mathbf{P}_{\varepsilon}^{\beta}d^{\varepsilon},\psi\right\rangle=0.

Then taking ψ=xdε\psi=\partial_{x}d^{\varepsilon}, we proceed as in the way to get the estimate (15) in proof of Theorem 3.3. Hence, dε=0d^{\varepsilon}=0 a.e. in (0,1)(0,1) because of the fact that

xdε(x,)2+dε(x,)H01(0,1)20.\left\|\partial_{x}d^{\varepsilon}\left(x,\cdot\right)\right\|^{2}+\left\|d^{\varepsilon}\left(x,\cdot\right)\right\|_{H_{0}^{1}\left(0,1\right)}^{2}\leq 0.

This completes the proof of the theorem. ∎

4 Convergence analysis

In this section, we are concerned about the rate of convergence of the regularized problem uβεu_{\beta}^{\varepsilon} towards a certain true solution uu. The estimate is pointwise in ε\varepsilon. We come up with the error estimation by using a Carleman weight function when looking for a scaled difference uβεuu_{\beta}^{\varepsilon}-u. This way is different from the proof of weak solvability of the regularized system (5)–(6) because of the presence of the perturbing operator acting on the true solution and of the large conditional estimate of the stabilized operator in the difference formulation. It is worth mentioning that the scaled difference equation is essentially a damped wave equation, where the damped terms play a crucial role in controlling all involved large quantities. The original notion behind the choice of Carleman weights for convergence analysis in the field of inverse and ill-posed problems is the designation of a strictly convex cost functional, which is usually called convexification; see e.g. the monograph [19] for a complete background of this method. In this work, we choose an exponentially decreasing weight such that it “maximizes” the presence of initial data u0εu_{0}^{\varepsilon}.

Theorem 4.1.

Assume (2) holds. Let uC([0,1];𝕎)u\in C([0,1];\mathbb{W}) be a unique solution of the Laplace system (1), where the space 𝕎\mathbb{W} is obtained by the perturbation 𝐐εβ\mathbf{Q}_{\varepsilon}^{\beta} in Definition 2.1. Let uβεu_{\beta}^{\varepsilon} be a unique weak solution of the regularized system (5)–(6) analyzed in Theorems 3.33.4. By choosing γ(β)1\gamma(\beta)\geq 1 such that 2C1log(γ(β))/3>12C_{1}\log\left(\gamma\left(\beta\right)\right)/3>1, the following error estimate holds:

xuβε(x,)xu(x,)2+uβε(x,)u(x,)H1(0,1)2\displaystyle\left\|\partial_{x}u_{\beta}^{\varepsilon}\left(x,\cdot\right)-\partial_{x}u\left(x,\cdot\right)\right\|^{2}+\left\|u_{\beta}^{\varepsilon}\left(x,\cdot\right)-u\left(x,\cdot\right)\right\|_{H^{1}(0,1)}^{2} (24)
43C12[4ε2+278C02C13γ2(β)log3(γ(β))xuC([0,1];𝕎)2]log2(γ(β))γ7C1x/3(β).\displaystyle\leq\frac{4}{3}C_{1}^{2}\left[4\varepsilon^{2}+\frac{27}{8}C_{0}^{2}C_{1}^{-3}\gamma^{-2}\left(\beta\right)\log^{-3}(\gamma(\beta))x\left\|u\right\|_{C\left([0,1];\mathbb{W}\right)}^{2}\right]\log^{2}\left(\gamma\left(\beta\right)\right)\gamma^{7C_{1}x/3}(\beta).
Proof.

Let w=(uβεu)exp(ρβx)w=\left(u_{\beta}^{\varepsilon}-u\right)\exp\left(-\rho_{\beta}x\right) for ρβ>0\rho_{\beta}>0 being chosen later. From (5) and (1), we compute the following difference equation:

wxxwyy+2ρβwx+ρβ2w=𝐏εβw+eρβx𝐐εβuin (0,1)×(0,1).\displaystyle w_{xx}-w_{yy}+2\rho_{\beta}w_{x}+\rho_{\beta}^{2}w=-\mathbf{P}_{\varepsilon}^{\beta}w+e^{-\rho_{\beta}x}\mathbf{Q}_{\varepsilon}^{\beta}u\quad\text{in }(0,1)\times(0,1). (25)

This equation is associated with the Dirichlet boundary condition and the initial conditions:

{w(x,0)=w(x,1)=0for x[0,1],w(0,y)=u0ε(y)u0(y),xw(0,y)=ρβw(0,y)for y[0,1].\displaystyle\begin{cases}w\left(x,0\right)=w\left(x,1\right)=0&\text{for }x\in[0,1],\\ w\left(0,y\right)=u_{0}^{\varepsilon}\left(y\right)-u_{0}(y),\partial_{x}w\left(0,y\right)=-\rho_{\beta}w(0,y)&\text{for }y\in[0,1].\end{cases} (26)

Multiplying both sides of (25) by wxw_{x} and integrating the resulting equation with respect to yy from 0 to 1, we have

ddxwx(x,)2\displaystyle\frac{d}{dx}\left\|w_{x}\left(x,\cdot\right)\right\|^{2} +ddxwy(x,)2+2ρβ2ddxw(x,)2+4ρβwx(x,)2\displaystyle+\frac{d}{dx}\left\|w_{y}\left(x,\cdot\right)\right\|^{2}+2\rho_{\beta}^{2}\frac{d}{dx}\left\|w\left(x,\cdot\right)\right\|^{2}+4\rho_{\beta}\left\|w_{x}\left(x,\cdot\right)\right\|^{2}
=2eρβx𝐐εβu,wx2𝐏εβw,wx=:I1+I2.\displaystyle=2e^{-\rho_{\beta}x}\left\langle\mathbf{Q}_{\varepsilon}^{\beta}u,w_{x}\right\rangle-2\left\langle\mathbf{P}_{\varepsilon}^{\beta}w,w_{x}\right\rangle=:I_{1}+I_{2}. (27)

Using the conditional estimates (3), (4) and applying the Cauchy–Schwarz inequality, we estimate I1I_{1} and I2I_{2} in (27) in the following manner:

I1\displaystyle I_{1} C02e2ρβxγ2(β)ρβ1u(x,)𝕎2+ρβwx(x,)2,\displaystyle\leq C_{0}^{2}e^{-2\rho_{\beta}x}\gamma^{-2}\left(\beta\right)\rho_{\beta}^{-1}\left\|u\left(x,\cdot\right)\right\|_{\mathbb{W}}^{2}+\rho_{\beta}\left\|w_{x}\left(x,\cdot\right)\right\|^{2},
I2\displaystyle I_{2} 2C1log(γ(β))(w(x,)wx(x,)+wy(x,)wx(x,))\displaystyle\leq 2C_{1}\log\left(\gamma(\beta)\right)\left(\left\|w\left(x,\cdot\right)\right\|\left\|w_{x}\left(x,\cdot\right)\right\|+\left\|w_{y}\left(x,\cdot\right)\right\|\left\|w_{x}\left(x,\cdot\right)\right\|\right)
C1log(γ(β))(w(x,)2+wy(x,)2)+2C1log(γ(β))wx(x,)2.\displaystyle\leq C_{1}\log\left(\gamma\left(\beta\right)\right)\left(\left\|w\left(x,\cdot\right)\right\|^{2}+\left\|w_{y}\left(x,\cdot\right)\right\|^{2}\right)+2C_{1}\log\left(\gamma\left(\beta\right)\right)\left\|w_{x}\left(x,\cdot\right)\right\|^{2}.

Henceforth, the left-hand side of (27) is bounded by

ρβ2ddxwx(x,)2+ρβ2ddxwy(x,)2+2ddxw(x,)2\displaystyle\rho_{\beta}^{-2}\frac{d}{dx}\left\|w_{x}\left(x,\cdot\right)\right\|^{2}+\rho_{\beta}^{-2}\frac{d}{dx}\left\|w_{y}\left(x,\cdot\right)\right\|^{2}+2\frac{d}{dx}\left\|w\left(x,\cdot\right)\right\|^{2}
C02e2ρβxγ2(β)ρβ3u(x,)𝕎2+C1log(γ(β))ρβ2(w(x,)2+wy(x,)2)\displaystyle\leq C_{0}^{2}e^{-2\rho_{\beta}x}\gamma^{-2}\left(\beta\right)\rho_{\beta}^{-3}\left\|u\left(x,\cdot\right)\right\|_{\mathbb{W}}^{2}+C_{1}\log\left(\gamma\left(\beta\right)\right)\rho_{\beta}^{-2}\left(\left\|w\left(x,\cdot\right)\right\|^{2}+\left\|w_{y}\left(x,\cdot\right)\right\|^{2}\right)
+ρβ2(2C1log(γ(β))3ρβ)wx(x,)2.\displaystyle+\rho_{\beta}^{-2}\left(2C_{1}\log\left(\gamma\left(\beta\right)\right)-3\rho_{\beta}\right)\left\|w_{x}\left(x,\cdot\right)\right\|^{2}. (28)

Now, choosing in (28) that ρβ=2C1log(γ(β))/3>1\rho_{\beta}=2C_{1}\log\left(\gamma\left(\beta\right)\right)/3>1, integrating the resulting estimate with respect to xx from 0 to x1x_{1}, we get

ρβ2(wx(x1,)2+wy(x1,)2)+w(x1,)2\displaystyle\rho_{\beta}^{-2}\left(\left\|w_{x}\left(x_{1},\cdot\right)\right\|^{2}+\left\|w_{y}\left(x_{1},\cdot\right)\right\|^{2}\right)+\left\|w\left(x_{1},\cdot\right)\right\|^{2}
ρβ2(wx(0,)2+wy(0,)2)+2w(0,)2+C02γ2(β)ρβ30x1e2ρβxu(x,)𝕎2\displaystyle\leq\rho_{\beta}^{-2}\left(\left\|w_{x}\left(0,\cdot\right)\right\|^{2}+\left\|w_{y}\left(0,\cdot\right)\right\|^{2}\right)+2\left\|w\left(0,\cdot\right)\right\|^{2}+C_{0}^{2}\gamma^{-2}\left(\beta\right)\rho_{\beta}^{-3}\int_{0}^{x_{1}}e^{-2\rho_{\beta}x}\left\|u\left(x,\cdot\right)\right\|_{\mathbb{W}}^{2}
+C1log(γ(β))0x1(w(x,)2+ρβ2wy(x,)2)𝑑x.\displaystyle+C_{1}\log(\gamma(\beta))\int_{0}^{x_{1}}\left(\left\|w\left(x,\cdot\right)\right\|^{2}+\rho_{\beta}^{-2}\left\|w_{y}\left(x,\cdot\right)\right\|^{2}\right)dx. (29)

Due to (2), we estimate that

ρβ2(wx(0,)2+wy(0,)2)+2w(0,)2=3u0εu02+ρβ2yu0εyu024ε2.\displaystyle\rho_{\beta}^{-2}\left(\left\|w_{x}\left(0,\cdot\right)\right\|^{2}+\left\|w_{y}\left(0,\cdot\right)\right\|^{2}\right)+2\left\|w\left(0,\cdot\right)\right\|^{2}=3\left\|u_{0}^{\varepsilon}-u_{0}\right\|^{2}+\rho_{\beta}^{-2}\left\|\partial_{y}u_{0}^{\varepsilon}-\partial_{y}u_{0}\right\|^{2}\leq 4\varepsilon^{2}. (30)

Thus, using Gronwall’s inequality we continue to estimate (29) as follows:

ρβ2(wx(x1,)2+wy(x1,)2)+w(x1,)2[4ε2+C02γ2(β)ρβ3x1uC([0,1];𝕎)2]γC1x1(β).\displaystyle\rho_{\beta}^{-2}\left(\left\|w_{x}\left(x_{1},\cdot\right)\right\|^{2}+\left\|w_{y}\left(x_{1},\cdot\right)\right\|^{2}\right)+\left\|w\left(x_{1},\cdot\right)\right\|^{2}\leq\left[4\varepsilon^{2}+C_{0}^{2}\gamma^{-2}\left(\beta\right)\rho_{\beta}^{-3}x_{1}\left\|u\right\|_{C\left([0,1];\mathbb{W}\right)}^{2}\right]\gamma^{C_{1}x_{1}}(\beta). (31)

It is now to use back-substitutions to get back the difference uβεuu_{\beta}^{\varepsilon}-u. In view of the fact that

ρβ2e2ρβx1xuβε(x1,)xu(x1,)2\displaystyle\rho_{\beta}^{-2}e^{-2\rho_{\beta}x_{1}}\left\|\partial_{x}u_{\beta}^{\varepsilon}\left(x_{1},\cdot\right)-\partial_{x}u\left(x_{1},\cdot\right)\right\|^{2} ρβ2(ρβw(x1,)+wx(x1,)2)2\displaystyle\leq\rho_{\beta}^{-2}\left(\rho_{\beta}\left\|w\left(x_{1},\cdot\right)\right\|+\left\|w_{x}\left(x_{1},\cdot\right)\right\|^{2}\right)^{2}
2(w(x1,)2+ρβ2wx(x1,)2),\displaystyle\leq 2\left(\left\|w\left(x_{1},\cdot\right)\right\|^{2}+\rho_{\beta}^{-2}\left\|w_{x}\left(x_{1},\cdot\right)\right\|^{2}\right),

we obtain the aimed estimate (24). Hence, we complete the proof of the theorem. ∎

Without loss of generality, we now assume that C1<6/7C_{1}<6/7 for our convergence analysis below. This assumption on C1C_{1} is somehow similar to what has been assumed in [16] for the parabolic case, where C1C_{1} and the time domain are related to each other. In subsection 5.1 below, we show a way to control the largeness of such C1C_{1} in the framework of the truncation projection. Besides, the quantity C1C_{1} can be largely selected due to the flexible choice of γ(β)\gamma(\beta).

Theorem 4.2.

Under the assumptions of Theorem 4.1, if one has C1<6/7C_{1}<6/7 and chooses γ=γ(β)1\gamma=\gamma(\beta)\geq 1 such that there holds

limε0εγ(β)log(γ(β))=K1(0,),\displaystyle\lim_{\varepsilon\to 0}\varepsilon\gamma\left(\beta\right)\log\left(\gamma\left(\beta\right)\right)=K_{1}\in\left(0,\infty\right), (32)

then the following error estimate holds

xuβε(x,)xu(x,)2+uβε(x,)u(x,)H1(0,1)2Cγ7C1x32(β).\displaystyle\left\|\partial_{x}u_{\beta}^{\varepsilon}\left(x,\cdot\right)-\partial_{x}u\left(x,\cdot\right)\right\|^{2}+\left\|u_{\beta}^{\varepsilon}\left(x,\cdot\right)-u\left(x,\cdot\right)\right\|_{H^{1}\left(0,1\right)}^{2}\leq C\gamma^{\frac{7C_{1}x}{3}-2}(\beta). (33)

Consequently, if instead of (32) we choose a stronger choice

limε0εγ(β)=K2(0,),\displaystyle\lim_{\varepsilon\to 0}\varepsilon\gamma\left(\beta\right)=K_{2}\in\left(0,\infty\right), (34)

then it holds that

uβε(x,)u(x,)2Cγ7C1x32(β).\displaystyle\left\|u_{\beta}^{\varepsilon}\left(x,\cdot\right)-u\left(x,\cdot\right)\right\|^{2}\leq C\gamma^{\frac{7C_{1}x}{3}-2}(\beta). (35)
Proof.

As a by-product of the rigorous estimate (24), if the choice (32) holds, one can deduce that

xuβε(x,)xu(x,)2+uβε(x,)u(x,)H1(0,1)2\displaystyle\left\|\partial_{x}u_{\beta}^{\varepsilon}\left(x,\cdot\right)-\partial_{x}u\left(x,\cdot\right)\right\|^{2}+\left\|u_{\beta}^{\varepsilon}\left(x,\cdot\right)-u\left(x,\cdot\right)\right\|_{H^{1}\left(0,1\right)}^{2}
43C12[4ε2log2(γ(β))γ2(β)+278C02C13log1(γ(β))xuC([0,1];𝕎)2]γ7C1x32(β)\displaystyle\leq\frac{4}{3}C_{1}^{2}\left[4\varepsilon^{2}\log^{2}\left(\gamma\left(\beta\right)\right)\gamma^{2}\left(\beta\right)+\frac{27}{8}C_{0}^{2}C_{1}^{-3}\log^{-1}\left(\gamma\left(\beta\right)\right)x\left\|u\right\|_{C\left([0,1];\mathbb{W}\right)}^{2}\right]\gamma^{\frac{7C_{1}x}{3}-2}(\beta)
43C12[4K12+278C02C13log1(γ(β))xuC([0,1];𝕎)2]γ7C1x32(β).\displaystyle\leq\frac{4}{3}C_{1}^{2}\left[4K_{1}^{2}+\frac{27}{8}C_{0}^{2}C_{1}^{-3}\log^{-1}\left(\gamma\left(\beta\right)\right)x\left\|u\right\|_{C\left([0,1];\mathbb{W}\right)}^{2}\right]\gamma^{\frac{7C_{1}x}{3}-2}(\beta).

Since C1<6/7C_{1}<6/7, we have γ7C1x1/32(β)0\gamma^{7C_{1}x_{1}/3-2}(\beta)\to 0 as β0\beta\to 0, which guarantees the strong convergence of the scheme. Notice that if in (31) we drop the first two terms on the right-hand side (i.e. the gradient terms), we, after back-substitution, arrive at

uβε(x,)u(x,)2[4ε2+C02γ2(β)ρβ3xuC([0,1];𝕎)2]γ7C1x/3.\displaystyle\left\|u_{\beta}^{\varepsilon}\left(x,\cdot\right)-u\left(x,\cdot\right)\right\|^{2}\leq\left[4\varepsilon^{2}+C_{0}^{2}\gamma^{-2}\left(\beta\right)\rho_{\beta}^{-3}x\left\|u\right\|_{C\left([0,1];\mathbb{W}\right)}^{2}\right]\gamma^{7C_{1}x/3}. (36)

Thus, using a stronger choice (34) (compared to (32)) we get

uβε(x,)u(x,)2[4K22+278C02C13log3(γ(β))xuC([0,1];𝕎)2]γ7C1x32(β).\displaystyle\left\|u_{\beta}^{\varepsilon}\left(x,\cdot\right)-u\left(x,\cdot\right)\right\|^{2}\leq\left[4K_{2}^{2}+\frac{27}{8}C_{0}^{2}C_{1}^{-3}\log^{-3}\left(\gamma\left(\beta\right)\right)x\left\|u\right\|_{C\left([0,1];\mathbb{W}\right)}^{2}\right]\gamma^{\frac{7C_{1}x}{3}-2}(\beta). (37)

Hence, we complete the proof of the theorem. ∎

Remark 4.3.

It is easy to see that if we adapt the choice (32) to the estimate (36), we then obtain a better rate of convergence than (35) in the following sense:

uβε(x,)u(x,)2Cγ7C1x32(β)log2(γ(β)).\displaystyle\left\|u_{\beta}^{\varepsilon}\left(x,\cdot\right)-u\left(x,\cdot\right)\right\|^{2}\leq C\gamma^{\frac{7C_{1}x}{3}-2}(\beta)\log^{-2}(\gamma(\beta)).

As a consequence of (33), we rely on the compact embedding H1(0,1)C[0,1]H^{1}(0,1)\subset C[0,1] to conclude the Hölder convergence in C([0,1]×[0,1])C([0,1]\times[0,1]), viz.

sup(x,y)[0,1]2|uβε(x,y)u(x,y)|2Cγ7C132(β).\sup_{\left(x,y\right)\in\left[0,1\right]^{2}}\left|u_{\beta}^{\varepsilon}\left(x,y\right)-u\left(x,y\right)\right|^{2}\leq C\gamma^{\frac{7C_{1}}{3}-2}(\beta).

As to the choice (32), we can particularly take γ(β)=β1\gamma(\beta)=\beta^{-1} with β=ε1/2\beta=\varepsilon^{1/2}. Meanwhile, choosing γ(β)=β1\gamma(\beta)=\beta^{-1} with β=ε\beta=\varepsilon fulfills the choice (34). In principle, they are of the Hölder rate of convergence. Last but not least, we remark that since 2C1log(γ(β))/3>12C_{1}\log\left(\gamma\left(\beta\right)\right)/3>1 is acquired in the convergence analysis, we suppose a sufficiently small noise level ε<e3/C1\varepsilon<e^{-3/C_{1}} for the particular choices have been made above.

5 Some crucial remarks and generalizations

5.1 A choice of the perturbing and stabilized operators in numerics

The existence of the perturbing and stabilized operators has been pointed out in [16], theoretically. From the simulation standpoint, we need to truncate high Fourier frequencies during the simulation part. Nevertheless, we know that truncation of frequencies should be an appropriate noise-dependent procedure. In this regard, convergence is no longer guaranteed because the numerical perturbing and stabilized operators we use may not satisfy their established definitions. Therefore, such theoretical operators are not practical. In this work, we rely on the so-called Fourier truncation method to propose the following operator:

𝐐εβu=2μj>116log2(γ(β))μju,ϕjϕj.\displaystyle\mathbf{Q}_{\varepsilon}^{\beta}u=2\sum_{\mu_{j}>\frac{1}{16}\log^{2}\left(\gamma\left(\beta\right)\right)}\mu_{j}\left\langle u,\phi_{j}\right\rangle\phi_{j}. (38)

We can show in (38) that 𝐐εβu2uH2(0,1)\left\|\mathbf{Q}_{\varepsilon}^{\beta}u\right\|\leq 2\left\|u\right\|_{H^{2}\left(0,1\right)} for any uH2(0,1)u\in H^{2}(0,1), which is obviously better than (3), but our convergence analysis cannot work without the decay behavior of 𝐐εβ\mathbf{Q}_{\varepsilon}^{\beta} in (3). Aided by the Parseval identity, we estimate 𝐐εβ\mathbf{Q}_{\varepsilon}^{\beta} in (38) as follows:

𝐐εβu2γ2(β)(ΔeΔ)u2,\displaystyle\left\|\mathbf{Q}_{\varepsilon}^{\beta}u\right\|^{2}\leq\gamma^{-2}\left(\beta\right)\left\|\left(-\Delta e^{\sqrt{-\Delta}}\right)u\right\|^{2}, (39)

where we have rewritten (3) as

𝐐εβu\displaystyle\mathbf{Q}_{\varepsilon}^{\beta}u =2μj>116log2(γ(T,β))eμjμjeμju,ϕjϕj.\displaystyle=2\sum_{\mu_{j}>\frac{1}{16}\log^{2}\left(\gamma\left(T,\beta\right)\right)}e^{-\sqrt{\mu_{j}}}\mu_{j}e^{\sqrt{\mu_{j}}}\left\langle u,\phi_{j}\right\rangle\phi_{j}.

In (39), we obtain 𝕎\mathbb{W} as a characterization of a Gevrey space Gσs/2G^{s/2}_{\sigma} with s=2,σ=1s=2,\sigma=1. This Gevrey space was postulated in [20, Definition 2.1] for the first time in the field of inverse and ill-posed problems. Cf. [21], we have the following relation for (1):

μje(1x)μj(u(x,),ϕj+ux(x,),ϕjμj)=μju(1,),ϕj+μjux(1,),ϕj.\displaystyle\mu_{j}e^{\left(1-x\right)\sqrt{\mu_{j}}}\left(\left\langle u\left(x,\cdot\right),\phi_{j}\right\rangle+\frac{\left\langle u_{x}\left(x,\cdot\right),\phi_{j}\right\rangle}{\sqrt{\mu_{j}}}\right)=\mu_{j}\left\langle u\left(1,\cdot\right),\phi_{j}\right\rangle+\sqrt{\mu_{j}}\left\langle u_{x}\left(1,\cdot\right),\phi_{j}\right\rangle. (40)

This means that it suffices to assume the true solution satisfies u(1,)H2(0,1)u(1,\cdot)\in H^{2}(0,1) and ux(1,)H1(0,1)u_{x}(1,\cdot)\in H^{1}(0,1) to fulfill the Gevrey space 𝕎\mathbb{W}. It is because of the fact that

supx[0,1](ΔeΔ)u(x,)2supx[0,1][jμj2e2(1x)μj(u(x,),ϕj+ux(x,),ϕjμj)2].\displaystyle\sup_{x\in\left[0,1\right]}\left\|\left(-\Delta e^{\sqrt{-\Delta}}\right)u\left(x,\cdot\right)\right\|^{2}\leq\sup_{x\in\left[0,1\right]}\left[\sum_{j\in\mathbb{N}}\mu_{j}^{2}e^{2\left(1-x\right)\sqrt{\mu_{j}}}\left(\left\langle u\left(x,\cdot\right),\phi_{j}\right\rangle+\frac{\left\langle u_{x}\left(x,\cdot\right),\phi_{j}\right\rangle}{\sqrt{\mu_{j}}}\right)^{2}\right]. (41)

Proofs of (40) and (41) can be found in A. On the order hand, this assumption is suitable when we consider the standard strong solution in H2((0,1)×(0,1))H^{2}((0,1)\times(0,1)) of the Laplace equation, the forward system of (1), using the standard Sobolev embedding H2(0,1)C1[0,1]H^{2}(0,1)\subset C^{1}[0,1].

Now, using (38) we obtain the following stabilized operator:

𝐏εβu=2μj116log2(γ(β))μju,ϕjϕj=2μj116log2(γ(β))μj1/2μj1/2u,ϕjϕj,\displaystyle\mathbf{P}_{\varepsilon}^{\beta}u=-2\sum_{\mu_{j}\leq\frac{1}{16}\log^{2}\left(\gamma\left(\beta\right)\right)}\mu_{j}\left\langle u,\phi_{j}\right\rangle\phi_{j}=-2\sum_{\mu_{j}\leq\frac{1}{16}\log^{2}\left(\gamma\left(\beta\right)\right)}\mu_{j}^{1/2}\mu_{j}^{1/2}\left\langle u,\phi_{j}\right\rangle\phi_{j}, (42)

which essentially leads to (4) with C1=1/2C_{1}=1/2. Observe that we can control C1C_{1} based upon the Fourier frequencies we want to cut off.

5.2 A linearized version of the QR scheme

Even though the QR scheme designed in (5)–(6) is a linear mapping with respect to the regularized solution uβεu_{\beta}^{\varepsilon}, it is still hard to apply it to the simulation regime. It is because of our concrete choice of operators formulated by a truncated Fourier series in subsection 5.1. It is then reasonable to propose a linearization for this QR scheme. Cf. [22], we investigate convergence of the following numerical scheme for (5)–(6):

k1:\displaystyle k\geq 1: 2x2uβε,k2y2uβε,k=𝐏εβuβε,k1,\displaystyle\quad\frac{\partial^{2}}{\partial x^{2}}u_{\beta}^{\varepsilon,k}-\frac{\partial^{2}}{\partial y^{2}}u_{\beta}^{\varepsilon,k}=-\mathbf{P}_{\varepsilon}^{\beta}u_{\beta}^{\varepsilon,k-1},
k=0:\displaystyle k=0: uβε,0=0.\displaystyle\quad u_{\beta}^{\varepsilon,0}=0. (43)

Here, for k1k\geq 1 we associate the PDE with the Dirichlet boundary condition and the initial conditions:

{uβε,k(x,0)=uβε,k(x,1)=0for x[0,1],uβε,k(0,y)=u0ε(y),xuβε,k(0,y)=0for y[0,1].\begin{cases}u_{\beta}^{\varepsilon,k}\left(x,0\right)=u_{\beta}^{\varepsilon,k}\left(x,1\right)=0&\text{for }x\in[0,1],\\ u_{\beta}^{\varepsilon,k}\left(0,y\right)=u_{0}^{\varepsilon}\left(y\right),\partial_{x}u_{\beta}^{\varepsilon,k}\left(0,y\right)=0&\text{for }y\in[0,1].\end{cases}

The weak solvability of this linearization scheme can be proceeded as in section 3. Therefore, we skip it in this part. However, we note that the right-hand side of the PDE for k=1k=1 vanishes since 𝐏εβuβε,0=0\mathbf{P}_{\varepsilon}^{\beta}u_{\beta}^{\varepsilon,0}=0. Thereby, the weak solvability for uβε,1u_{\beta}^{\varepsilon,1} is attained for any ε>0\varepsilon>0. In other words, for any ε\varepsilon one has

uβε,1C([0,1];H1(0,1))+xuβε,1C([0,1];L2(0,1))C.\displaystyle\left\|u_{\beta}^{\varepsilon,1}\right\|_{C\left(\left[0,1\right];H^{1}\left(0,1\right)\right)}+\left\|\partial_{x}u_{\beta}^{\varepsilon,1}\right\|_{C\left(\left[0,1\right];L^{2}\left(0,1\right)\right)}\leq C. (44)

Our convergence analysis below shows that the refinement in xx should be dependent of the noise level ε\varepsilon. This means that the linearization is a local-in-xx approximation of the regularized solution uβεu^{\varepsilon}_{\beta}. Nevertheless, since the convergence of uβεu^{\varepsilon}_{\beta} to the true solution uu is global in xx (cf. section 4), the local approximation under consideration does not affect the whole convergence of the QR scheme. It is because of the fact that for every ε>0\varepsilon>0 we can divide the domain of xx into many finite sub-domains and repeat the linearization procedure in every sub-domain. Hence, it suffices to consider the linearization scheme (43) in a sub-domain [0,x¯][0,1][0,\bar{x}]\subset[0,1] and below, we single out the choice of x¯:=x¯β>0\bar{x}:=\bar{x}_{\beta}>0 based upon γ(β)\gamma(\beta) to ensure the strong convergence of our linearization in x¯\mathbb{H}_{\bar{x}}, where

x¯:={uC([0,x¯];H1(0,1)):uxC([0,x¯];L2(0,1))}.\mathbb{H}_{\bar{x}}:=\left\{u\in C\left(\left[0,\bar{x}\right];H^{1}\left(0,1\right)\right):u_{x}\in C\left(\left[0,\bar{x}\right];L^{2}\left(0,1\right)\right)\right\}.

.

Theorem 5.1.

The approximate solution uβε,ku_{\beta}^{\varepsilon,k} defined in (43) is strongly convergent in x¯\mathbb{H}_{\bar{x}}. Furthermore, for each ε>0\varepsilon>0 we can find a sufficiently small ηβ(0,1)\eta_{\beta}\in(0,1) such that

uβε,kuβεx¯ηβk1ηβuβε,1x¯.\left\|u_{\beta}^{\varepsilon,k}-u_{\beta}^{\varepsilon}\right\|_{\mathbb{H}_{\bar{x}}}\leq\frac{\eta_{\beta}^{k}}{1-\eta_{\beta}}\left\|u_{\beta}^{\varepsilon,1}\right\|_{\mathbb{H}_{\bar{x}}}.
Proof.

Similar to our proofs above, we herein rely on energy estimates for the difference vk+1=uβε,k+1uβε,kv^{k+1}=u_{\beta}^{\varepsilon,k+1}-u_{\beta}^{\varepsilon,k}. Different from the difference equation (25), this time we do not have the presence of the perturbing operator 𝐐εβ\mathbf{Q}_{\varepsilon}^{\beta}. The difference equation for vk+1v^{k+1} reads as

2x2vk+12y2vk+1=𝐏εβvk,\frac{\partial^{2}}{\partial x^{2}}v^{k+1}-\frac{\partial^{2}}{\partial y^{2}}v^{k+1}=-\mathbf{P}_{\varepsilon}^{\beta}v^{k},

along with the zero Dirichlet boundary condition and zero initial data. Multiplying the difference equation by vxk+1v_{x}^{k+1} and integrating the resulting equation from 0 to 11, we have

ddx(vxk+1(x,)2+vyk+1(x,)2)\displaystyle\frac{d}{dx}\left(\left\|v_{x}^{k+1}\left(x,\cdot\right)\right\|^{2}+\left\|v_{y}^{k+1}\left(x,\cdot\right)\right\|^{2}\right)
=2𝐏εβvk,vk+12C1log(γ(β))(vk2+vyk2)1/2vxk+1.\displaystyle=-2\left\langle\mathbf{P}_{\varepsilon}^{\beta}v^{k},v^{k+1}\right\rangle\leq 2C_{1}\log\left(\gamma\left(\beta\right)\right)\left(\left\|v^{k}\right\|^{2}+\left\|v_{y}^{k}\right\|^{2}\right)^{1/2}\left\|v_{x}^{k+1}\right\|. (45)

Integrating (45) from 0 to xx, we estimate that

vxk+1(x,)2+vyk+1(x,)2\displaystyle\left\|v_{x}^{k+1}\left(x,\cdot\right)\right\|^{2}+\left\|v_{y}^{k+1}\left(x,\cdot\right)\right\|^{2}
C1log(γ(β))0x(vk(s,)2+vyk(s,)2+vxk+1(s,)2)𝑑s\displaystyle\leq C_{1}\log\left(\gamma\left(\beta\right)\right)\int_{0}^{x}\left(\left\|v^{k}\left(s,\cdot\right)\right\|^{2}+\left\|v_{y}^{k}\left(s,\cdot\right)\right\|^{2}+\left\|v_{x}^{k+1}\left(s,\cdot\right)\right\|^{2}\right)ds
C1log(γ(β))x¯vkC([0,x¯];H1(0,1))2+C1log(γ(β))0xvxk+1(s,)2𝑑s.\displaystyle\leq C_{1}\log\left(\gamma\left(\beta\right)\right)\bar{x}\left\|v^{k}\right\|_{C\left(\left[0,\bar{x}\right];H^{1}\left(0,1\right)\right)}^{2}+C_{1}\log\left(\gamma\left(\beta\right)\right)\int_{0}^{x}\left\|v_{x}^{k+1}\left(s,\cdot\right)\right\|^{2}ds.

Therefore, using the Gronwall inequality we obtain

supx[0,x¯](vxk+1(x,)2+vyk+1(x,)2)\displaystyle\sup_{x\in\left[0,\bar{x}\right]}\left(\left\|v_{x}^{k+1}\left(x,\cdot\right)\right\|^{2}+\left\|v_{y}^{k+1}\left(x,\cdot\right)\right\|^{2}\right)
C1log(γ(β))x¯(vkC([0,x¯];H1(0,1))2+vxkC([0,x¯];L2(0,1))2)γC1x¯(β).\displaystyle\leq C_{1}\log\left(\gamma\left(\beta\right)\right)\bar{x}\left(\left\|v^{k}\right\|_{C\left(\left[0,\bar{x}\right];H^{1}\left(0,1\right)\right)}^{2}+\left\|v_{x}^{k}\right\|_{C\left(\left[0,\bar{x}\right];L^{2}\left(0,1\right)\right)}^{2}\right)\gamma^{C_{1}\bar{x}}\left(\beta\right).

Choosing now x¯\bar{x} small enough such that

ηβ2:=2C1log(γ(β))x¯γC1x¯(β)<1,\displaystyle\eta_{\beta}^{2}:=2C_{1}\log\left(\gamma\left(\beta\right)\right)\bar{x}\gamma^{C_{1}\bar{x}}\left(\beta\right)<1, (46)

we then find that

vk+1C([0,x¯];H1(0,1))2+vxk+1C([0,x¯];L2(0,1))2ηβ2(vkC([0,x¯];H1(0,1))2+vxkC([0,x¯];L2(0,1))2).\left\|v^{k+1}\right\|_{C\left(\left[0,\bar{x}\right];H^{1}\left(0,1\right)\right)}^{2}+\left\|v_{x}^{k+1}\right\|_{C\left(\left[0,\bar{x}\right];L^{2}\left(0,1\right)\right)}^{2}\leq\eta_{\beta}^{2}\left(\left\|v^{k}\right\|_{C\left(\left[0,\bar{x}\right];H^{1}\left(0,1\right)\right)}^{2}+\left\|v_{x}^{k}\right\|_{C\left(\left[0,\bar{x}\right];L^{2}\left(0,1\right)\right)}^{2}\right).

Henceforth, for r1r\geq 1 it holds that

uβε,k+ruβε,kC([0,x¯];H1(0,1))+xuβε,k+rxuβε,kC([0,x¯];L2(0,1))\displaystyle\left\|u_{\beta}^{\varepsilon,k+r}-u_{\beta}^{\varepsilon,k}\right\|_{C\left(\left[0,\bar{x}\right];H^{1}\left(0,1\right)\right)}+\left\|\partial_{x}u_{\beta}^{\varepsilon,k+r}-\partial_{x}u_{\beta}^{\varepsilon,k}\right\|_{C\left(\left[0,\bar{x}\right];L^{2}\left(0,1\right)\right)}
j=1r(uβε,k+juβε,k+j1C([0,x¯];H1(0,1))+xuβε,k+jxuβε,k+j1C([0,x¯];L2(0,1)))\displaystyle\leq\sum_{j=1}^{r}\left(\left\|u_{\beta}^{\varepsilon,k+j}-u_{\beta}^{\varepsilon,k+j-1}\right\|_{C\left(\left[0,\bar{x}\right];H^{1}\left(0,1\right)\right)}+\left\|\partial_{x}u_{\beta}^{\varepsilon,k+j}-\partial_{x}u_{\beta}^{\varepsilon,k+j-1}\right\|_{C\left(\left[0,\bar{x}\right];L^{2}\left(0,1\right)\right)}\right)
j=1rηβk+j1(uβε,1uβε,0C([0,x¯];H1(0,1))+xuβε,1xuβε,0C([0,x¯];L2(0,1)))\displaystyle\leq\sum_{j=1}^{r}\eta_{\beta}^{k+j-1}\left(\left\|u_{\beta}^{\varepsilon,1}-u_{\beta}^{\varepsilon,0}\right\|_{C\left(\left[0,\bar{x}\right];H^{1}\left(0,1\right)\right)}+\left\|\partial_{x}u_{\beta}^{\varepsilon,1}-\partial_{x}u_{\beta}^{\varepsilon,0}\right\|_{C\left(\left[0,\bar{x}\right];L^{2}\left(0,1\right)\right)}\right)
ηβk(1ηβr)1ηβ(uβε,1C([0,x¯];H1(0,1))+xuβε,1C([0,x¯];L2(0,1))).\displaystyle\leq\frac{\eta_{\beta}^{k}\left(1-\eta_{\beta}^{r}\right)}{1-\eta_{\beta}}\left(\left\|u_{\beta}^{\varepsilon,1}\right\|_{C\left(\left[0,\bar{x}\right];H^{1}\left(0,1\right)\right)}+\left\|\partial_{x}u_{\beta}^{\varepsilon,1}\right\|_{C\left(\left[0,\bar{x}\right];L^{2}\left(0,1\right)\right)}\right). (47)

This shows that {uβε,k}k\left\{u_{\beta}^{\varepsilon,k}\right\}_{k\in\mathbb{N}} is a Cauchy sequence in the space x¯\mathbb{H}_{\bar{x}}. Thus, there exists uniquely uβεx¯u_{\beta}^{\varepsilon}\in\mathbb{H}_{\bar{x}} such that uβε,kuβεu_{\beta}^{\varepsilon,k}\to u_{\beta}^{\varepsilon} strongly in x¯\mathbb{H}_{\bar{x}} as kk\to\infty. When rr\to\infty in (47), we have

uβε,kuβεx¯ηβk1ηβuβε,1x¯.\left\|u_{\beta}^{\varepsilon,k}-u_{\beta}^{\varepsilon}\right\|_{\mathbb{H}_{\bar{x}}}\leq\frac{\eta_{\beta}^{k}}{1-\eta_{\beta}}\left\|u_{\beta}^{\varepsilon,1}\right\|_{\mathbb{H}_{\bar{x}}}.

Combining this convergence with the linearity of the stabilized operator, we arrive at

𝐏εβuβε,k𝐏εβuβεC1log(γ(β))uβε,kuβεH1(0,1)ηβkC1log(γ(β))1ηβuβε,1x¯.\left\|\mathbf{P}_{\varepsilon}^{\beta}u_{\beta}^{\varepsilon,k}-\mathbf{P}_{\varepsilon}^{\beta}u_{\beta}^{\varepsilon}\right\|\leq C_{1}\log\left(\gamma\left(\beta\right)\right)\left\|u_{\beta}^{\varepsilon,k}-u_{\beta}^{\varepsilon}\right\|_{H^{1}\left(0,1\right)}\leq\frac{\eta_{\beta}^{k}C_{1}\log\left(\gamma\left(\beta\right)\right)}{1-\eta_{\beta}}\left\|u_{\beta}^{\varepsilon,1}\right\|_{\mathbb{H}_{\bar{x}}}.

Thanks to the choice (46) and to the fact that (44) holds, 𝐏εβuβε,k𝐏εβuβε\mathbf{P}_{\varepsilon}^{\beta}u_{\beta}^{\varepsilon,k}\to\mathbf{P}_{\varepsilon}^{\beta}u_{\beta}^{\varepsilon} strongly in L2(0,1)L^{2}(0,1) as kk\to\infty. Hence, the limit function uβεx¯u_{\beta}^{\varepsilon}\in\mathbb{H}_{\bar{x}} obtained above is actually the solution of the regularized system (5)–(6). We complete the proof of the theorem. ∎

It is worth mentioning that we can choose x¯=12γ2C1(β)\bar{x}=\frac{1}{2}\gamma^{-2C_{1}}(\beta) to fulfill the choice (46) because

2C1log(γ(β))γC1x¯x¯=log(γC1(β))γC1(β)γC1(x¯1)<1.2C_{1}\log\left(\gamma\left(\beta\right)\right)\gamma^{C_{1}\bar{x}}\bar{x}=\log\left(\gamma^{C_{1}}\left(\beta\right)\right)\gamma^{-C_{1}}\left(\beta\right)\gamma^{C_{1}\left(\bar{x}-1\right)}<1.

Now, we state the strong convergence of uβε,ku_{\beta}^{\varepsilon,k} towards the true solution uu by a combination of Theorems 4.25.1.

Theorem 5.2.

Under the assumptions of Theorem 4.2, if the choice of γ(β)\gamma(\beta) in (32) holds, then

xuβε,k(x,)xu(x,)2+uβε,k(x,)u(x,)H1(0,1)2C[γ7C132(β)+ηβ2k].\left\|\partial_{x}u_{\beta}^{\varepsilon,k}\left(x,\cdot\right)-\partial_{x}u\left(x,\cdot\right)\right\|^{2}+\left\|u_{\beta}^{\varepsilon,k}\left(x,\cdot\right)-u\left(x,\cdot\right)\right\|_{H^{1}\left(0,1\right)}^{2}\leq C\left[\gamma^{\frac{7C_{1}}{3}-2}\left(\beta\right)+\eta_{\beta}^{2k}\right].

Also, if the choice (34) holds, then

uβε,k(x,)u(x,)2C[γ7C132(β)+ηβ2k].\left\|u_{\beta}^{\varepsilon,k}\left(x,\cdot\right)-u\left(x,\cdot\right)\right\|^{2}\leq C\left[\gamma^{\frac{7C_{1}}{3}-2}\left(\beta\right)+\eta_{\beta}^{2k}\right].

5.3 Generalizations

5.3.1 Non-homogeneous cases

Even though the inverse problem (1) has a simple form, in this part we show that solving more general equations is simply the same. Cf. [3], the Cauchy problem for a non-homogeneous elliptic equation can be reduced to the simplified one (1). Indeed, consider the following boundary value determination inverse problem:

{wxx+wyy=f(x,y)in (0,1)×(0,1),w(x,0)=g0(x),w(x,1)=g1(x)for x[0,1],w(0,y)=w0(y),wx(0,y)=w1(y)for y[0,1].\displaystyle\begin{cases}w_{xx}+w_{yy}=f\left(x,y\right)&\text{in }\left(0,1\right)\times\left(0,1\right),\\ w\left(x,0\right)=g_{0}(x),w\left(x,1\right)=g_{1}(x)&\text{for }x\in[0,1],\\ w\left(0,y\right)=w_{0}\left(y\right),w_{x}\left(0,y\right)=w_{1}\left(y\right)&\text{for }y\in[0,1].\end{cases} (48)

In this regard, we can look for w=u+u~w=u+\tilde{u}, where uu satisfies the inverse problem (1) and u~\tilde{u} obeys the following boundary value problem:

{u~xx+u~yy=f(x,y)in (0,1)×(0,1),u~(x,0)=g0(x),u~(x,1)=g1(x)for x[0,1],u~x(0,y)=w1(y)for y[0,1],u~(1,y)=(1y)g0(1)+yg1(1)for y[0,1].\displaystyle\begin{cases}\tilde{u}_{xx}+\tilde{u}_{yy}=f\left(x,y\right)&\text{in }\left(0,1\right)\times\left(0,1\right),\\ \tilde{u}\left(x,0\right)=g_{0}\left(x\right),\tilde{u}\left(x,1\right)=g_{1}\left(x\right)&\text{for }x\in\left[0,1\right],\\ \tilde{u}_{x}\left(0,y\right)=w_{1}\left(y\right)&\text{for }y\in\left[0,1\right],\\ \tilde{u}\left(1,y\right)=\left(1-y\right)g_{0}\left(1\right)+yg_{1}\left(1\right)&\text{for }y\in\left[0,1\right].\end{cases} (49)

It is clear that (49) is a two-dimensional elliptic equation with mixed non-homogeneous boundary conditions. Therefore, it is obviously well-posed with respect to all inputs. Henceforth, instead of solving the inverse problem (48), we can simply investigate the simpler case (1), which facilitates a lot of computational issues in the inversion method. Multidimensional non-homogeneous problems can also be transformed into the Cauchy problem for the Laplace equation using the same transformation as above; cf. e.g. [5].

5.3.2 Convergence with large noise (ε1\varepsilon\geq 1)

This is now the first time we attempt to show convergence of the QR scheme with large noise. Here, we are interested in answering the question whether or not our QR scheme is convergent when ε1\varepsilon\gg 1. To do so, we only need to modify our establishment for the perturbing and stabilized operators along with a new auxiliary function. Instead of using γ(β)1\gamma(\beta)\geq 1 as we have done with the case ε0\varepsilon\to 0, we now consider an auxiliary function τ:(0,1)\tau:(0,1)\to\mathbb{R} such that for β(0,1)\beta\in(0,1) there holds

τ(β)1,limβ0τ(β)=0.\tau(\beta)\leq 1,\quad\lim_{\beta\to 0}\tau(\beta)=0.

Thereby, we plug γ(β)=1τ(β)\gamma(\beta)=\frac{1}{\tau(\beta)} into Definitions 2.12.2 and proceed the mathematical analysis as in the case ε0\varepsilon\to 0. In short, we state the principle convergence analysis in the following theorem, while the convergence regarding the choice of τ\tau in terms of large noise ε\varepsilon follows immediately.

Theorem 5.3.

Suppose the measurement assumption (2) holds with ε1\varepsilon\gg 1. If we choose τ(β)1\tau(\beta)\leq 1 such that 2C1log(1τ(β))/3>12C_{1}\log\left(\frac{1}{\tau\left(\beta\right)}\right)/3>1, then the following error estimate holds:

xuβε(x,)xu(x,)2+uβε(x,)u(x,)H1(0,1)2\displaystyle\left\|\partial_{x}u_{\beta}^{\varepsilon}\left(x,\cdot\right)-\partial_{x}u\left(x,\cdot\right)\right\|^{2}+\left\|u_{\beta}^{\varepsilon}\left(x,\cdot\right)-u\left(x,\cdot\right)\right\|_{H^{1}(0,1)}^{2} (50)
43C12[4ε2+278C02C13τ2(β)log3(1τ(β))xuC([0,1];𝕎)2]log2(1τ(β))τ7C1x/3(β).\displaystyle\leq\frac{4}{3}C_{1}^{2}\left[4\varepsilon^{2}+\frac{27}{8}C_{0}^{2}C_{1}^{-3}\tau^{2}\left(\beta\right)\log^{-3}\left(\frac{1}{\tau(\beta)}\right)x\left\|u\right\|_{C\left([0,1];\mathbb{W}\right)}^{2}\right]\log^{2}\left(\frac{1}{\tau(\beta)}\right)\tau^{-7C_{1}x/3}(\beta).

Observe in (50) that when ε1\varepsilon\gg 1 we use the small quantity τ2(β)\tau^{2}\left(\beta\right) to control ε2\varepsilon^{2} as well as the term log2(1τ(β))\log^{2}\left(\frac{1}{\tau(\beta)}\right). Thereupon, the choices of τ(β)\tau(\beta) can be taken as in (32) and (34), respectively. Finally, for intermediate noise like ε1\varepsilon\to 1, we can still get the convergence by taking a shifted auxiliary function τ:=τ1\tau:=\tau-1, for instance.

6 Numerical results

In this part, we take into account a finite difference solution of our proposed QR scheme in the linearized version we have studied above. Since the discretization for the domain of xx has to be dependent of ε\varepsilon in the sense of (46), its mesh is understood as “fine mesh” in this part. Meanwhile, we use “coarse mesh” for the domain of yy. For each ε>0\varepsilon>0, we consider a uniform grid of mesh-points xm=mΔxx_{m}=m\Delta x, where mM\mathbb{N}\ni m\leq M and Δx\Delta x is the equivalent mesh-width in xx. In the same manner, we take yn=nΔyy_{n}=n\Delta y with nN\mathbb{N}\ni n\leq N. As mentioned in (42), the stabilized operator is chosen as

𝐏εβu=2μj116log2(γ(β))μju,ϕjϕj,\mathbf{P}_{\varepsilon}^{\beta}u=-2\sum_{\mu_{j}\leq\frac{1}{16}\log^{2}\left(\gamma\left(\beta\right)\right)}\mu_{j}\left\langle u,\phi_{j}\right\rangle\phi_{j},

which gives C1=1/2C_{1}=1/2. Since we are interested in the choice (34), we follow the arguments in Remark 4.3 to choose γ(β)=β1\gamma(\beta)=\beta^{-1} and β=ε\beta=\varepsilon. By this way, we condition that Δx=12γ2C1(β)=ε/2\Delta x=\frac{1}{2}\gamma^{-2C_{1}}(\beta)=\varepsilon/2 and for simplicity, we take Δx=Δy\Delta x=\Delta y.

In this linearization regime, we only need k=3k=3 because, cf. Theorem 5.2, the convergence is eventually dominated by the Hölder rate in γ\gamma. For example, take ε=102\varepsilon=10^{-2} we find in (46) that

ηβ2=12log(1ε)εεε/40.0233.\eta_{\beta}^{2}=\frac{1}{2}\log\left(\frac{1}{\varepsilon}\right)\frac{\varepsilon}{\varepsilon^{\varepsilon/4}}\approx 0.0233.

Therefore, we compute that ηβ2k1.2×105\eta_{\beta}^{2k}\approx 1.2\times 10^{-5}, while the Hölder rate gives γ7C132(β)=ε5/60.0215\gamma^{\frac{7C_{1}}{3}-2}(\beta)=\varepsilon^{5/6}\approx 0.0215.

As to the Dirichlet eigen-elements we have mentioned in Remark 2.3, it is trivial to get that

ϕj(y)=2sin(jπy),μj=j2π2for j.\phi_{j}\left(y\right)=\sqrt{2}\sin\left(j\pi y\right),\quad\mu_{j}=j^{2}\pi^{2}\quad\text{for }j\in\mathbb{N}.

Prior to the derivation of the discrete version of (43), we note that we are not concerned with highly oscillatory integrals usually met in infinite series (cf. e.g. [23]) because of the very low Fourier domain after truncation. Indeed, for ε=102\varepsilon=10^{-2} one has

μj116log2(1ε)1.33,\mu_{j}\leq\frac{1}{16}\log^{2}\left(\frac{1}{\varepsilon}\right)\approx 1.33,

which means j0.37j\leq 0.37, i.e. j=0j=0 in this case. It also holds up to ε=105\varepsilon=10^{-5}. Note that when j=0j=0, the stabilized operator vanishes. Below, we only consider ε=101,102\varepsilon=10^{-1},10^{-2} (relatively meaning 10%,1%10\%,1\%) as usually met in reality. In this regard, the number of mesh-points in xx and yy is not large and less expensive for real-world measurements.

For ease of presentation, we neglect the presence of β\beta. Taking the starting point um,nε,0=0u_{m,n}^{\varepsilon,0}=0, we seek a discrete solution um,nε,kuε,k(xm,yn)u_{m,n}^{\varepsilon,k}\approx u^{\varepsilon,k}\left(x_{m},y_{n}\right) satisfying the following equation:

um+1,nε,k2um,nε,k+um1,nε,k(Δx)2=um,n+1ε,k2um,nε,k+um,n1ε,k(Δy)2+2Δyj14πlog(1ε)l=0Nμjum,lε,k1ϕj(yl)ϕj(yn).\displaystyle\frac{u_{m+1,n}^{\varepsilon,k}-2u_{m,n}^{\varepsilon,k}+u_{m-1,n}^{\varepsilon,k}}{\left(\Delta x\right)^{2}}=\frac{u_{m,n+1}^{\varepsilon,k}-2u_{m,n}^{\varepsilon,k}+u_{m,n-1}^{\varepsilon,k}}{\left(\Delta y\right)^{2}}+2\Delta y\sum_{j\leq\frac{1}{4\pi}\log\left(\frac{1}{\varepsilon}\right)}\sum_{l=0}^{N}\mu_{j}u_{m,l}^{\varepsilon,k-1}\phi_{j}\left(y_{l}\right)\phi_{j}\left(y_{n}\right). (51)

Letting r=ΔxΔyr=\frac{\Delta x}{\Delta y}, (51) is equivalent to

um+1,nε,k\displaystyle u_{m+1,n}^{\varepsilon,k} =2um,nε,kum1,nε,k+r2(um,n+1ε,k2um,nε,k+um,n1ε,k)\displaystyle=2u_{m,n}^{\varepsilon,k}-u_{m-1,n}^{\varepsilon,k}+r^{2}\left(u_{m,n+1}^{\varepsilon,k}-2u_{m,n}^{\varepsilon,k}+u_{m,n-1}^{\varepsilon,k}\right)
+2(Δx)2Δyj14πlog(1ε)l=0Nμjum,lε,k1ϕj(yl)ϕj(yn)for 1mM1,1nN1.\displaystyle+2\left(\Delta x\right)^{2}\Delta y\sum_{j\leq\frac{1}{4\pi}\log\left(\frac{1}{\varepsilon}\right)}\sum_{l=0}^{N}\mu_{j}u_{m,l}^{\varepsilon,k-1}\phi_{j}\left(y_{l}\right)\phi_{j}\left(y_{n}\right)\;\text{for }1\leq m\leq M-1,1\leq n\leq N-1.

Henceforth, the fully discrete version of (43) can be written in matrix form as

𝐔m+1ε,k=𝐊𝐔mε,k𝐔m1ε,k+𝐅(𝐔mε,k1)for 1mM1,\displaystyle\mathbf{U}_{m+1}^{\varepsilon,k}=\mathbf{K}\mathbf{U}_{m}^{\varepsilon,k}-\mathbf{U}_{m-1}^{\varepsilon,k}+\mathbf{F}\left(\mathbf{U}_{m}^{\varepsilon,k-1}\right)\quad\text{for }1\leq m\leq M-1, (52)

where we have denoted by 𝐔mε,k=(um,1ε,k,um,2ε,k,,um,N1ε,k)TN1\mathbf{U}_{m}^{\varepsilon,k}=\left(u_{m,1}^{\varepsilon,k},u_{m,2}^{\varepsilon,k},\ldots,u_{m,N-1}^{\varepsilon,k}\right)^{\text{T}}\in\mathbb{R}^{N-1} and 𝐊𝕄(N1)×(N1)\mathbf{K}\in\mathbb{M}^{\left(N-1\right)\times\left(N-1\right)}, 𝐅N1\mathbf{F}\in\mathbb{R}^{N-1} given by

𝐊=[22r2r200r222r2r200r222r2r200r222r2],𝐅(𝐔mε,k1)=[𝐅(𝐔mε,k1)(y1)𝐅(𝐔mε,k1)(y2)𝐅(𝐔mε,k1)(yN1)].\displaystyle\mathbf{K}=\begin{bmatrix}2-2r^{2}&r^{2}&0&\cdots&0\\ r^{2}&2-2r^{2}&r^{2}&\cdots&0\\ 0&\ddots&\ddots&\ddots&\vdots\\ \vdots&&r^{2}&2-2r^{2}&r^{2}\\ 0&\cdots&0&r^{2}&2-2r^{2}\end{bmatrix},\;\mathbf{F}\left(\mathbf{U}_{m}^{\varepsilon,k-1}\right)=\begin{bmatrix}\mathbf{F}\left(\mathbf{U}_{m}^{\varepsilon,k-1}\right)\left(y_{1}\right)\\ \mathbf{F}\left(\mathbf{U}_{m}^{\varepsilon,k-1}\right)\left(y_{2}\right)\\ \vdots\\ \vdots\\ \mathbf{F}\left(\mathbf{U}_{m}^{\varepsilon,k-1}\right)\left(y_{N-1}\right)\end{bmatrix}.

Here, the function 𝐅(𝐔mε,k1)\mathbf{F}\left(\mathbf{U}_{m}^{\varepsilon,k-1}\right) can be computed in the previous step of linearization procedure, whose elements read as

𝐅(𝐔mε,k1)(yn)\displaystyle\mathbf{F}\left(\mathbf{U}_{m}^{\varepsilon,k-1}\right)\left(y_{n}\right) =2(Δx)2Δyj14πlog(1ε)l=0Nμjum,lε,k1ϕj(yl)ϕj(yn)\displaystyle=2\left(\Delta x\right)^{2}\Delta y\sum_{j\leq\frac{1}{4\pi}\log\left(\frac{1}{\varepsilon}\right)}\sum_{l=0}^{N}\mu_{j}u_{m,l}^{\varepsilon,k-1}\phi_{j}\left(y_{l}\right)\phi_{j}\left(y_{n}\right)
=2(Δx)2Δyj14πlog(1ε)μj[ϕj(y1)ϕj(y2)ϕj(yN1)]T[um,1ε,k1um,2ε,k1um,N1ε,k1]ϕj(yn).\displaystyle=2\left(\Delta x\right)^{2}\Delta y\sum_{j\leq\frac{1}{4\pi}\log\left(\frac{1}{\varepsilon}\right)}\mu_{j}\begin{bmatrix}\phi_{j}\left(y_{1}\right)\\ \phi_{j}\left(y_{2}\right)\\ \vdots\\ \vdots\\ \phi_{j}\left(y_{N-1}\right)\end{bmatrix}^{\text{T}}\begin{bmatrix}u_{m,1}^{\varepsilon,k-1}\\ u_{m,2}^{\varepsilon,k-1}\\ \vdots\\ \vdots\\ u_{m,N-1}^{\varepsilon,k-1}\end{bmatrix}\phi_{j}\left(y_{n}\right).

Note that due to the Dirichlet condition in yy, we have um,0ε,k=um,Nε,k=0u_{m,0}^{\varepsilon,k}=u_{m,N}^{\varepsilon,k}=0 for 0mM0\leq m\leq M. Moreover, using the initial conditions we endow (52) with

𝐔0ε,k=𝐔1ε,k=(u0ε(y1),u0ε(y2),,u0ε(yN1))T,\displaystyle\mathbf{U}_{0}^{\varepsilon,k}=\mathbf{U}_{1}^{\varepsilon,k}=\left(u_{0}^{\varepsilon}\left(y_{1}\right),u_{0}^{\varepsilon}\left(y_{2}\right),\ldots,u_{0}^{\varepsilon}\left(y_{N-1}\right)\right)^{\text{T}}, (53)

which is attained by our measured data.

It is well-known that the system (52)–(53) satisfies the von Neumann stability condition when r1r\leq 1 due to the explicit finite difference regime we choose. This choice is based on the fact that Δx\Delta x is already very small by its dependence on ε\varepsilon and that the explicit scheme is easier to use in implementation. As to the choice of numerical integration involved in 𝐏εβ\mathbf{P}_{\varepsilon}^{\beta} we simply rely on the Riemann sum because we intend to take Δy\Delta y slightly small for better images’ resolution. The number of yny_{n} is not a matter here since we can apply some quadrature methods to get good accuracy for, e.g., N10N\leq 10. This is already postulated in our previous work [23] and is not our main interest in this work.

Our numerical illustrations consist of two (2) tests, where we suppose to know the true solutions with different shapes. We remark that these true solutions satisfy the non-homogeneous elliptic equation, which turns out that we solve the general inverse problem (48) instead of just (1). The algorithm that transforms (48) into (1) is already given in subsection 5.3. With the analytical solutions, we can easily compute all inputs involved in (48). On the other hand, we take into account the following type of additive measured data of (1):

u0ε(y)=u(0,y)+rand(y)ε=w0(y)u~(0,y)+rand(y)ε,\displaystyle u_{0}^{\varepsilon}\left(y\right)=u\left(0,y\right)+\text{rand}\left(y\right)\varepsilon=w_{0}\left(y\right)-\tilde{u}\left(0,y\right)+\text{rand}\left(y\right)\varepsilon, (54)
yu0ε(y)=yw0(y)yu~(0,y)+rand(y)ε,\displaystyle\partial_{y}u_{0}^{\varepsilon}\left(y\right)=\partial_{y}w_{0}\left(y\right)-\partial_{y}\tilde{u}\left(0,y\right)+\text{rand}\left(y\right)\varepsilon, (55)

where w0(y)w_{0}(y) is known from the analytical solution, u~(0,y)\tilde{u}\left(0,y\right) can be found by solving (49), and rand is a uniformly distributed random number such that maxy[0,1]rand(y)1\max_{y\in\left[0,1\right]}\text{rand}\left(y\right)\leq 1. By this way, we fulfill the assumption (2).

Lastly, we define the 2\ell^{2} and relative errors as

E2=1(M+1)(N+1)m=0Mn=0N|wβε(xm,yn)wtrue(xm,yn)|2,\displaystyle E_{\ell^{2}}=\sqrt{\frac{1}{(M+1)(N+1)}\sum_{m=0}^{M}\sum_{n=0}^{N}\left|w_{\beta}^{\varepsilon}\left(x_{m},y_{n}\right)-w_{\text{true}}\left(x_{m},y_{n}\right)\right|^{2}},
Erel=m=0Mn=0N|wβε(xm,yn)wtrue(xm,yn)|2m=0Mn=0N|wtrue(xm,yn)|2×100%.\displaystyle E_{\text{rel}}=\frac{\sqrt{\sum_{m=0}^{M}\sum_{n=0}^{N}\left|w_{\beta}^{\varepsilon}\left(x_{m},y_{n}\right)-w_{\text{true}}\left(x_{m},y_{n}\right)\right|^{2}}}{\sqrt{\sum_{m=0}^{M}\sum_{n=0}^{N}\left|w_{\text{true}}\left(x_{m},y_{n}\right)\right|^{2}}}\times 100\%.
Remark 6.1.

In practice, we just need to use the measured data (54) to estimate its gradient (55) because measurements are usually expensive. At the discretization level, one has

yu0ε(yn)\displaystyle\partial_{y}u_{0}^{\varepsilon}\left(y_{n}\right) u0ε(yn+1)u0ε(yn)Δy=u0(yn+1)+rand(yn+1)εu0(yn)rand(yn)εΔy\displaystyle\approx\frac{u_{0}^{\varepsilon}\left(y_{n+1}\right)-u_{0}^{\varepsilon}\left(y_{n}\right)}{\Delta y}=\frac{u_{0}\left(y_{n+1}\right)+\text{rand}\left(y_{n+1}\right)\varepsilon-u_{0}\left(y_{n}\right)-\text{rand}\left(y_{n}\right)\varepsilon}{\Delta y}
yu0(yn)+ε(Δy)1(rand(yn+1)rand(yn)).\displaystyle\approx\partial_{y}u_{0}\left(y_{n}\right)+\varepsilon\left(\Delta y\right)^{-1}\left(\text{rand}\left(y_{n+1}\right)-\text{rand}\left(y_{n}\right)\right).

If we take (Δy)1=C1log(γ(β))/3=log(ε1)/3\left(\Delta y\right)^{-1}=C_{1}\log\left(\gamma(\beta)\right)/3=\log\left(\varepsilon^{-1}\right)/3 , we have

|ε(Δy)1(rand(yn+1)rand(yn))|ερβ,\left|\varepsilon\left(\Delta y\right)^{-1}\left(\text{rand}\left(y_{n+1}\right)-\text{rand}\left(y_{n}\right)\right)\right|\leq\varepsilon\rho_{\beta},

where we have recalled ρβ\rho_{\beta} in the proof of Theorem 4.1. Theoretically, the bound ερβ\varepsilon\rho_{\beta} is acceptable because cf. (30), the whole error bound for initial data in the proof remains unchanged. Note that by this ε\varepsilon dependence, the number of measurement points in yy is small. Therefore, as we have mentioned above, one should apply, e.g., the Gauss–Legendre method to get a fine numerical integration for 𝐏εβ\mathbf{P}_{\varepsilon}^{\beta}. This reveals a facing challenge of inverse problems in real-world applications.

6.1 Test 1: sinusoidal humps

In this test, we reconstruct the heat distribution satisfying the inverse problem (48), where we suppose that the true analytical solution is

wtrue(x,y)=sin(6x)sin(6y).w_{\text{true}}\left(x,y\right)=\sin\left(6x\right)\sin\left(6y\right).

This test models sinusoidal humps, which is one of the classical examples in simulation. To validate the proposed scheme, we depict in Figure 1 the computed solutions with ε=10%\varepsilon=10\% and ε=1%\varepsilon=1\% and also, we compare them with the true solution illustrated therein. We can see that the computed solutions are very close to the true one. Besides, we briefly report that for ε=10%\varepsilon=10\% the 2\ell^{2} error is 0.090.09 and the relative error is 15.2%15.2\%. For ε=1%\varepsilon=1\%, the errors reduce to 0.060.06 and 10.8%10.8\%, respectively.

Refer to caption
(a) Computed (ε=101\varepsilon=10^{-1})
Refer to caption
(b) Computed (ε=101\varepsilon=10^{-1})
Refer to caption
(c) Computed (ε=102\varepsilon=10^{-2})
Refer to caption
(d) Computed (ε=102\varepsilon=10^{-2})
Refer to caption
(e) True
Refer to caption
(f) True
Figure 1: (a)–(b) Reconstructed solution in Test 1 with ε=101,M=N=20\varepsilon=10^{-1},M=N=20. (c)–(d) Reconstructed solution in Test 1 with ε=102,M=N=200\varepsilon=10^{-2},M=N=200. (e)–(f) Illustrations of the true solution in Test 1.

6.2 Test 2: a box-shaped protrusion

In the second test, we suppose that

wtrue(x,y)=10.001+(x0.5)4+(y0.5)4,w_{\text{true}}\left(x,y\right)=\frac{1}{0.001+\left(x-0.5\right)^{4}+\left(y-0.5\right)^{4}},

which resembles a scaled “witch of Agnesi”. Unlike Test 1, the true solution in this test attains a really huge maximal value. Therefore, the 2\ell^{2} error can be relatively large. Figure 2 shows illustrations of the computed and true solutions. We observe that for a very coarse mesh in both xx and yy (M=N=20M=N=20) and large noise (10%) the 3D image of the computed solution is very accurate. We remark that the projection of the true solution onto the plane {z=0}\left\{z=0\right\} is box-shaped. Then in this particular comparison, the accuracy of the approximate solution is also enjoyed when ε=1%\varepsilon=1\%. Furthermore, as to the numerical errors we find that for ε=10%\varepsilon=10\% the 2\ell^{2} and relative errors are 5.555.55 and 1.9%1.9\%, respectively. When ε=1%\varepsilon=1\%, they become smaller with 2.452.45 and merely 0.81%0.81\%.

Refer to caption
(a) Computed (ε=101\varepsilon=10^{-1})
Refer to caption
(b) Computed (ε=101\varepsilon=10^{-1})
Refer to caption
(c) Computed (ε=102\varepsilon=10^{-2})
Refer to caption
(d) Computed (ε=102\varepsilon=10^{-2})
Refer to caption
(e) True
Refer to caption
(f) True
Figure 2: (a)–(b) Reconstructed solution in Test 2 with ε=101,M=N=20\varepsilon=10^{-1},M=N=20. (c)–(d) Reconstructed solution in Test 2 with ε=102,M=N=200\varepsilon=10^{-2},M=N=200. (e)–(f) Illustrations of the true solution in Test 2.

7 Conclusions

We have demonstrated that the modified quasi-reversibility method we have developed so far for regularization of time-reversed parabolic problems is well-adapted to solve the Cauchy problem for elliptic equations. In this regard, we have proposed two conditional estimates for our perturbing and stabilized operators, which are crucial for the strong convergence of the regularization scheme for this problem. The notion behind this method is to turn the inverse problem into the forward-like problem, which can be solved by several numerical methods. In this case, we solve the Cauchy problem for the elliptic equation by a linear wave equation. We have exploited the energy method to prove weak solvability of the regularized problem. Moreover, driven by a Carleman-type function, we have obtained the Hölder rate of convergence towards the “ideal” true solution with distinctive cases of noise levels: small (ε1\varepsilon\ll 1), large (ε1\varepsilon\gg 1) and intermediate. This shows the flexibility of our method.

Some upcoming topics should be investigated in the near future.

  1. 1.

    The present choice of the perturbing and stabilized operators is really dependent of information of the eigen-elements. However, solving the Sturm–Liouville eigenvalue problems posed in complex-geometry materials (cf. e.g. [24, 25]) is rather challenging, compared to standard media like rectangle and circle we usually consider in theory. Therefore, our first forthcoming target is about a numerical solution of the regularized system in more complicated domains.

  2. 2.

    To show that our regularization results have general applicability in applied sciences, we are inclined to adapt this new method to cope with many other types of inverse problems. For instance, it is promising to study this method for coefficient inverse problems [26] and inverse fractional PDEs [27, 28].

  3. 3.

    If we take into account the strong solution of the regularized problem, our QR scheme is convergent in H2H^{2}. As in [29, 30], it is then interesting to deduce an error estimation in H2H^{2}.

Appendix A Proofs of (40) and (41)

To prove (40) and (41), we recall the formulation of the Fourier coefficients of the true solution, which usually indicates the natural ill-posedness of the Cauchy problem for elliptic equations. Cf. [21], we compute for the Laplace system (1) that

u(x,),ϕj=cosh(μjx)u0,ϕj,ux(x,),ϕj=μjsinh(μjx)u0,ϕj,\displaystyle\left\langle u\left(x,\cdot\right),\phi_{j}\right\rangle=\cosh\left(\sqrt{\mu_{j}}x\right)\left\langle u_{0},\phi_{j}\right\rangle,\quad\left\langle u_{x}\left(x,\cdot\right),\phi_{j}\right\rangle=\sqrt{\mu_{j}}\sinh\left(\sqrt{\mu_{j}}x\right)\left\langle u_{0},\phi_{j}\right\rangle,

which yield

u(x,),ϕj+ux(x,),ϕjμj=eμjxu0,ϕj,\displaystyle\left\langle u\left(x,\cdot\right),\phi_{j}\right\rangle+\frac{\left\langle u_{x}\left(x,\cdot\right),\phi_{j}\right\rangle}{\sqrt{\mu_{j}}}=e^{\sqrt{\mu_{j}}x}\left\langle u_{0},\phi_{j}\right\rangle,
u(1,),ϕj+ux(1,),ϕjμj=eμju0,ϕj.\displaystyle\left\langle u\left(1,\cdot\right),\phi_{j}\right\rangle+\frac{\left\langle u_{x}\left(1,\cdot\right),\phi_{j}\right\rangle}{\sqrt{\mu_{j}}}=e^{\sqrt{\mu_{j}}}\left\langle u_{0},\phi_{j}\right\rangle.

Henceforth, we obtain the relation (40). Proof of (41) follows because

1μju(x,),ϕjux(x,),ϕj=cosh(μjx)sinh(μjx)|u0,ϕj|20.\frac{1}{\sqrt{\mu_{j}}}\left\langle u\left(x,\cdot\right),\phi_{j}\right\rangle\left\langle u_{x}\left(x,\cdot\right),\phi_{j}\right\rangle=\cosh\left(\sqrt{\mu_{j}}x\right)\sinh\left(\sqrt{\mu_{j}}x\right)\left|\left\langle u_{0},\phi_{j}\right\rangle\right|^{2}\geq 0.

Acknowledgment

V.A.K was partly supported by the Research Foundation-Flanders (FWO) in Belgium under the project named “Approximations for forward and inverse reaction-diffusion problems related to cancer models”. V.A.K is very grateful for the guidance and support of Prof. Dr. Loc Hoang Nguyen (Charlotte, USA) during the fellowship at UNCC.

References

  • [1] S. I. Kabanikhin, Definitions and examples of inverse and ill-posed problems, Journal of Inverse and Ill-posed Problems 16 (4) (2008) 317–357. doi:10.1515/jiip.2008.019.
  • [2] A. Leitão, An iterative method for solving elliptic cauchy problems, Numerical Functional Analysis and Optimization 21 (5-6) (2000) 715–742. doi:10.1080/01630560008816982.
  • [3] Z. Qian, C.-L. Fu, Z.-P. Li, Two regularization methods for a Cauchy problem for the Laplace equation, Journal of Mathematical Analysis and Applications 338 (1) (2008) 479–489. doi:10.1016/j.jmaa.2007.05.040.
  • [4] N. H. Tuan, D. D. Trong, P. H. Quan, A note on a Cauchy problem for the Laplace equation: Regularization and error estimates, Applied Mathematics and Computation 217 (7) (2010) 2913–2922. doi:10.1016/j.amc.2010.09.019.
  • [5] L. Eldén, V. Simoncini, A numerical solution of a Cauchy problem for an elliptic equation by Krylov subspaces, Inverse Problems 25 (6) (2009) 065002. doi:10.1088/0266-5611/25/6/065002.
  • [6] D. N. Hào, D. Lesnic, The Cauchy problem for Laplace’s equation via the conjugate gradient method, IMA Journal of Applied Mathematics 65 (2) (2000) 199–217. doi:10.1093/imamat/65.2.199.
  • [7] T. Takeuchi, M. Yamamoto, Tikhonov regularization by a reproducing kernel Hilbert space for the Cauchy problem for an elliptic equation, SIAM Journal on Scientific Computing 31 (1) (2008) 112–142. doi:10.1137/070684793.
  • [8] M. V. Klibanov, Carleman estimates for the regularization of ill-posed Cauchy problems, Applied Numerical Mathematics 94 (2015) 46–74. doi:10.1016/j.apnum.2015.02.003.
  • [9] H.-J. Reinhardt, H. Han, D. N. Hào, Stability and regularization of a discrete approximation to the Cauchy problem for Laplace’s equation, SIAM Journal on Numerical Analysis 36 (3) (1999) 890–905. doi:10.1137/s0036142997316955.
  • [10] R. S. Falk, P. B. Monk, Logarithmic convexity for discrete harmonic functions and the approximation of the Cauchy problem for Poisson’s equation, Mathematics of Computation 47 (175) (1986) 135. doi:10.2307/2008085.
  • [11] L. Eldén, F. Berntsson, A stability estimate for a Cauchy problem for an elliptic partial differential equation, Inverse Problems 21 (5) (2005) 1643–1653. doi:10.1088/0266-5611/21/5/008.
  • [12] M. V. Klibanov, Carleman estimates for global uniqueness, stability and numerical methods for coefficient inverse problems, Journal of Inverse and Ill-Posed Problems 21 (4) (2013) 477–560. doi:10.1515/jip-2012-0072.
  • [13] T. Wei, Y. C. Hon, J. Cheng, Computation for MultiDimensional Cauchy problem, SIAM Journal on Control and Optimization 42 (2) (2003) 381–396. doi:10.1137/s0363012901389391.
  • [14] Y. Sun, Indirect boundary integral equation method for the Cauchy problem of the Laplace equation, Journal of Scientific Computing 71 (2) (2016) 469–498. doi:10.1007/s10915-016-0308-4.
  • [15] R. Chapko, B. T. Johansson, On the numerical solution of a Cauchy problem for the Laplace equation via a direct integral equation approach, Inverse Problems and Imaging 6 (1) (2012) 25–38. doi:10.3934/ipi.2012.6.25.
  • [16] H. T. Nguyen, V. A. Khoa, V. A. Vo, Analysis of a quasi-reversibility method for a terminal value quasi-linear parabolic problem with measurements, SIAM Journal on Mathematical Analysis 51 (1) (2019) 60–85. doi:10.1137/18m1174064.
  • [17] R. Lattès, J. L. Lions, Méthode de Quasi-réversibilité et Applications, Paris, Dunod, 1967.
  • [18] A. B. Ferrari, E. S. Titi, Gevrey regularity for nonlinear analytic parabolic equations, Communications in Partial Differential Equations 23 (1–2) (1998) 424–448. doi:10.1080/03605309808821336.
  • [19] L. Beilina, M. V. Klibanov, Approximate Global Convergence and Adaptivity for Coefficient Inverse Problems, Springer US, 2012. doi:10.1007/978-1-4419-7805-9.
  • [20] N. H. Tuan, L. D. Thang, V. A. Khoa, T. Tran, On an inverse boundary value problem of a nonlinear elliptic equation in three dimensions, Journal of Mathematical Analysis and Applications 426 (2) (2015) 1232–1261. doi:10.1016/j.jmaa.2014.12.047.
  • [21] V. A. Khoa, T. T. Hung, Regularity bounds for a Gevrey criterion in a kernel-based regularization of the Cauchy problem of elliptic equations, Applied Mathematics Letters 69 (2017) 75–81. doi:10.1016/j.aml.2017.02.009.
  • [22] N. T. Long, A. P. N. Dinh, T. N. Diem, Linear recursive schemes and asymptotic expansion associated with the Kirchoff–Carrier operator, Journal of Mathematical Analysis and Applications 267 (1) (2002) 116–134. doi:10.1006/jmaa.2001.7755.
  • [23] V. A. Khoa, M. T. N. Truong, N. H. M. Duy, N. H. Tuan, The Cauchy problem of coupled elliptic sine–Gordon equations with noise: Analysis of a general kernel-based regularization and reliable tools of computing, Computers & Mathematics with Applications 73 (1) (2017) 141–162. doi:10.1016/j.camwa.2016.11.001.
  • [24] D. S. Grebenkov, B.-T. Nguyen, Geometrical structure of Laplacian eigenfunctions, SIAM Review 55 (4) (2013) 601–667. doi:10.1137/120880173.
  • [25] X. Liu, S. Oishi, Verified eigenvalue evaluation for the Laplacian over polygonal domains of arbitrary shape, SIAM Journal on Numerical Analysis 51 (3) (2013) 1634–1654. doi:10.1137/120878446.
  • [26] M. V. Klibanov, D.-L. Nguyen, L. H. Nguyen, A coefficient inverse problem with a single measurement of phaseless scattering data, SIAM Journal on Applied Mathematics 79 (1) (2019) 1–27. doi:10.1137/18m1168303.
  • [27] D. D. Trong, D. N. D. Hai, N. D. Minh, Optimal regularization for an unknown source of space-fractional diffusion equation, Applied Mathematics and Computation 349 (2019) 184–206. doi:10.1016/j.amc.2018.12.030.
  • [28] K. Šišková, M. Slodička, Identification of a source in a fractional wave equation from a boundary measurement, Journal of Computational and Applied Mathematics 349 (2019) 172–186. doi:10.1016/j.cam.2018.09.020.
  • [29] M. V. Klibanov, F. Santosa, A computational quasi-reversibility method for Cauchy problems for Laplace’s equation, SIAM Journal on Applied Mathematics 51 (6) (1991) 1653–1675. doi:10.1137/0151085.
  • [30] L. Bourgeois, Convergence rates for the quasi-reversibility method to solve the Cauchy problem for Laplace’s equation, Inverse Problems 22 (2) (2006) 413–430. doi:10.1088/0266-5611/22/2/002.