Brought to you by:
Paper

Direct algorithm for reconstructing small absorbers in thermoacoustic tomography problem from a single data

and

Published 25 May 2020 © 2020 IOP Publishing Ltd
, , Citation Hanin Al Jebawy and Abdellatif El Badia 2020 Inverse Problems 36 065010 DOI 10.1088/1361-6420/ab8862

0266-5611/36/6/065010

Abstract

This work deals with the inverse thermoacoustic tomography (TAT) problem. It is a biomedical, multi-wave imaging technique, based on the photoacoustic effect (generation of sound from light) that was discovered in 1880 by Alexander Graham Bell. The inverse problem we are concerned in is the inverse source problem of recovering small absorbers in a bounded domain ${\Omega}\subset {\mathbb{R}}^{3}$ without following the quantitative thermoacoustic tomography approach. In our work, we follow a direct algebraic algorithm, that was first developed in Badia and Ha-Duong (2001 Inverse Problems 17 1127), which allows us to reconstruct the number of absorbers, their locations, and some information about the conductivity coefficient directly from the wave equation with constant sound speed, and using a single data, without the knowledge of the coupling Grüneisen's coefficient measuring the strength of the photoacoustic effect. Finally, we establish the corresponding Hölder stability result.

Export citation and abstract BibTeX RIS

1. Introduction and statement of the problem

Thermoacoustic tomography (TAT) is a biomedical imaging technique similar to photo-acoustic tomography (PAT), which is based on the photoacoustic effect (production of sound from light) that was discovered in 1880 by Alexander Graham Bell. However, the main difference between the two modalities is in the type of the radiation used. In photoacoustic tomography, a laser (high frequency radiation) is delivered into the biological tissue to be imaged [13], while in thermoacoustic tomography a radio wave (low frequency radiation) is used. The propagation of this electromagnetic radiation is modeled by the wave equation, see [4]

Equation (1.1)

where Ω is an open, bounded and connected domain in ${\mathbb{R}}^{3}$, with a smooth boundary Γ; c is the light speed in tissues; μ is the permeability; σ = σ(x) is the conductivity to be determined; and the incoming data w is the source radiation. Some of the delivered energy is absorbed by the underlying medium and converted into heat, which in turn generates acoustic waves. The absorbed radiation by the underlying medium is given by [5],

Equation (1.2)

Then, using ultrasonic transducers, the acoustic signal is measured at the boundary in order to localize the absorbing subregions. The acoustic wave generated is described by the pressure p that satisfies the boundary value problem

Equation (1.3)

equipped with the observation

Equation (1.4)

where cs is the acoustic speed, chosen to be constant, and β(x) is the Grüneisen's coefficient that measures the efficiency of conversion from radio waves into acoustic waves.

As explained in [6, 7], tumors have a significantly different electrical properties from normal tissues. In fact, relatively few blood vessels can be found in the inner regions of tumors, which is responsible for lack in nutrients 'supply necessary for cells' survival, and accumulation of metabolic wastes, leading to cell death and formation of scar tissues. This results in higher heterogeneity and much less uniform arrangement of cells, which leads to a significant difference in conductivity between healthy tissues and cancerous ones. Moreover, various studies show that for low frequencies, which is the case in TAT, a significant difference in conductivity is detected between healthy tissues and tumors. A study by Smith et al [8] on liver tissue, for example, shows that tumor conductivity is 6–7 times higher than that of healthy liver cells. Another studies such as [6] on liver tissues, and [9] on breast cells show also a remarkable difference in conductivity between normal tissues and cancerous ones. Thus, measuring the conductivity σ would be of great importance and makes TAT an effective method in detecting tumors.

This problem is highly considered in literature, and different methods are used for solving it (see [10, 11] and references therein), in which the main tool used is the quantitative thermoacoustic tomography approach (qTAT) [1, 4, 12, 13], which is divided into two main steps: the first step is solving the inverse source problem for the wave equation (1.3) to reconstruct the absorbed electromagnetic energy H(x, t), and the second step is to reconstruct the conductivity coefficient σ using the relation (1.2) and equation (1.1). The main difficulty of this approach is that the inverse source problem for the reconstruction of the absorbed electromagnetic energy is highly ill-posed for general source term H(x, t), in particular it lacks uniqueness, see for example [14, 15] and references therein. For this purpose, many studies consider the source term H to be separable, that is H(x, t) = H0(x)Φ(x, t), where Φ is known and H0 is the unknown function that shall be determined (see [15, 16]). Also, in the quantitative photo-acoustic tomography problem, for example, the source term is assumed to satisfy H(x, t) = H0(x)δ(t), where δ is the Dirac delta function (see [4, 17, 18] and references therein). Moreover, the coefficient β is often assumed to be constant and known.

However, in our work we do not follow the quantitative thermoacoustic tomography approach, and we propose a new direct identification process based on the algebraic algorithm for the reconstruction of the conductivity coefficient σ directly from the wave equation (1.3). This algorithm was first developed in [14], and then generalized in [19]. Moreover, in our algorithm we do not require the knowledge of the coupling coefficient β, which is assumed throughout this paper to be variable and unknown with β(x) ∈ L(Ω). Furthermore, different studies in literature have proposed numerical algorithms to solve the problem. In [20], for example, the authors have considered the TAT problem in ${\mathbb{R}}^{n}$ with variable sound speed, and proposed a numerical reconstruction method based on the back projection algorithms. Also, TAT was studied through spherical or circular Radon transform, see for example [3, 21].

In this paper, we consider the inverse thermoacoustic problem of recovering the conductivity coefficient σ, defined piece-wisely. That is

where σj are functions belonging to the spaces ${L}^{\infty }\left({\mathcal{D}}_{j}\right)$, and ${\mathcal{D}}_{j}$ are small domains representing the absorbers (tumors) and defined as follows

Equation (1.5)

where Sj ∈ Ω are the locations of the sub-domains ${\mathcal{D}}_{j}$, and ${B}_{j}\subset {\mathbb{R}}^{3}$ are bounded domains containing the origin. The constant epsilon is the common order of magnitude of the diameters of ${\mathcal{D}}_{j}$ supposed to be a sufficiently small positive real number, taken without loss of generality, smaller than 1. Moreover, the domains ${\mathcal{D}}_{j}$ are assumed to be far from the boundary Γ and their centers are supposed to be mutually distinct. We denote by

the background medium. The difference in conductivity between healthy tissues and cancerous ones allows us to assume that the conductivity coefficient σ0 in the background medium ${\mathcal{D}}_{0}$ is very small compared to that in the absorbers ${\left\{{\mathcal{D}}_{j}\right\}}_{j=1,\dots ,m}$. More precisely, σ0 = O(epsilonκ) for some κ > 0. This is an important assumption that we will use in our reconstruction algorithm later. The inverse problem we are interested in, consists in determining the number of absorbers m, the centers Sj, and some characteristics related to ${\mathcal{D}}_{j}$, using only one acoustic boundary measurement ${p}_{{\Gamma}{\times}\left(0,T\right)}$. In addition, we study the stability of reconstruction of the centers Sj.

The rest of this paper is organized as follows. In section 2, we propose our algebraic algorithm for the reconstruction of the number of small domains ${\mathcal{D}}_{j}$, their locations Sj and some information about the coefficients σj. Then, we give in section 3 the corresponding stability result.

2. Identification algorithm

In this section, we present our identification algorithm of the number of sub-domains ${\mathcal{D}}_{j}$ and their centers Sj based on the idea of the so called reciprocity gap functional and by using specific test functions that allow us to get the desired results.

2.1. Reciprocity gap formula

First, we set

where H is given in (1.2), and we consider the adjoint problem

Equation (2.1)

Then, multiplying equation (1.3) by v, and integrating by parts, we obtain using Green's formula

Equation (2.2)

We note that

Let

then

If we denote the so called reciprocity gap functional by

Equation (2.3)

then, from (2.2), we get

Furthermore, the large difference in conductivity between healthy tissues and cancerous ones, see [6, 7, 22, 23], permits us to assume that σ0 = O(epsilonκ) for some κ > 0, assumed in our work to be κ > 4. Consequently, we have

Equation (2.4)

To solve our inverse problem, we need to know the values of p(x, T) and pt(x, T), unless we impose that the solution v of (2.1) satisfies

Equation (2.5)

To overcome this difficulty, and for a source term of the form ${\sum }_{j=1}^{N}{\lambda }_{j}\left(t\right){\delta }_{{\mathbf{S}}_{j}}$, the authors in [14] opted for the determination of the values of the unknowns pt(x, T) and p(x, T) using the fact that the intensities λj(t) become null after a certain time T* < T and by applying the exact controllability approach. In here, this approach is not employed where we opt for the choice of the condition (2.5). Indeed, we introduce what one calls test functions, the solutions of (2.1) verifying the condition (2.5).

2.2. Algebraic relations

In the following, we will show how an appropriate choice of test functions unveils the desired information. The general idea behind our identification method is the projection of the problem onto chosen test functions. This idea is not new, but our approach employs a specific family of functions that allows a practical solution.Let ${\alpha }_{i},{\beta }_{i},{\gamma }_{i}\in \mathbb{R}$, for i = 1, 2 such that

Consider now the one-dimensional wave equation,

Equation (2.6)

Then, employing the test functions

Equation (2.7)

the relation (2.4) becomes

Equation (2.8)

Thus, using the change of variables x = Sj + epsilonτ with τ = (τ1, τ2, τ3), one obtains

Equation (2.9)

where

Equation (2.10)

and consequently,

Equation (2.11)

where

Equation (2.12)

and

Our goal now is to solve the equation (2.11) for the reconstruction of the unknown coefficients ${\nu }_{j}^{k,a}$ and the centers ${P}_{j}^{a}$. We observe that $\forall {n}_{0}\in \mathbb{N}$, the system (2.11), with n = 0, ..., n0, has n0 + 1 equations and m(n0 + 2) unknowns. Then, the number of unknowns is variable with n0, and it increases with the increase of the number of equations. Thus, it is impossible to uniquely reconstruct the unknowns no matter how many equations are taken. To overcome this difficulty, we will truncate these equations from a small error. First, for a given positive epsilon < 1, we choose a fixed positive integer K such that K < κ − 4 and epsilonK+4 is small enough, and we set

Equation (2.13)

Then, according to (2.11), we can see that for nK one has

and for n > K,

where

From (2.13), we notice that the number of unknowns is now fixed at m(K + 1). Thus, we can increase the number of equations, by taking n large enough, to be able to uniquely determine the unknowns. This is the objective of the next subsection.

Remark 2.1. Assuming that the observation time T is large enough, $T{ >}\frac{\text{diam}\,{\Omega}}{{c}_{\text{s}}}$, then a special solution of (2.6) would be

where ${\rho }_{\varepsilon }\in {C}_{0}^{\infty }\left(\mathbb{R}\right)$ is a mollifier function supported in the interval [−ɛ, ɛ] and satisfying ${\int }_{\mathbb{R}}{\rho }_{\varepsilon }\left(s\right)\mathrm{d}s=1$, and θ ∈ [θmin + ɛ, θmaxɛ] with

It is obvious that for every θ and ɛ fixed, the function ρ belongs to the functional space C((γ1, γ2) × (0, T)). Moreover, for each z fixed, ρ(z, t) is supported in the time interval $\left[\theta -\frac{z}{{c}_{\text{s}}}-\varepsilon ,\theta -\frac{z}{{c}_{\text{s}}}+\varepsilon \right]$. Therefore, taking θ ∈ [θmin + ɛ, θmaxɛ] we get ρ(., T) = ρt(., T) = 0.

The mollifier function ρ was used by the authors in [24] for solving the inverse problem of reconstruction of time-varying point sources for wave equation.

Before solving the equation (2.11), we need to know if the projections ${P}_{j}^{a}$ are mutually distinct, which is necessary in order to use the algebraic method proposed below. Indeed, one can remark that there is only a finite number of planes containing the origin such that at least two points among Sj, are projected onto the same point on this plane. So, if a basis is chosen randomly, one is almost sure that the points Sj are projected onto distinct points on every coordinate plane. Therefore, without loss of generality, we assume that:

(H) The projections onto the xy, yz and xz-planes of the points Sj are mutually distinct.

2.3. Reconstruction algorithm

Before presenting our identification algorithm we introduce, for simplicity, some notations and definitions that are used throughout this subsection.

First, denote by ${P}_{j}^{b}$ and ${P}_{j}^{c}$, the projections of Sj onto the yz and xz-complex planes respectively. Then, using in (2.4), the following test functions

Equation (2.14)

where ρ1, and ρ2 are the solutions of (2.6) with the space variable z being replaced by x and y respectively, leads to the following algebraic quantities

Equation (2.15)

and

Equation (2.16)

where

and

Finally, bringing together the equations (2.13), (2.15) and (2.16), we can write

Equation (2.17)

On the other hand, we assume that we know an upper bound M of the number of sources m, then $\overline{J}=M\left(K+1\right)$ is an upper bound of the number J = m(K + 1). We define for r = a, b, c, and k = 0, ..., K, the complex vectors

of sizes $\overline{J},m$ and m(K + 1) respectively, and consider, for all $n\in \mathbb{N}$, the complex matrices ${A}_{n}^{r}$ of size $\overline{J}{\times}J$,

Equation (2.18)

where, for k ∈ {0, ..., K}, ${V}_{n}^{k,r}$ are the confluent $\overline{J}{\times}m$ Vandermonde matrices

Now, let us introduce the Hankel matrix

Equation (2.19)

and the following multi-diagonal matrix

Equation (2.20)

where

We propose an identification processes in four steps. First, we determine the absorbers' number m using the Hankel matrix defined in (2.19), then we obtain the projections ${P}_{j}^{r}$ using a companion matrix. Next, we get the coefficients ${\nu }_{j}^{k,r}$ as a solution of a linear system via the Hankel matrix ${H}_{J}^{r}$, and finally we end up with the locations Sj by solving a similar linear system and using the coefficients ${\nu }_{j}^{k,r}$.

Step 1: determination of the number of absorbers

The first step consists in determining the number of sources by means of the following theorem.

Theorem 2.2. Let ${H}_{\overline{J}}^{r}$ be the Hankel matrix defined in (2.19) where $\overline{J}$ is a known upper bound of J = m(K + 1). Assume that, for r = a, b, c, the projected points ${P}_{j}^{r}$ of Sj are mutually distinct, then we have

The proof of this theorem is similar to that of theorem 2.8 in [19]. It is based on the decomposition of the Hankel matrix ${H}_{\overline{J}}^{r}$ given in lemma 2.9 in [19]. For the convenience of the reader, we recall this lemma here for which we present a sketch of the proof.

Lemma 2.3. Let ${H}_{\overline{J}}^{r}$ be the Hankel matrix defined in (2.19), Lr the multi-diagonal matrix defined in (2.20), ${A}_{0}^{r}$ the Vandermonde matrix defined in (2.18), and ${\left({A}_{0}^{r}\right)}^{t}$ its transpose matrix. Then,

Equation (2.21)

Proof. Indeed, from (2.18), we notice that

Moreover,

Therefore, the algebraic relation (2.17) can be rewritten in a matrix form as

Equation (2.22)

Furthermore, if we denote by Tr the block J × J upper triangular complex matrix

Equation (2.23)

with

then, using the Pascal formula $\left(\genfrac{}{}{0.0pt}{}{n}{j-1}\right)+\left(\genfrac{}{}{0.0pt}{}{n}{j}\right)=\left(\genfrac{}{}{0.0pt}{}{n+1}{j}\right)$, it follows that

and

From the definition of the matrix ${A}_{n}^{r}$, we obtain

and therefore, from (2.22), one gets

Equation (2.24)

Now, thanks to (2.24), one can verify by a simple calculation the following relationship

which ends the proof of Lemma 2.3. □

Proof of Theorem 2.2. We can easily check that $\text{rank}\left({A}_{0}^{r}\right)=\text{rank}{\left({A}_{0}^{r}\right)}^{t}=J$ if and only if the projections ${P}_{j}^{r}$ are mutually distinct, for j = 1, ..., m. Moreover, the matrix Lr is nonsingular if and only if ${\nu }_{j}^{K,r}\ne 0\quad \forall j=1,\dots ,\;m$. Therefore, under the hypothesis (H), and assuming that the coefficients ${\nu }_{j}^{K,r}\ne 0$, we deduce that ${A}_{0}^{r}{L}^{r}$ is surjective and therefore we have $\text{rank}\left({A}_{0}^{r}{L}^{r}{\left({A}_{0}^{r}\right)}^{t}\right)=\text{rank}{\left({A}_{0}^{r}\right)}^{t}=J$, which leads to the desired result. □

Step 2: reconstruction of the projections ${P}_{j}^{r}$

The second step consists in determining the projections onto the xy-, xz- and yz-complex planes of the centers Sj.

Henceforth, we replace $\overline{J}$ by J in the quantities defined above. Thus, from (2.24), we can easily derive that

where we have set

Equation (2.25)

Here, the matrix ${A}_{0}^{r}$ is invertible since the projected points ${P}_{j}^{r}$ are assumed distinct. Moreover, since $\text{rank}\left({H}_{J}^{r}\right)=J$, the family ${\left({\xi }_{n}^{r}\right)}_{n=0,\dots ,J-1}$ forms a basis of ${\mathbb{C}}^{J}$, so the J × J complex matrix Br is given explicitly by

Equation (2.26)

where the vector ${C}^{r}={\left({c}_{0}^{r},\dots ,\;{c}_{J-1}^{r}\right)}^{t}$ is obtained by solving the linear system

Equation (2.27)

Indeed, as the family ${\left({\xi }_{n}^{r}\right)}_{n=0,\dots ,J-1}$ forms a basis of ${\mathbb{C}}^{J}$, given ξJ, there exists a vector ${C}^{r}={\left({c}_{0}^{r},\dots ,\;{c}_{J-1}^{r}\right)}^{t}$ such that

which is equivalent to (2.27) since the matrix ${H}_{J}^{r}=\left({\xi }_{0}^{r},\dots ,\;{\xi }_{J-1}^{r}\right)$.

Finally, the projections ${P}_{j}^{r}$ are given by the following theorem.

Theorem 2.4. Let Br, for r = a, b, c, be the companion matrices defined in (2.26). Assume that the projected points ${P}_{j}^{r}$ of Sj are distinct and away from the origin and that ${\nu }_{j}^{K,r}\ne 0\;$ for j = 1, ..., m. Then,

  • (a)  
    Br admits m eigenvalues of multiplicity K + 1;
  • (b)  
    The m eigenvalues of multiplicity K + 1 are the projections ${P}_{j}^{r}$ of the point sources Sj.

Proof. The proof of this theorem follows from (2.23) and (2.25). □

Remark 2.5. We know according to formulas (2.23), (2.25) and hypothesis (H), that the matrix Br admits m distinct eigenvalues having the same multiplicity K + 1, where K is given. Then, numerically, one should obtain m distinct eigenvalues, m being the rank of the Hankel matrix ${H}_{\overline{J}}^{r}$ defined in (2.19). The rank of ${H}_{\overline{J}}^{r}$ can be obtained numerically as the number of its non-zero singular values.

Step 3: determination of the vectors Λr

In this step, we shall determine the quantities ${\nu }_{j}^{k,r}$, which we will use later to reconstruct the locations Sj. In fact, in order to determine ${\nu }_{j}^{k,r}$, it is sufficient to solve the linear systems

Equation (2.28)

Step 4: complete reconstruction of locations Sj

In step 2, we were able to reconstruct modulo epsilonK+4 the projections of the centers Sj on the xy-, yz- and xz-planes. However, this is not enough. In fact, the combination of different projections does not uniquely determine the centers Sj, see remark 2.6. For this purpose, we fix r = a and we reconstruct, modulo epsilon4, the third components zj that correspond to the previously calculated projections ${P}_{j}^{a}$, on which we need that the coefficients ${\nu }_{j}^{0,a}$ calculated in the previous step to satisfy

Equation (2.29)

We propose an algorithm for the reconstruction of the components zj using specific wave functions. For this purpose, for $n\in \mathbb{N}$, we take ${v}_{n}^{a}$ defined in (2.7) and we consider the following wave functions,

Equation (2.30)

Equation (2.31)

which are solutions of (2.1) satisfying the condition (2.5). Therefore, substituting ${g}_{n}^{a}$ in (2.4) we get

where for every j = 1, ..., m and k = 1, ..., K,

Equation (2.32)

Similar to the previous section, we have for nK

Equation (2.33)

and for n > K

Equation (2.34)

Now, if we choose for $n\in \mathbb{N}$

Equation (2.35)

and we define the following quantities

and

where J = m(K + 1), we get using the formula (2.35) the following system of equations

Equation (2.36)

where ${A}_{0}^{a}$ is defined in (2.18), and ${\tilde {\xi }}_{0}^{a}$ can be approximated using the equations (2.34) and (2.35). Under the assumption that the projections ${P}_{j}^{a}$ are mutually distinct, then ${A}_{0}^{a}$ is invertible, and system (2.36) admits a unique solution. By solving this system, we can reconstruct the quantities ${\tilde {{\nu }_{j}}}^{k,a}$.On the other hand, if we substitute ${h}_{n}^{a}$ in (2.4) we get

Applying the change of variables Dj = Sj + epsilonBj one gets,

where γj(τ, t) is defined in (2.10). Therefore, one has

Hence,

Now, if we define for $n\in \mathbb{N}$

Equation (2.37)

and we take

and

we get, modulo O(epsilon4), the following system

Equation (2.38)

where Pm is the m × m Vandermonde matrix defined by

Since the projections ${P}_{j}^{a}$ are assumed to be mutually distinct, thus Pm is invertible and equation (2.38) admits a unique solution. By solving this equation we can reconstruct ${\nu }_{j}^{0,a}{z}_{j}$ up to O(epsilon4), and since ${\nu }_{j}^{0,a}$ are previously calculated and chosen to be different from zero, thus we can calculate the values of zj. Finally, combining this result with that of section 2.3, we get the locations of the centers Sj.

Remark 2.6. 

  • (a)  
    The projections on the xy-, yz-, and xz-planes do not uniquely determine the centers Sj. For example, the two sets of points
    and
    have the same projections
    and
    on the xy-, yz- and xz-planes respectively. Thus, knowing the projections Pa, Pb and Pc is not enough to uniquely determine the corresponding locations Sj.
  • (b)  
    The last step allows us to uniquely determine the centers Sj. However, we lose some accuracy as the obtained values of zj are given modulo epsilon4.

Summary: Identification algorithm

To sum up, our identification algorithm can be summarized as follows. First, we choose a fixed integer K > 0 such that epsilonK+4 is small enough and we suppose that we know an upper bound $\overline{J}$ of the number J = m(K + 1). Then, we proceed in the following steps:

  • Step 1. We fix r = a, b or c, and we estimate for $n=0,\dots ,\;2\overline{J}-2$ the coefficients ${\alpha }_{n}^{r}$ by $R\left(f,{v}_{n}^{r}\right)$ defined in (2.3), which introduces an accuracy error O(epsilonK+4) in our identification algorithm. Then we construct the Hankel matrix ${H}_{\overline{J}}^{r}$ defined in (2.19).
  • Step 2. We calculate J as the rank of the Hankel matrix ${H}_{\overline{J}}^{r}$, which can be numerically estimated using, for example, the singular value decomposition (SVD) method, and we deduce the number of absorbers m using the relation J = m(K + 1).
  • Step 3. We construct ${C}^{r}={\left({c}_{0}^{r},\dots ,\;{c}_{J-1}^{r}\right)}^{t}$ by solving the system ${H}_{J}^{r}{C}^{r}={\xi }_{J}^{r}$, where ${\xi }_{J}^{r}={\left({\alpha }_{J}^{r},\dots ,\;{\alpha }_{2J-1}^{r}\right)}^{t}$. The projections ${P}_{j}^{r}$ of the centers Sj then can be determined as the eigenvalues of the matrix Br defined in (2.26), which can be constructed using the previously calculated coefficients ${\left\{{c}_{j}^{r}\right\}}_{0{\leqslant}j{\leqslant}J-1}$.
  • Step 4. The coefficients ${\nu }_{j}^{k,a}$ can be obtained by solving the system (2.28).
  • Step 5. We fix r = a and we estimate for n = 0, 1, ..., J − 1 the coefficients ${\tilde {\alpha }}_{n}^{a}$ by $R\left(f,{g}_{n}^{a}\right)$ defined in (2.3), where ${g}_{n}^{a}$ is given by (2.30), then we calculate the coefficients ${\tilde {\nu }}_{j}^{k,a}$ defined in (2.32) by solving system (2.36).
  • Step 6. We calculate $R\left(f,{h}_{n}^{a}\right)$ for n = 1, ..., m, where ${h}_{n}^{a}$ is defined by (2.31). Then using these values and the previously calculated coefficients ${\tilde {\nu }}_{j}^{k,a}$, we may calculate the coefficients ${\tilde {R}}_{n}$ defined in (2.37).
  • Step 7. Finally, by solving (2.38) we get the values of ${\nu }_{j}^{0,a}{z}_{j}$ for j = 1, ..., m. Then under the assumption that the coefficients ${\nu }_{j}^{0,a}$ calculated in step 4 are different from zero, we may calculate the values of zj that correspond to the projections ${P}_{j}^{a}$, and thus we deduce the complete location of the centers Sj.

Remark 2.7. From the previous calculations, one can see that the resolution of the problem depends on the conductivity of the background medium σ0. More precisely, on how much σ0 is small. In other words, from equations (2.11)–(2.14), we notice that to solve the problem, it is necessary to have κ > 4 + K, for some positive integer K, where κ is given by the relation σ0 = O(epsilonκ). On the other hand, the conductivity coefficient σ0 plays an important role in the resolution of the problem in higher dimensions, on which we notice that our reconstruction algorithm can be generalized to higher dimensions ${\mathbb{R}}^{N}$, if we can find a positive integer K such that

Detailed explanations are given in the appendix.

3. Stability

The main goal in this section is to study the stability of reconstruction of the centers of the absorbers based on the method proposed in [25], on which the authors have used specific wave functions, then, using the reciprocity gap functional and suitable inequalities estimates, one gets the desired result.

3.1. Definitions and notations

Before giving our stability results, we first introduce several notations and definitions needed throughout the rest of the paper.

First, let S = (x, y, z) ∈ Ω and d(Γ, S) be the Euclidean distance between the boundary Γ and S. Let Dj, j = 1, ..., m, be m absorbers with centers

We set

which is strictly positive, since Sj ∈ Ω. Then, we define the set

and let

As denoted before, we take

to be the projections of the centers Sj onto the xy and yz-planes respectively, and set

Then, we introduce the following real coefficients

and take

Equation (3.1)

Thus, δ is strictly positive according to hypothesis (H). Finally, for two points configurations ${R}^{\ell }={\left({R}_{j}^{\ell }\right)}_{1{\leqslant}j{\leqslant}{N}_{\ell }}$ for = 1, 2, we recall the Hausdorff distance between R1 and R2, defined as follows

3.2. Stability estimate for the locations of the absorbers

Let

be two absorber configurations with ${\gamma }_{j}^{\ell }$ defined in (2.10), and consider the following constants

Equation (3.2)

and

Equation (3.3)

where ${\nu }_{i,1}^{0,a}$ and ${\nu }_{j,2}^{0,a}$ are as defined in (2.12):

and

and the function ρ is the solution of (2.6). Then take ϱ > 0 such that

Equation (3.4)

where ${\Vert}\cdot {{\Vert}}_{{H}^{1}\left({\gamma }_{1},{\gamma }_{2}\right)}$ is the classical H1 norm defined by

Now we are able to state our stability result for the absorbers' centers.

Theorem 3.1. Let pk, for k = 1, 2, be the solutions of (1.3) corresponding to two different sets of absorption domains, ${D}_{j}^{k}={\mathbf{S}}_{j}^{k}+{\epsilon}{B}_{j}^{k},1{\leqslant}j{\leqslant}{m}_{k}$, and characterized by the configurations ${\left({\gamma }_{j}^{k},{\mathbf{S}}_{j}^{k}\right)}_{1{\leqslant}j{\leqslant}{m}_{k}}$ with ${\nu }_{j,k}^{0,a}\ne 0$. For k = 1, 2, we take ${f}^{k}{:=}{p}_{{\Gamma}{\times}\left(0,T\right)}^{k}$ to be the corresponding measurements on the boundary Σ = Γ × (0, T), and ${S}^{k}={\left\{{\mathbf{S}}_{j}^{k}\right\}}_{1{\leqslant}j{\leqslant}{m}_{k}}$. Assuming that ${\mathbf{S}}_{j}^{k}\in {{\Omega}}_{\alpha }$, then the following estimate holds,

The proof of theorem 3.1 is based on the following lemma.

Lemma 3.2. Let ${P}^{r,k}={\left\{{P}_{j}^{r,k}\right\}}_{1{\leqslant}j{\leqslant}{m}_{k}}$, for r = a, b, be respectively the corresponding projections on the xy- and yz-planes of Sk. Under the assumptions of theorem 3.1, we have, for r = a, b and k = 1, 2,

Equation (3.5)

Proof. The proof is based on the idea of choosing specific wave functions. In fact, if we take for 1 ⩽ m2,

and

where ρ is defined in (2.6). Then, we observe that Ψ is a solution of (2.1) satisfying (2.5). Thus,

Equation (3.6)

where $\mathcal{R}\left(f,{{\Psi}}_{\ell }\right)$ is defined in (2.3). Also, substituting Ψ in (2.4) we get

Applying the change of variable ${D}_{j}^{2}={\mathbf{S}}_{j}^{2}+{\epsilon}{B}_{j}^{2}$, we obtain

Since κ > 4, we have

Similarly, one has

Subtracting the two previous sums, we get

Therefore, from (3.1), (3.2) and (3.6), we obtain

Moreover, from (3.3) and (3.4), one can easily check that

Finally, we get

Equation (3.7)

Similarly, by considering, first,

and then,

we get

Equation (3.8)

Taking the maximum between (3.7) and (3.8), we get (3.5) for r = a. To prove the second inequality, we consider the following functions

where ρ1 is the solution of (2.6) with the space variable z replaced by x. □

Now we are able to give the proof of theorem 3.1.

Proof. Using the fact that

and

we have

Using the previous lemma, we get the desired result. □

Remark 3.3 (Numerical simulations). The numerical implementation of our algorithm and some extensions of this work, in particular, the investigation of stability in case of variable sound speed, will be considered in some future work. For this reason, we leave the numerical analysis of this inverse problem for some future investigations. Moreover, we would like to compare our results to those obtained in [1] even if they were in the context of photoacoustics.

4. Conclusion

In this paper, we consider the inverse TAT problem, an effective imaging technique for detecting tumors based on the photo-acoustic effect. Unlike PAT, in TAT the biological tissue is subjected to a low frequency radiation which results in a remarkable difference in conductivity between healthy tissues and cancerous ones. In our work, we take advantage of this difference to detect tumors, represented by small absorbers ${\left\{{D}_{j}\right\}}_{1{\leqslant}j{\leqslant}m}$. More precisely, we consider the inverse source problem for the wave equation (1.3), on which we reconstruct these absorbers from the boundary measurements. Unlike prior methods, our method does not follow the quantitative thermoacoustic tomography approach. Instead, we develop a direct reconstruction process based on the algebraic algorithm proposed in [14]. This method allows us, directly from the wave equation, and using only one Cauchy data, to reconstruct modulo epsilonK+4 the number of absorbers m, their centers Sj, and some information on the conductivity coefficient σj. Moreover, our algorithm does not require the knowledge of the Grüneisen coefficient β(x), which is supposed throughout this paper to be variable and unknown with β(x) ∈ L(Ω). Finally we end up with the corresponding stability result.

Acknowledgments

We would like to thank Professor Kui REN for useful discussion on photoacoustic and thermoacoustic problems. The present work has been realized as part of Multi-Ondes projects receiving financial supports from ANR-France.

Appendix:

We show here how the algebraic algorithm we proposed can be generalized to higher dimensions ${\mathbb{R}}^{N}$, under some conditions related to the physical properties of the medium, more precisely, to the conductivity of the background medium σ0.

Let ${\gamma }_{i}^{k}\in \mathbb{R}$, for i = 1, 2 and k = 1, ..., N, such that

We denote by ${\mathbf{S}}_{j}=\left({x}_{j}^{1},{x}_{j}^{2},\dots ,\;{x}_{j}^{N}\right)$ the centers of the absorbers Dj = Sj + epsilonBj, for j = 1, ..., m.

Now, let x = (x1, x2, ..., xn) be a vector of Ω. First, we consider the one-dimensional wave equation

Then, we define, for $n\in \mathbb{N}$, the test functions

that are solutions of the adjoint problem

Equation (4.1)

Following the same procedure as in section 2.2, we get

where R(f, v) is defined in (2.3). Applying the change of variable x = Sj + epsilonτ, where τ = (τ1, ..., τN), one obtains

Equation (4.2)

where

and

Therefore, we have for every $n\in \mathbb{N}$,

Equation (4.3)

where

Equation (4.4)

Our goal now is to solve the equation (4.3) to reconstruct the number of absorbers m, the projections ${P}_{j}^{1}$ of the centers Sj on the x1x2-plane, and the coefficients ${\nu }_{j}^{k,1}$. Again, the equation (4.3) do not allow us to uniquely solve the problem, as the number of unknowns is greater than that of equations. For this purpose, we define for a positive integer K,

on which we need, as in case of ${\mathbb{R}}^{3}$, the condition

to be satisfied in order to get the relations

and

where

Now, assuming that the projections are mutually distinct, one can apply the algebraic algorithm (steps 1–4), as in the case of ${\mathbb{R}}^{3}$, to reconstruct, modulo O(epsilonN+K+1), the number of absorbers m, the projections ${P}_{j}^{1}$ and the coefficients ${\nu }_{j}^{k,1}$. Similarly, we can reconstruct all the projections of the centers Sj on the xixk-planes, for 1 ⩽ i, kN.

As in the case of ${\mathbb{R}}^{3}$, we can uniquely determine the complete locations of the centers Sj modulo O(epsilonN+1). For this purpose, we consider for k = 3, ..., N, the one-dimensional wave equations

Equation (4.5)

and we define, for $n\in \mathbb{N}$, the test functions

Equation (4.6)

and

Equation (4.7)

First, taking k = 3, and following the steps 4, 5, 6 and 7 of the algebraic algorithm, one can determine the components ${x}_{j}^{3}$ up to O(epsilonN+1), which leads to obtaining the projections ${P}_{j}^{2}$ of the centers Sj on the x2x3-plane, and then, by recurrence, we obtain, modulo O(epsilonN+1), the locations Sj for all k = 4, ..., N, and j = 1, ..., m.

Please wait… references are loading.
10.1088/1361-6420/ab8862