Izvestiya: Mathematics
RUS  ENG    JOURNALS   PEOPLE   ORGANISATIONS   CONFERENCES   SEMINARS   VIDEO LIBRARY   PACKAGE AMSBIB  
General information
Latest issue
Forthcoming papers
Archive
Impact factor
Guidelines for authors
Submit a manuscript

Search papers
Search references

RSS
Latest issue
Current issues
Archive issues
What is RSS



Izv. RAN. Ser. Mat.:
Year:
Volume:
Issue:
Page:
Find






Personal entry:
Login:
Password:
Save password
Enter
Forgotten password?
Register


Izvestiya: Mathematics, 2022, Volume 86, Issue 6, Pages 1123–1142
DOI: https://doi.org/10.4213/im9253e
(Mi im9253)
 

This article is cited in 2 scientific papers (total in 2 papers)

Completeness of asymmetric products of harmonic functions and uniqueness of the solution to the Lavrent'ev equation in inverse wave sounding problems

M. Yu. Kokurin

Mari State University, Ioshkar-Ola
References:
Abstract: We prove that the family of all pairwise products of regular harmonic functions in a domain $D \subset \mathbb{R}^3$ and Newtonian potentials of points located on a ray outside $D$ is complete in $L_2(D)$. This result is used for justification of uniqueness of a solution to the linear integral equation to which inverse problems of wave sounding in $\mathbb{R}^3$ are reduced. The corresponding inverse problems are shown to be uniquely solvable in spatially non-overdetermined settings where the dimension of the spatial data support coincides with that of the support of the sought-for function. Uniqueness theorems are used for establishing that the axial symmetry of the input data for the inverse problems under consideration implies that of the solutions to these problems.
Keywords: harmonic function, completeness, inverse problem, hyperbolic equation, linear integral equation, uniqueness of solution, axial symmetry.
Funding agency Grant number
Russian Science Foundation 20-11-20085
Supported by the Russian Science Foundation (project no. 20-11-20085).
Received: 03.08.2021
Revised: 16.11.2021
Russian version:
Izvestiya Rossiiskoi Akademii Nauk. Seriya Matematicheskaya, 2022, Volume 86, Issue 6, Pages 101–122
DOI: https://doi.org/10.4213/im9253
Bibliographic databases:
Document Type: Article
UDC: 517.57+517.518.32+519.968.21
Language: English
Original paper language: Russian

§ 1. Statement of the problem. The main theorem

Let $\mathcal{H}(D)$ be the class of all functions from $C^2(\overline{D})$ harmonic in a bounded domain $D \subset \mathbb{R}^n$, $n \geqslant 2$, and $\mathcal{H}_j = \mathcal{H}_j (D)\subset \mathcal{H}(D)$ be some families of functions from $\mathcal{H}(D)$, $j=1,2$. Below, by a domain we will always mean an open connected set.

The study of uniqueness of solutions to various classes of coefficient inverse problems for equations of mathematical physics is generally facilitated by the ability to answer (to some extent) the following approximation theory problem: under which conditions the family of all pairwise products

$$ \begin{equation*} \mathcal{H}_1 \cdot \mathcal{H}_2 = \{ u_1 u_2\colon u_1 \in \mathcal{H}_1,\, u_2 \in \mathcal{H}_2 \} \end{equation*} \notag $$
(generated by $\mathcal{H}_1$, $\mathcal{H}_2$) is such that, for a fixed function $h\in L_p(D)$, $p\in [2,\infty]$, the equality
$$ \begin{equation} \int_D h(x) u_1(x) u_2(x)\, dx=0 \quad \forall \, u_j \in \mathcal{H}_j,\quad j=1,2, \end{equation} \tag{1.1} $$
implies that $h=0$ almost everywhere in $D$? In the case $p=2$, equality (1.1) means that the family $\mathcal{H}_1 \cdot \mathcal{H}_2$ is dense in $L_2(D)$, and hence any function $v\in L_2(D)$ is arbitrarily well approximated in $L_2(D)$ by finite linear combinations of functions of the form $u_{1 k}(x) u_{2 k}(x)$, where $u_{j k} \in \mathcal{H}_j(D)$, $j=1,2$.

Let us briefly recall some known results in this direction. The first result here is due to A. P. Calderón (1980), who considered the simplest case where each of two families of harmonic functions is just the entire class $\mathcal{H}(D)$.

Proposition 1 (see [1], [2], § 5.3). Let $n \geqslant 2$, let $\mathcal{H}_1=\mathcal{H}_2=\mathcal{H}(D)$, and let $h\in L_2(D)$. Then (1.1) implies that $h=0$ almost everywhere in $D$.

From the point of view of approximation theory, it seems natural in this result to reduce the families $\mathcal{H}_1$, $\mathcal{H}_2$ without impairing the completeness property of pairwise products. Below, we will see that such a reduction plays an important role in establishing uniqueness of solutions for a number of applied inverse problems with minimal requirements to the dimension of their input data set. We note the following result pertaining to harmonic functions with fixed values on a part of the boundary. We set

$$ \begin{equation*} O_{r}(x_0)=\{ x\in \mathbb{R}^n\colon |x-x_0|< r \}. \end{equation*} \notag $$

Proposition 2 (see [3], Theorem 7.1). Let $n \geqslant 2$, let $x_0 \in \partial D$, $\partial D \in C^{\infty}$, and let

$$ \begin{equation*} \mathcal{H}_1=\mathcal{H}_2=\{ u\in \mathcal{H}(D)\colon u(x)=0,\, x\in \partial D \setminus O_{\varepsilon}(x_0) \}. \end{equation*} \notag $$
Then there exists $\delta=\delta(\varepsilon)>0$ such that if $h\in L_{\infty}(D)$, then equality (1.1) implies $h=0$ almost everywhere in $D\cap O_{\delta}(x_0)$.

There are generalizations of Proposition 1 in which the Laplace equation $\Delta u=0$ is replaced by more general equations $P(-i\partial)u+c(x)u=0$, $x\in D$. In this regard, we mention the following result for elliptic second-order equations. Let $c_j \in L_{\infty}(D)$ and let $\mathcal{H}_j$ be the family of all generalized solutions to the equation $\Delta u+c_j(x) u=0$ that lie in $L_2(D)$, $j=1,2$.

Proposition 3 (see [2], Corollary 5.3.5). Let $n \geqslant 3$, $h\in L_{\infty}(D)$, and let (1.1) be met. Then $h=0$ almost everywhere in $D$.

It is worth pointing out that in Proposition 3 and in its generalizations from [2], § 5.3, the families $\mathcal{H}_1$, $\mathcal{H}_2$ are formed by the solutions of generally different equations.

It is easily seen that, in the above examples, the families $\mathcal{H}_1$, $\mathcal{H}_2$ coincide with the set of all solutions of the equation under consideration in the domain (possibly, with a boundary condition on a part of the boundary $\partial D$). In the case of interest, in which both families are generated by the same Laplace equation, these families are equal. In this sense, Proposition 1 establishes completeness of the symmetric products of harmonic functions. It is also known that the completeness property in Proposition 1 remains valid also in the case where one of the involved families $\mathcal{H}_1$, $\mathcal{H}_2$ is much narrower than the other one. In these cases, we will speak about asymmetric products of functions harmonic in the domain. The next result, which strengthens Proposition 1, established completeness of the asymmetric products of harmonic functions in the above sense.

Proposition 4 (see [4]). Let $n \geqslant 3$, let $L$ be a straight line in $\mathbb{R}^n$, $L \cap \overline{D}=\varnothing$, let $Y$ be an open set on $L$, and let

$$ \begin{equation*} \mathcal{H}_1=\mathcal{H}(D), \qquad \mathcal{H}_2=\biggl\{ \frac{1}{|x-y|^{n-2}}\colon y\in Y \biggr\}. \end{equation*} \notag $$
Then the family $\mathcal{H}_1 \cdot \mathcal{H}_2$ is dense in $L_{2}(D)$.

In Proposition 4, the class $\mathcal{H}_2$ is formed only by Newtonian potentials of points of an open one-dimensional subset of a straight line. Note that a similar result also holds for the Helmholtz equation [5]. The motivation of the present paper is to remove the condition $L \cap \overline{D}=\varnothing$, which played an important role in the proof of Proposition 4, and which, at the same time, does not seem essential. Below, we will consider the case $n=3$, which has direct applications to inverse wave sounding problems in $\mathbb{R}^3$.

The key result of the present paper is the following theorem, which strengthens Proposition 4 for $n=3$.

Theorem 1. Let $L$ be a straight line in $\mathbb{R}^3$, let $Y$ be an open set lying in the unbounded component of the complement $L\setminus \overline{D}$, and let

$$ \begin{equation*} \mathcal{H}_1=\mathcal{H}(D), \qquad \mathcal{H}_2=\biggl\{ \frac{1}{|x-y|}\colon y\in Y \biggr\}. \end{equation*} \notag $$
Then the family $\mathcal{H}_1 \cdot \mathcal{H}_2$ is dense in $L_{2}(D)$.

The scheme of our further exposition is as follows. Theorem 1 will be proved in § 2. In subsequent sections, we will consider applications of Theorem 1 to the study of uniqueness of solutions to inverse wave sounding problems in various settings. In these applications, Theorem 1 will be used in analysis of uniqueness of the solution to the Lavrent’ev linear integral equation, which is naturally related to the inverse problems under considerations. In § 3, we will briefly recall the derivation of this equation, and then employ Theorem 1 to verify uniqueness of the solution of the coefficient inverse problem for a hyperbolic equation in the spatially non-overdetermined setting. In § 4, an inverse problem of wave sounding by means of time-harmonic fields in the phaseless setting is investigated. We will show that Theorem 1 (in contrast to other results in this field) is capable of reducing the dimension of the data support, and also provides a non-overdetermined statement of the inverse problem. In § 5, for the inverse problems under consideration, we will establish theorems on the relation of axial symmetry of input data with that of the probed inhomogeneity. The corresponding results are similar to S. N. Karp’s theorem and its variants, which link the spherical symmetry of the scattering amplitude with that of the inhomogeneity under study.

§ 2. Proof of Theorem 1

Let us formulate the result to be proved in a form convenient for further analysis. In view of (1.1), it will be sufficient to verify that the relation

$$ \begin{equation} \int_D \frac{h(x) u(x)\, dx}{|x-y|}=0 \quad \forall \, y\in Y, \quad u\in \mathcal{H}(D), \end{equation} \tag{2.1} $$
with $h\in L_2(D)$ implies $h(x)=0$ almost everywhere in $D$. Let $f(y)$ be the left-hand side of (2.1). The function $f(y)$ is harmonic with respect to $y \in \mathbb{R}^3 \setminus \overline{D}$, and hence, is real analytic outside $\overline{D}$. It follows that the restriction of $f$ to $L \setminus \overline{D}$ is also real analytic. Hence the equality $f(y)=0$, $y\in Y$, extends by analyticity to the entire unbounded component of the straight line $L$ containing the interval $Y$ (see [6], Ch. I, § 1). On the straight line $L$, consider an arbitrary point and a ball with centre at this point whose interior contains $\overline{D}$. Without loss of generality it can be also assumed that the chosen point is the origin, and the radius of the ball containing $\overline{D}$ is $R<1$. So, $|x|\leqslant R$ for all $x\in \overline{D}$. Assume, for definiteness, that $L=\{ (0,0,x_3)\colon x_3 \in \mathbb{R}\}$. We set $L_a^+=\{ x\in L\colon x_3 \geqslant a\}$. It can be assumed that $f(y)=0$ if $y$ belongs to the unbounded component of $L_0^+ \setminus \overline{D}$. This component contains the ray $L_R^+$, and hence by (2.1) we have
$$ \begin{equation} \int_D \frac{h(x) u(x)\, dx}{|x-y|}=0 \quad \forall \, y\in L_R^+, \quad u\in \mathcal{H}(D). \end{equation} \tag{2.2} $$

Let us now proceed with the proof of Theorem 1 with due account of the above remarks.

Proof of Theorem 1. We need to prove that equality (2.2) implies $h(x)=0$ almost everywhere in $D$. We equip $\mathbb{R}^3=(x_1,x_2,x_3)$ with the spherical coordinates $x_1=\rho\sin \theta \cos\varphi$, $x_2=\rho\sin \theta \sin\varphi$, $x_3=\rho\cos\theta$, where $\theta \in [0,\pi]$, $\varphi \in [0,2\pi)$. Consider the spherical functions $\{Y_{lm} \}$, $l=0,1,\dots$, $|m|\leqslant l$, defined by (see [7], § 23, [8], § 5.2)
$$ \begin{equation} \begin{aligned} \, Y_{lm}(\theta,\varphi)&=e^{i m\varphi} \sqrt{\frac{2l+1}{4\pi}\, \frac{(l-m)!}{(l+m)!}}\, P_l^m (\cos\theta), \qquad 0\leqslant m \leqslant l, \\ Y_{l, -m}(\theta,\varphi)&=(-1)^m Y_{lm}(\theta,-\varphi). \end{aligned} \end{equation} \tag{2.3} $$
Here,
$$ \begin{equation*} P_l^m(t)=(1-t^2)^{m/2}\, \biggl( \frac{d}{dt} \biggr)^m P_l(t) \end{equation*} \notag $$
is the associated Legendre function, $P_l(t)=P_l^0(t)$ is the Legendre polynomial of degree $l$ (see [7], § 23). From (2.3) we have
$$ \begin{equation} P_l (\cos \theta)= \sqrt{ \frac{4\pi}{2l+1}}\, Y_{l0}(\theta,\varphi), \qquad \varphi\in [0,2\pi). \end{equation} \tag{2.4} $$

Given $y \in L_1^{+}$, we set $\tau=|y|^{-1}$. Then $\tau$ assumes arbitrary values in $(0,1]$. Since $L_1^{+} \subset L_R^+$, equality (2.2) holds for all $y\in L_1^{+}$. For arbitrary points $x\in D$, $y\in L_1^{+}$, we get

$$ \begin{equation*} \frac{1}{|x-y|}=\frac{\tau}{|\tau x- y^0|}, \qquad y^0=(0,0,1). \end{equation*} \notag $$
Let $(\rho,\theta,\varphi)$ be the spherical coordinate of a point $x\in D$. Since $|\tau x|\leqslant|x|\leqslant R< 1=|y^0|$ for all $x\in D$, we have (see [7], § 8)
$$ \begin{equation} \frac{1}{|x-y|}=\frac{\tau}{\sqrt{1-2\tau\rho \cos\theta+(\tau \rho)^2}}=\tau \sum_{l=0}^{\infty} P_l(\cos\theta) (\tau \rho)^l. \end{equation} \tag{2.5} $$

Substituting (2.5) into (2.2), we find that

$$ \begin{equation} \sum_{l=0}^{\infty} \tau^l \biggl( \int_D h(x) u(x) \rho^l P_l(\cos\theta)\, dx \biggr)=0 \end{equation} \tag{2.6} $$
for all $\tau \in (0,1]$. Here, the change in the order of integration and summation is legitimate because of the uniform convergence in (2.5) with respect to $x \in D$, $\tau\in [0,1]$. Since $|P_l(t)|\leqslant 1$ for $t\in [-1,1]$ (see [7], § 7) and $\rho\leqslant R<1$ for $x\in D$, the power series in (2.6) has positive radius of convergence with respect to $\tau$. As a result,
$$ \begin{equation} \int_D h(x) u(x) \rho^l P_l(\cos\theta)\, dx =0 \end{equation} \tag{2.7} $$
for all integer $l \geqslant 0$ and all functions $u\in \mathcal{H}(D)$. Using (2.4), we rewrite equality (2.7) as
$$ \begin{equation} \int_D h(x) u(x) \rho^l Y_{l0}(\theta,\varphi)\, dx =0, \qquad l\geqslant 0. \end{equation} \tag{2.8} $$
As a regular harmonic function $u$ in (2.8), consider the harmonic polynomial
$$ \begin{equation*} u(x)=\rho^L Y_{LM}(\theta,\varphi) \end{equation*} \notag $$
with arbitrary integer $L\geqslant 0$ and $|M|\leqslant L$. So, for all integer $l, L \geqslant 0$, and $|M|\leqslant L$,
$$ \begin{equation} \int_D h(x) \rho^{L+l} Y_{LM}(\theta,\varphi) Y_{l0}(\theta,\varphi)\, dx =0. \end{equation} \tag{2.9} $$

We set

$$ \begin{equation*} S=\{ x\in \mathbb{R}^3\colon |x|=1 \}. \end{equation*} \notag $$
For any integrable compactly supported function $F=F(x)$, we have (see [9], Ch. I, § 4.14)
$$ \begin{equation} \int_{\mathbb{R}^3} F(x)\, dx=\int_S \biggl( \int_0^{\infty} \rho^2 F(\rho\eta)\, d\rho \biggr)\, d\sigma =\int_0^{\infty} \rho^2 \biggl( \int_S F(\rho\eta)\, d\sigma \biggr) \, d\rho, \end{equation} \tag{2.10} $$
where $\eta$ varies over the sphere $S$, and $d\sigma$ is the Lebesgue measure on $S$. Assuming that $h$ is extended by zero outside $D$ and using (2.10), we can write equality (2.9) as
$$ \begin{equation} \int_0^{\infty} \rho^{L+l+2} \biggl( \int_S h(\rho\eta) Y_{LM}(\theta,\varphi) Y_{l0}(\theta,\varphi)\, d\sigma \biggr)\, d\rho=0. \end{equation} \tag{2.11} $$
In (2.11), the vector $\eta=x/|x|\in S$ has spherical coordinates $(1,\theta,\varphi)$. Note that by the above assumptions the outer integral in (2.11) is taken in fact over the interval $[0,R].$ The orthonormal system of spherical functions $\{ Y_{lm} \}$, $l=0,1,\dots$, $|m|\leqslant l$, is complete in $L_2(S)$, and hence, for almost all $\rho\geqslant 0$,
$$ \begin{equation} h(\rho\eta) =\sum_{\lambda=0}^{\infty} \sum_{\mu=-\lambda}^{\lambda} a_{\lambda \mu}(\rho) Y_{\lambda\mu}(\theta,\varphi), \qquad \sum_{\lambda=0}^{\infty} \sum_{\mu=-\lambda}^{\lambda}|a_{\lambda \mu}(\rho)|^2 <\infty. \end{equation} \tag{2.12} $$
Indeed, $h(r\eta)$ lies in $L_2$ qua a function of $(r,\theta,\varphi)$. Hence its restriction to the plane $r=\rho$ belongs to the space $L_2(S)$ for almost all $\rho \geqslant 0$ (see, for example, Lemma 1.3.8 in [10]).

From (2.11) and (2.12) it follows that, for all integer $L,l \geqslant 0$, and $|M|\leqslant L$,

$$ \begin{equation} \int_0^{\infty} \rho^{L+l+2} \biggl\{ \sum_{\lambda=0}^{\infty} \sum_{\mu=-\lambda}^{\lambda} a_{\lambda \mu}(\rho) \biggl( \int_S Y_{\lambda\mu}(\theta,\varphi) Y_{LM}(\theta,\varphi) Y_{l0}(\theta,\varphi)\, d\sigma \biggr) \biggr\}\, d\rho=0. \end{equation} \tag{2.13} $$
Here, the change in the order of integration and summation is legitimate inasmuch as the system of spherical functions is closed in $L_2(S)$ (see [11], Ch. VII, § 3).

In our further analysis, we will require certain properties of the inner integral in (2.13). Integrals of this kind appear in various fields of quantum mechanics for evaluation of matrix elements of vector physical quantities (see [12], Ch. XIV). Let us recall some properties of this integral (see, for example, [8], § 5.9, [12], § 107). First of all, this integral can be explicitly written as

$$ \begin{equation*} \begin{aligned} \, &\int_S Y_{\lambda\mu}(\theta,\varphi) Y_{LM}(\theta,\varphi) Y_{l0}(\theta,\varphi)\, d\sigma \\ &\qquad=\sqrt{\frac{(2\lambda+1)(2L+1)(2l+1)}{4\pi}}\, \begin{pmatrix} \lambda & L & l \\ 0 & 0 & 0 \end{pmatrix}\begin{pmatrix} \lambda & L & l \\ \mu & M & 0 \end{pmatrix}. \end{aligned} \end{equation*} \notag $$
So, using (2.13) we obtain
$$ \begin{equation} \int_0^{\infty} \rho^{L+l+2} \biggl\{ \sum_{\lambda=0}^{\infty} \sum_{\mu=-\lambda}^{\lambda} b_{\lambda \mu}(\rho)\begin{pmatrix} \lambda & L & l\\ 0 & 0 & 0 \end{pmatrix}\begin{pmatrix} \lambda & L & l\\ \mu & M & 0 \end{pmatrix} \biggr\}\, d\rho=0, \end{equation} \tag{2.14} $$
where we set
$$ \begin{equation*} b_{\lambda \mu}(\rho)=\sqrt{2\lambda+1}\, a_{\lambda \mu}(\rho). \end{equation*} \notag $$
Here and in what follows,
$$ \begin{equation} \begin{pmatrix} j_1 & j_2 & j_3 \\ m_1 & m_2 & m_3 \end{pmatrix}=\begin{pmatrix} j_3 & j_1 & j_2 \\ m_3 & m_1 & m_2 \end{pmatrix}=\begin{pmatrix} j_2 & j_3 & j_1 \\ m_2 & m_3 & m_1 \end{pmatrix} \end{equation} \tag{2.15} $$
denotes the $3jm$-symbol (the Wigner coefficient), whose explicit representation for arbitrary $m_1$, $m_2$, $m_3$ is too cumbersome and hence is not given (for more details, see [8], § 8.1, § 8.2, [12], § 106, and also (2.18)). Coefficients (2.15) are real functions of the integer arguments $(j_1,\dots,m_3)$. We will also need the Clebsch–Gordan coefficients $C_{j_1 m_1 \, j_2 m_2}^{j_3 m_3}$, which are related to the Wigner coefficients (2.15) as follows:
$$ \begin{equation} C_{j_1 m_1 \, j_2 m_2}^{j_3 m_3}=(-1)^{j_1-j_2+m_3} \sqrt{2j_3+1}\, \begin{pmatrix} j_1 & j_2 & j_3 \\ m_1 & m_2 & -m_3 \end{pmatrix}. \end{equation} \tag{2.16} $$

Here, the integer $j_k$, $m_k$ satisfy $j_k \geqslant 0$, $|m_k|\leqslant j_k$, $1\leqslant k\leqslant 3$. According to [8], pp. 235–237, the Wigner coefficient (2.15) vanishes if at least one of the constraints

$$ \begin{equation} |j_1-j_2|\leqslant j_3 \leqslant j_1+j_2, \qquad m_1+m_2+m_3=0 \end{equation} \tag{2.17} $$
is violated. In addition (see [8], § 8.5.2, [12], § 106), for integer $p\geqslant 0$,
$$ \begin{equation} \begin{aligned} \, &\begin{pmatrix} j_1 & j_2 & j_3 \\ 0 & 0 & 0 \end{pmatrix} \nonumber \\ &=\begin{cases} \biggl( \dfrac{(j_1+j_2-j_3)!\, (j_1-j_2+j_3)!\, (-j_1+j_2+j_3)!}{(2p+1)!} \biggr)^{1/2} & \\ \quad\times\dfrac{(-1)^p p!}{(p-j_1)!\, (p-j_2)!\, (p-j_3)!}, &j_1+j_2+j_3=2p, \\ 0, &j_1+j_2+j_3=2p+1. \end{cases} \end{aligned} \end{equation} \tag{2.18} $$
For reference, we also recall the integral representation of the Clebsch–Gordan coefficients (see [8], § 8.3)
$$ \begin{equation} \begin{aligned} \, &C_{a\alpha\, b\beta}^{c\gamma}=\frac{(-1)^{a-c+\beta}}{2^{J+1}} \nonumber \\ &\times \biggl( \frac{(c+\gamma)!\, (J-2c)!\, (J+1)!\, (2c+1)}{(a-\alpha)!\, (a+\alpha)!\, (b-\beta)!\, (b+\beta)!\, (c-\gamma)!\, (J-2a)!\, (J-2b)!} \biggr)^{1/2} \nonumber \\ &\times\int_{-1}^1 (1-t)^{a-\alpha} (1+t)^{b-\beta} \biggl( \frac{d}{dt} \biggr)^{c-\gamma} [(1-t)^{J-2a} (1+t)^{J-2b}]\, dt, \qquad J=a+b+c. \end{aligned} \end{equation} \tag{2.19} $$

We will use (2.16) and (2.19) to verify that

$$ \begin{equation} \begin{pmatrix} N-2s & N-s & s\\ \mu & -\mu & 0 \end{pmatrix} \neq 0 \end{equation} \tag{2.20} $$
for any integer $N, s\geqslant 0$ and $\mu$ such that
$$ \begin{equation*} |\mu|\leqslant N-2s. \end{equation*} \notag $$

From (2.16) it follows that (2.20) differs from the Clebsch–Gordan coefficient $C_{N-2s \,\, \mu \,\, N-s \,\, -\mu}^{s 0}$ by a nonzero factor. In turn, by (2.19) this factor is equal, up to a nonzero factor, to the integral

$$ \begin{equation*} \int_{-1}^1 (1-t)^{N-s-\mu} (1+t)^{N-s+\mu}\, dt >0. \end{equation*} \notag $$
Now the required result (2.20) follows.

For an arbitrary fixed $\mu_0 \in \mathbb{Z}$, we set $M=-\mu_0$ in (2.14). We will assume that

$$ \begin{equation} L-l \geqslant|\mu_0|. \end{equation} \tag{2.21} $$
From (2.15), (2.17), (2.18) we have
$$ \begin{equation*} \begin{pmatrix} \lambda & L & l \\ \mu & -\mu_0 & 0 \end{pmatrix}= \begin{pmatrix} L & l & \lambda \\ -\mu_0 & 0 & \mu \end{pmatrix}=0 \end{equation*} \notag $$
for any $\lambda$ if $\mu\neq \mu_0$, and with the violated condition $|L-l|\leqslant \lambda \leqslant L+l$ for $\mu=\mu_0$.

Using (2.21), we conclude that in the inner sum in (2.14) only the terms with $\mu=\mu_0$ may nonvanish, and in the outer sum, only the terms for which $L-l\leqslant \lambda \leqslant L+ l$. Hence (2.14) assumes the form

$$ \begin{equation} \int_0^{\infty} \rho^{L+l+2} \biggl\{ \sum_{\lambda=L-l}^{L+l} b_{\lambda \mu_0}(\rho) \begin{pmatrix} \lambda & L & l \\ 0 & 0 & 0 \end{pmatrix}\begin{pmatrix} \lambda & L & l \\ \mu_0 & -\mu_0 & 0 \end{pmatrix} \biggr\}\, d\rho=0. \end{equation} \tag{2.22} $$
Equality (2.22) holds for all nonnegative integers $L$, $l$ satisfying condition (2.21).

We now fix $K=L+l$ so that

$$ \begin{equation} K \geqslant|\mu_0|. \end{equation} \tag{2.23} $$
Let us analyze equalities (2.22) corresponding to various representations of $K$ as the sum $L+l$ under condition (2.21).

We first consider the case $L=K$, $l=0$. In this case, (2.21) holds by (2.23). The interval of variation of $\lambda$ in (2.22) reduces to the single point $\lambda=K$. Therefore, the sum in (2.22) reduces to the single term

$$ \begin{equation*} b_{K\mu_0}(\rho) \begin{pmatrix} K & K & 0 \\ 0 & 0 & 0 \end{pmatrix}\begin{pmatrix} K & K & 0\\ \mu_0 & -\mu_0 & 0 \end{pmatrix}. \end{equation*} \notag $$
Here the first Wigner coefficient is nonzero by (2.18), and the second one is nonzero by (2.20) for $N=K$, $s=0$, $\mu=\mu_0$. So, the function $b_{K \mu_0}(\rho)$ enters (2.22) with nonzero coefficient, and hence
$$ \begin{equation} \int_0^{\infty} \rho^{K+2} b_{K\mu_0}(\rho)\, d\rho=0. \end{equation} \tag{2.24} $$

Now let $L=K-1$, $l=1$. Here, in addition to (2.23), we have

$$ \begin{equation*} K-2 \geqslant|\mu_0|. \end{equation*} \notag $$
This inequality secures (2.21). In this case, the interval of variation of $\lambda$ in (2.22) is $K-2\leqslant \lambda \leqslant K$, and the sum in (2.22) has the form
$$ \begin{equation*} \begin{aligned} \, &b_{K-2 \, \mu_0}(\rho)\begin{pmatrix} K-2 & K-1 & 1 \\ 0 & 0 & 0 \end{pmatrix}\begin{pmatrix} K-2 & K-1 & 1 \\ \mu_0 & -\mu_0 & 0 \end{pmatrix} \\ &\qquad+ b_{K-1 \, \mu_0}(\rho) \begin{pmatrix} K-1 & K-1 & 1 \\ 0 & 0 & 0 \end{pmatrix}\begin{pmatrix} K-1 & K-1 & 1 \\ \mu_0 & -\mu_0 & 0 \end{pmatrix} \\ &\qquad+b_{K\mu_0}(\rho) \begin{pmatrix} K & K-1 & 1 \\ 0 & 0 & 0 \end{pmatrix}\begin{pmatrix} K & K-1 & 1 \\ \mu_0 & -\mu_0 & 0 \end{pmatrix}. \end{aligned} \end{equation*} \notag $$
Here the coefficient multiplying $b_{K-1 \,\mu_0}(\rho)$ vanishes by (2.18), and the coefficient multiplying $b_{K-2 \, \mu_0}(\rho)$ is nonzero. This follows from (2.18) and (2.20), where in (2.20) one should put $N=K$, $s=1$, $\mu=\mu_0$. Using (2.22) and the already proved equality (2.24), we conclude that
$$ \begin{equation} \int_0^{\infty} \rho^{K+2} b_{K-2 \,\mu_0}(\rho)\, d\rho=0. \end{equation} \tag{2.25} $$

Similarly, for $L=K-2$, $l=2$, under the assumption

$$ \begin{equation*} K-4 \geqslant|\mu_0|, \end{equation*} \notag $$
we get in (2.22) the sum of functions $b_{K-4 \, \mu_0}(\rho)$, $b_{K-3 \, \mu_0}(\rho)$, $b_{K-2 \, \mu_0}(\rho)$, $b_{K-1\, \mu_0}(\rho)$, $b_{K\, \mu_0}(\rho)$, in which the coefficients
$$ \begin{equation*} \begin{gathered} \, \begin{pmatrix} K-3 & K-2 & 2 \\ 0 & 0 & 0 \end{pmatrix}\begin{pmatrix} K-3 & K-2 & 2 \\ \mu_0 & -\mu_0 & 0 \end{pmatrix}, \\ \begin{pmatrix} K-1 & K-2 & 2 \\ 0 & 0 & 0 \end{pmatrix}\begin{pmatrix} K-1 & K-2 & 2 \\ \mu_0 & -\mu_0 & 0 \end{pmatrix} \end{gathered} \end{equation*} \notag $$
multiplying $b_{K-3 \, \mu_0}(\rho)$, $b_{K-1 \, \mu_0}(\rho)$ are all zero, because the first terms in the products vanish in view of (2.18). On the other hand, the coefficient
$$ \begin{equation*} \begin{pmatrix} K-4 & K-2 & 2 \\ 0 & 0 & 0 \end{pmatrix}\begin{pmatrix} K-4 & K-2 & 2 \\ \mu_0 & -\mu_0 & 0 \end{pmatrix} \end{equation*} \notag $$
multiplying $b_{K-4 \, \mu_0}(\rho)$ is nonzero. This follows from (2.18) and (2.20). In view of (2.24) and (2.25), from (2.22) we have
$$ \begin{equation*} \int_0^{\infty} \rho^{K+2} b_{K-4 \, \mu_0}(\rho)\, d\rho=0. \end{equation*} \notag $$

Continuing in this way, we find that

$$ \begin{equation} \int_0^{\infty} \rho^{K+2} b_{K-2j \, \mu_0}(\rho)\, d\rho=0 \end{equation} \tag{2.26} $$
for all $j=0,1,\dots$ such that $K-2j\geqslant| \mu_0|$. In the proof of (2.26) we used the relation
$$ \begin{equation*} \begin{pmatrix} K-2j & K-j & j \\ 0 & 0 & 0 \end{pmatrix}\begin{pmatrix} K-2j & K-j & j \\ \mu_0 & -\mu_0 & 0 \end{pmatrix} \neq 0, \qquad j=0,1,\dots, \biggl[\frac{K-|\mu_0|}2\biggr], \end{equation*} \notag $$
which follows from (2.18) and (2.20). Here the square brackets denote the integral part of a number.

We now fix an integer $\lambda_0 \geqslant 0$ such that $\lambda_0 \geqslant|\mu_0|$. From (2.26) for $K=\lambda_0$, $j=0$ we have

$$ \begin{equation*} \int_0^{\infty} \rho^{\lambda_0+2} b_{\lambda_0\mu_0}(\rho)\, d\rho=0. \end{equation*} \notag $$
Letting $K=\lambda_0+2$ and $j=1$ in (2.26), and then putting $K=\lambda_0+4$ and $j=2$, we conclude that
$$ \begin{equation*} \int_0^{\infty} \rho^{\lambda_0+4} b_{\lambda_0\mu_0}(\rho)\, d\rho=0, \qquad \int_0^{\infty} \rho^{\lambda_0+6} b_{\lambda_0\mu_0}(\rho)\, d\rho=0. \end{equation*} \notag $$
Continuing similarly, we obtain
$$ \begin{equation} \int_0^{\infty} \rho^{\lambda_0+2j+2} b_{\lambda_0\mu_0}(\rho)\, d\rho=0, \qquad j=0,1,\dots\,. \end{equation} \tag{2.27} $$
The series with the terms $(\lambda_0+2j+2)^{-1}$, $j=0,1,\dots$, is divergent, and so, using (2.27) and Müntz’s theorem [13], § 27, we conclude that $b_{\lambda_0\mu_0}(\rho)=0$ for almost all $\rho \geqslant 0$. Hence
$$ \begin{equation*} a_{\lambda_0\mu_0}(\rho)=0. \end{equation*} \notag $$
Since $\lambda_0 \geqslant 0$ and $|\mu_0|\leqslant \lambda_0$ are arbitrary, from (2.12) we have $h(\rho\eta)=0$ almost everywhere on $S$ for almost all $\rho \geqslant 0$, that is, $h(x)=0$ almost everywhere in the domain $D$. This completes the proof of Theorem 1.

Let us now consider applications of this theorem to uniqueness of solutions to some coefficient inverse problems.

§ 3. The Lavrent’ev equation

Let us study the inverse wave sounding problem with bounded inhomogeneity and a set of point sources not contained in this inhomogeneity [14]. The wave field $u(x,t)=u(x,t;y)$ excited by a source at a point $y$ is determined by the solution of the Cauchy problem

$$ \begin{equation} \begin{gathered} \, \frac{1}{c^2(x)} u_{tt}(x,t) = \Delta u(x,t) - \delta(x - y) g(t),\qquad x\in\mathbb{R}^3, \quad t\geqslant 0 ; \\ u(x,0)=u_t(x,0) = 0, \quad x\in\mathbb{R}^3. \nonumber \end{gathered} \end{equation} \tag{3.1} $$
Here $c(x) > 0$ is the signal propagation velocity at $x\in\mathbb{R}^3$. A typical situation for many applied inverse problems of wave tomography is that $c(x)\equiv c_0$ outside an a priori given bounded domain $D\subset \mathbb{R}^3$, where the constant $c_0$ is known, and $c(x)$ for $x\in D$ should be determined. This model corresponds to sounding of a bounded inhomogeneity in a homogeneous medium. In what follows, we will assume that ${\mathbb R}^n \setminus \overline{D}$ is a domain, $c \in L_{\infty}(D)$, and $g$ is continuous and satisfies
$$ \begin{equation} \int_0^{\infty} g(t)\,dt \neq 0, \qquad |g(t)|\leqslant C_0 e^{-\beta t},\quad \beta > 0,\quad t \geqslant 0. \end{equation} \tag{3.2} $$
In order to find $c(x)$ for $x\in D$, the wave field $u = u(x,t;y)$ is measured for $t > 0$ at the points $x=z \in Z$, where $Z\subset \mathbb{R}^3$ is the set of detectors, $Z\cap \overline D = \varnothing$. We will assume that a set of sources $y\in Y$, $Y\cap \overline D = \varnothing$, is used for sounding. The sets $Y$ and $Z$ are assumed to be domains on manifolds in $\mathbb{R}^3$, $Y \cap Z=\varnothing$. We denote this inverse problem by $\{ Y, Z \}$. Since one knows a priori that the observed wave field corresponds to some velocity function $c(x)$, the above inverse problem has at least one solution. So, below we will be concerned with the problem of whether this function can be uniquely recovered from the observation data. In the present paper, the ‘uniqueness of solution’ for $\{ Y, Z\}$ and for other inverse problems will be understood in this sense.

For a broad class of partial differential equations, Lavrent’ev [15], [16] proposed an approach to solution of nonlinear coefficient inverse problems capable of reducing such problems to linear integral equations. Let us illustrate his approach for the inverse problem $\{ Y, Z \}$. For an integrable function $f = f(t)$, $t \geqslant 0$, we recall that the Laplace transform is defined by $\widetilde{f}(p) = \int_0^{\infty} e^{-pt} f(t)\, dt$. We will assume that all the functions $u(z,t;y)$, $y\in Y$, and their derivatives with respect to $t$ up to the second order inclusively decay sufficiently rapidly as $t\to +\infty$ for $z\in Z$, so that their Laplace transforms exist. In addition, we will assume that $u(x,t;y)\to 0$ as $|x|\to \infty$ uniformly in $y\in Y$ and $t > 0$. The conditions on $c(x)$ to satisfy these requirements were discussed in [17], [18]. Setting

$$ \begin{equation} \xi(x) = \frac{1}{c^2(x)} - \frac{1}{c_0^2},\qquad x\in D, \end{equation} \tag{3.3} $$
we rewrite equation (3.1) as
$$ \begin{equation} \Delta u - \frac{1}{c_0^2}\, u_{tt}(x,\,t) = \xi(x) u_{tt}(x,t) + \delta(x - y) g(t). \end{equation} \tag{3.4} $$
According to (3.3), the function $c$ is uniquely defined from $\xi$, and so, we will deal with finding $\xi(x)$ for $x\in D$. Applying the Laplace transform with respect to time to both sides of (3.4), this gives
$$ \begin{equation} \Delta \widetilde{u}(x,p;y) - \frac{p^2}{c_0^2}\,\widetilde{u}(x,p;y) = p^2\xi(x)\widetilde{u}(x,p;y) + \widetilde{g}(p)\delta(x - y),\qquad x\in \mathbb{R}^3. \end{equation} \tag{3.5} $$
By the assumptions, $\widetilde{u}(x,p;y)\to 0$ as $|x|\to\infty$ for $y\in Y$, $p\geqslant 0$. Let
$$ \begin{equation} G(x, x_0; p) = -\frac{e^{-(p/c_0)|x - x_0|}}{4\pi|x - x_0|} \end{equation} \tag{3.6} $$
be the fundamental solution (Green function) of the equation
$$ \begin{equation*} \Delta u(x) - \frac{p^2}{c_0^2} u(x) = \delta(x - x_0), \qquad p \geqslant 0, \end{equation*} \notag $$
such that $u(x)\to 0$ as $|x|\to \infty$. From (3.5) in view of (3.6) we have
$$ \begin{equation} \widetilde{u}(x,p;y) = \widetilde{g}(p) G(x,y;p)+p^2 \int_{D} G(x, x'; p) \widetilde{u}(x',p; y) \xi(x')\, dx' ,\qquad x\in\mathbb{R}^3. \end{equation} \tag{3.7} $$
We set
$$ \begin{equation} H(x,y;p)=\widetilde{g}(p) G(x,y;p). \end{equation} \tag{3.8} $$
Twice differentiating equality (3.7) with respect to $p$ and using (3.8), we get
$$ \begin{equation} \begin{aligned} \, &\widetilde{u}_{pp}(x,p;y)=H_{pp}(x,y;p)+2\int_D G(x,x';p) \widetilde{u}(x',p;y) \xi(x')\, dx' \nonumber \\ &\qquad +4p \int_D [ G_p(x,x';p) \widetilde{u}(x',p;y)+G(x,x';p) \widetilde{u}_p(x',p;y)] \xi(x')\, dx' \nonumber \\ &\qquad+p^2 \int_D [G_{pp}(x,x';p) \widetilde{u}(x',p;y)+ 2 G_p(x,x';p) \widetilde{u}_p(x',p;y) \nonumber \\ &\qquad\qquad\qquad +G(x,x';p) \widetilde{u}_{pp}(x',p;y)] \xi(x')\, dx'. \end{aligned} \end{equation} \tag{3.9} $$
Putting $p=0$ in (3.7) and (3.9), we obtain
$$ \begin{equation*} \begin{gathered} \, \widetilde{u}(x,0;y)=H(x,y;0)=\widetilde{g}(0) G(x,y;0), \\ \begin{aligned} \, \widetilde{u}_{pp}(x,0;y) &=\int_0^{\infty} t^2 u(x,t;y)\, dt=H_{pp}(x,y;0)+2\int_D G(x,x';0) \widetilde{u}(x',0;y) \xi(x')\, dx' \\ &=H_{pp}(x,y;0)+ 2\widetilde{g}(0) \int_D G(x,x';0) G(x',y;0) \xi(x')\, dx'. \end{aligned} \end{gathered} \end{equation*} \notag $$
Now from (3.6) it follows that the required function $\xi(x)$ satisfies the linear integral equation
$$ \begin{equation} \int_{D} \frac{\xi(x)\, dx}{|x- y||x-z|} = f_0(y,z),\qquad (y,z)\in Y\times Z, \end{equation} \tag{3.10} $$
which is referred to as the Lavrent’ev equation in the present paper. The right-hand side of equation (3.10) has the form
$$ \begin{equation} f_0(y,z)=\frac{8\pi^2}{\widetilde{g}(0)} \biggl( \int_0^{\infty} t^2 u(z,t;y)\, dt- H_{pp}(z,y;0) \biggr). \end{equation} \tag{3.11} $$
Note that $\widetilde{g}(0) \neq 0$ by the first condition in (3.2). For the second term in brackets on the right of (3.11), using (3.8) we have
$$ \begin{equation} H_{pp}(z,y;0)=-\frac{\widetilde{g}(0)|y-z|}{4\pi c_0^2}+\frac{\widetilde{g}'(0)}{2\pi c_0}- \frac{\widetilde{g}''(0)}{4\pi|y-z|}. \end{equation} \tag{3.12} $$

The magnitude of the wave field $u(z,t;y)$ averaged in time with weight $t^2$ for $(y,z)\in Y\times Z$ serves as input data for evaluation of $\xi$.

The inverse sounding problem of an inhomogeneity by point time-harmonic sources with frequencies $\omega\in(0,\omega_0]$ can also be reduced to equation (3.10); for more details, see [19], Ch. VII, § 1, [20], [21]. In the case of injective operator of equation (3.10), the solution of the original nonlinear inverse problem is unique. This property is satisfied, for example, if $Y$, $Z$ are domains on a plane disjoint from $D$ or bounded surfaces surrounding $D$. It is worth pointing out that, in all above cases, the dimension of the spatial data support $Y\times Z$ in (3.10) is 4, whereas the sought-for function $\xi$ depends only on 3 variables. Thus, the inverse problems in these papers are overdetermined. The above construction also includes the case $y=z$, which corresponds to the problem of reconstruction of an inhomogeneity from inverse scattering data. In this case, uniqueness of the solution can be verified in the non-overdetermined case, where $Y=Z$ is a domain in $\mathbb{R}^3$ disjoint from $\overline{D}$ [21].

Let us clarify how the injectivity of the integral operator (3.10) is related to the approximation problems considered in § § 1, 2. One has to verify that the homogeneous equation

$$ \begin{equation} \int_{D} \frac{\xi(x)\, dx}{|x- y||x - z|} = 0,\qquad (y,z)\in Y\times Z, \end{equation} \tag{3.13} $$
which corresponds to (3.10), has only the trivial solution. In this case, both equation (3.10) and the original inverse problem $\{ Y, Z \}$ have a unique solution. Equation (3.13) is symmetric with respect to $y,z$, and hence so is the inverse problem $\{ Z, Y \}$.

It is easily seen that if families of harmonic functions

$$ \begin{equation*} \mathcal{H}_1 =\biggl\{ \frac{1}{|x-y|}\colon y \in Y \biggr\}, \qquad \mathcal{H}_2 =\biggl\{ \frac{1}{|x-z|}\colon z \in Z \biggr\} \end{equation*} \notag $$
are such that $\mathcal{H}_1 \cdot \mathcal{H}_2$ is complete in $L_2(D)$, then equation (3.10) has only the zero solution.

Let us indicate one general method for choosing such sets $Y$, $Z$.

We say that $\mathcal{M} \subset \mathbb{R}^3 \setminus \overline{D}$ is a uniqueness set (for harmonic functions in $\mathbb{R}^3 \setminus \overline{D}$) if, for any regular harmonic (in $\mathbb{R}^3 \setminus \overline{D}$) function $u$ such that

$$ \begin{equation*} \lim_{|x|\to \infty} u(x)=0, \end{equation*} \notag $$
the equality $u|_{\mathcal{M}}=0$ implies that $u \equiv 0$ in $\mathbb{R}^3 \setminus \overline{D}$. As an example of a uniqueness set, we mention a closed surface disjoint from $\overline{D}$ or a domain on the plane disjoint from $\overline{D}$. Any domain on an analytic surface (for example, a sphere) which has no common points with $\overline{D}$ is also a uniqueness set.

The next proposition justifies the application of Theorem 1 for a proof of unique solvability of the homogeneous equation (3.13).

Proposition 5 (see [22]). Let $\mathcal{M}$ be a uniqueness set. Then the family of linear combinations of the functions $\{|x-w|^{-1}\colon w \in \mathcal{M} \}$ is $L_2(D)$-dense in $\mathcal{H}(D)$.

From Proposition 5 it follows that if $Z$ is a uniqueness set in the above sense, then equality (3.13) implies

$$ \begin{equation} \int_{D} \frac{\xi(x) u(x)\, dx}{|x-y|}=0\quad \forall \, y\in Y, \quad u\in \mathcal{H}(D). \end{equation} \tag{3.14} $$
If $Y$ is also a uniqueness set, then by applying Proposition 5 to (3.14) we get
$$ \begin{equation*} \int_{D} \xi(x) u(x) v(x)\, dx=0\quad \forall \, u,v\in \mathcal{H}(D). \end{equation*} \notag $$
Now an appeal to Proposition 1 shows that $\xi=0$ almost everywhere in $D$. The above overdetermined statements of the problem $\{ Y,Z \}$ correspond to two-dimensional uniqueness sets taken as $Y$ and $Z$. At the same time, Proposition 4 and Theorem 1 show that (3.14) implies the required equality $\xi=0$ almost everywhere in $D$ also in cases where $Y$ is not a uniqueness set. Indeed, combining (3.14) and (2.1) and using Theorem 1, we get the following result.

Theorem 2. Let $Z$ be a uniqueness set, $L$ be a straight line in $\mathbb{R}^3$, $Y$ be an open set lying in the unbounded component of $L\setminus \overline{D}$. Then, for any right-hand side $f_0$, equation (3.10) has at most one solution. If, in addition $Y \cap Z =\varnothing$, then the inverse problems $\{ Y, Z \}$ and $\{ Z, Y \}$ are also uniquely solvable.

If as $Z$ we take a two-dimensional uniqueness set (for example, a domain on the plane, or a closed surface disjoint from $\overline{D}$), then, under the conditions of Theorem 2, the dimension of the spatial data support $Y\times Z$ in problems $\{ Y, Z \}$ and $\{ Z, Y \}$ is 3, which equals the dimension of support $D$ of the sought-for function. In this sense, the inverse problems under consideration turn out to be spatially non-overdetermined. Note that here a complete data set $\{ u(z,t;y)\colon y\in Y$, $z\in Z$, $t>0 \}$ required for evaluation of the function $f_0$ in (3.10) involves the variable $t$, and hence its dimension is one plus $\operatorname{dim} D$. In [23], a related result was established on uniqueness for an equation obtained by transformation of the Lavrent’ev equation to a differential form with subsequent approximation via the Galerkin method.

§ 4. The inverse sounding problem in the phaseless setting

In this section, we will consider the inverse problem for the Helmholtz equation

$$ \begin{equation} \Delta u(x)+ k^2 n(x)u(x)=-\delta(x-y), \qquad x\in \mathbb{R}^3. \end{equation} \tag{4.1} $$
For equation (4.1), the radiation condition at infinity is specified,
$$ \begin{equation} \lim_{r\to \infty} r \biggl( \frac{\partial u}{\partial r}- iku \biggr)=0, \qquad r=| x|. \end{equation} \tag{4.2} $$
Here, $k>0$ is the wave number. It is assumed that the refraction coefficient $n(x) \equiv 1$, for $x \in \mathbb{R}^3\setminus D$, $n \in L_{\infty}(D)$, where $D$ is a given bounded domain in $\mathbb{R}^3$. System (4.1), (4.2) describes a scalar time-harmonic wave field $U(x,t)=\operatorname{Re}(e^{-ikt} u(x))$ in an inhomogeneous medium generated by a point time-harmonic source and described by the complex amplitude $u(x)$.

The classical statement of the inverse scattering problem for (4.1), (4.2) calls for recovery of the function $n(x)$, $x\in D$, from known values of the solution $u=u(x;y,k)$ of problem (4.1), (4.2) for $x=z\in Z$, $y \in Y$, $k\in [k_1,k_2]$, with appropriate sets $Y$, $Z$ and for chosen values $0<k_1\leqslant k_2$. As above, $Z$ is the set of detectors of the scattered wave field generated by a source at $y\in Y$, $Y$ is the set of point sounding sources, where we assume that $(Y \cup Z) \cap \overline{D}=\varnothing$. Inverse scattering problems in this and related settings with various choices of sets of sources and detectors have been extensively studied (see, for example, the books [2], [20], [24] and the references given there). The scattering amplitude in the far-zone, that is, the principal term in the asymptotics of $u(x;y,k)$ as $|x|\to \infty$, frequently plays the role of input data in inverse sounding problems. Here, an incoming plane wave may also serve as a wave field source. At the same time, experimental measurement of a complex function $u(x;y,k)$, for which both modulus (amplitude) $|u(x;y,k)|$ and the phase $\operatorname{arg} u(x;y,k)$ of the wave field should be obtained, is a very restrictive requirement in many applied problems. In many cases, experimental observations provide only the amplitude of a signal scattered by the inhomogeneity, that is, $|u(x;y,k)|$. In this regard, following [25]–[29], we will consider the problem of recovery of the function $n(x)$, $x\in D$, from a known amplitude $|u(z;y,k)|$ of the wave field, where $z\in Z$, $y \in Y$, $k\in [k_1,k_2]$, $0<k_1<k_2$. We will assume that $Y \cap Z=\varnothing$. We are interested in unique recovery of this function from the input data. For brevity, this inverse problem will be denoted by $[Y,Z]$.

We will next show that analysis of unique solvability of $[Y,Z]$ can also be reduced to the study of injectivity of the integral operator from the Lavrent’ev equation (3.10).

According to [24], § 8.2, problem (4.1), (4.2) is equivalent to the integral equation

$$ \begin{equation} u(x)=k^2 \int_{D} \Phi(x,x';k) (n(x')-1) u(x')\, dx'+\Phi(x,y;k), \qquad x\in D, \end{equation} \tag{4.3} $$
where
$$ \begin{equation} \Phi(x,x_0;k)=\frac{e^{i k|x-x_0|}}{4\pi|x-x_0|}. \end{equation} \tag{4.4} $$

Under the condition $n\in L_{\infty}(D)$, the integral operator on the right of (4.3) is completely continuous in $L_2(D)$ for any $k \in \mathbb{C}$ (see [30], § 8.5); moreover, the right-hand side of (4.3) depends analytically on the parameter $k$. According to Theorems 8.4, 8.7 in [24], problems (4.3) and (4.1), (4.2) are uniquely solvable for all real $k\geqslant 0$ and for complex $|k|\leqslant \delta$ with sufficiently small $\delta>0$. By theorem on holomorphic families of Fredholm operators (see [24], § 8.5, [17], Ch. VI, § 4), for fixed $x\neq y$, the function $u(x;y,k)$ extends holomorphically with respect to $k$ to a neighbourhood of the positive real axis. Retaining the same notation for the extended function of $k=\kappa_1+i \kappa_2$, we note that the real and imaginary parts ($\operatorname{Re} u(x;y,k)$, $\operatorname{Im} u(x;y,k)$) are regular harmonic functions of the variables $(\kappa_1,\kappa_2)$. Hence, for $x\neq y$, the function

$$ \begin{equation*} |u(x;y,k)|^2= (\operatorname{Re} u(x;y,k))^2 + (\operatorname{Im} u(x;y,k))^2 \end{equation*} \notag $$
is real analytic with respect to $k>-\delta$.

We set

$$ \begin{equation*} \xi(x)=n(x)-1, \qquad x\in \mathbb{R}^3. \end{equation*} \notag $$
By the above assumptions, $\xi(x) \equiv 0$ for $x\in \mathbb{R}^3 \setminus D$. We write equation (4.1) as
$$ \begin{equation} \Delta u(x) +k^2 u(x)=-k^2 \xi(x)u(x)-\delta(x-y). \end{equation} \tag{4.5} $$
The solution $u(x)=u(x;y,k)$ of problem (4.2), (4.5) satisfies the integral equation
$$ \begin{equation} u(x; y,k)=k^2 \int_{D} \Phi (x,x';k) \xi(x') u(x'; y,k)\, d x'+\Phi(x,y;k), \qquad x\in \mathbb{R}^3 \setminus\{ y \}. \end{equation} \tag{4.6} $$
From (4.6), for the above $x$, we have
$$ \begin{equation*} \begin{aligned} \, &u(x; y,k)=k^2 \int_{D} \Phi(x,x';k) \xi(x') \\ &\qquad\times\biggl(k^2 \int_{D} \Phi(x',x'';k) \xi (x'') u(x''; y,k)\, d x''+\Phi(x',y;k) \biggr) \, dx'+\Phi(x,y;k) \\ &=k^2 \int_{D} \Phi(x,x';k) \Phi(x',y;k) \xi(x')\, d x' \\ &\qquad+k^4 \int_{D} \Phi(x,x';k) \xi(x') \biggl( \int_{D} \Phi(x',x'';k) \xi(x'') u(x''; y,k) \, d x'' \biggr)\, dx'+ \Phi(x,y;k). \end{aligned} \end{equation*} \notag $$
Using the identity
$$ \begin{equation*} |z_1+z_2|^2=|z_1|^2+2 \operatorname{Re}(\overline{z}_1 z_2)+|z_2|^2, \qquad z_1,z_2 \in \mathbb{C}, \end{equation*} \notag $$
we find that, for all $y\in Y$, $z\in Z$ and $k\geqslant 0$,
$$ \begin{equation} \begin{aligned} \, &|u(z; y,k)|^2=|\Phi(z,y;k)|^2 \nonumber \\ &\qquad+2 k^2 \operatorname{Re}\biggl(\overline{\Phi(z,y;k)} \int_{D} \Phi(z,x';k) \Phi(x',y;k) \xi(x')\, d x'\biggr)+O(k^4). \end{aligned} \end{equation} \tag{4.7} $$
From (4.7) it follows that, for $y\in Y$, $z\in Z$,
$$ \begin{equation} \operatorname{Re}\biggl( \overline{\Phi(z,y;0)} \int_{D} \Phi(z,x';0) \Phi(x',y;0) \xi(x')\, d x'\biggr) =f_1(y,z), \end{equation} \tag{4.8} $$
where
$$ \begin{equation} f_1(y,z)=\lim_{k\to 0} (2 k^2)^{-1} \bigl(|u(z; y,k)|^2-|\Phi(z,y;k)|^2\bigr). \end{equation} \tag{4.9} $$

Note that the function $\Phi(z,y;0)=(4\pi|z-y|)^{-1}$ is real-valued. Hence from (4.4), (4.8) it follows that the real-valued function $\xi$ satisfies the Lavrent’ev equation

$$ \begin{equation} \begin{gathered} \, \int_{D} \frac{\xi(x)\, dx}{|x-y||x-z|} =f_2(y,z), \qquad (y,z) \in Y \times Z; \\ f_2(y,z) =64 \pi^3|y-z|f_1(y,z). \end{gathered} \end{equation} \tag{4.10} $$

The function $|u(z;y,k)|^2$ is real analytic in $k>-\delta$, and hence $|u(z;y,k)|^2$ for all $k>-\delta$, $(y,z)\in Y \times Z$ is uniquely determined from the values of $|u(z;y,k)|^2$ for $k \in [k_1,k_2]$ and $(y,z)\in Y \times Z$. So, the function in the right-hand side of equation (4.10) is uniquely defined by the input data of the inverse problem $[Y,Z]$. Hence the solution $[Y,Z]$ is unique if and only if the corresponding homogeneous equation (3.13) has only the trivial solution. The next result now follows from Theorem 1.

Theorem 3. Let $Z$ be a uniqueness set, $L$ be a straight line in $\mathbb{R}^3$, $Y$ be an open set lying in the unbounded component of the complement $L\setminus \overline{D}$, $Y\cap Z=\varnothing$. Then the inverse problems $[ Y, Z ]$ and $[ Z, Y ]$ have a unique solution.

Choosing as $Z$ a two-dimensional uniqueness set and employing Theorem 3, we obtain that the nonlinear coefficient inverse problems $[Y, Z]$ and $[Z, Y]$ are uniquely solvable in the spatially non-overdetermined setting, where the dimension of the spatial data support $Y\times Z$ coincides with that of the support $D$ of the original function. It is worth pointing out that in the preceding papers [26]–[29] uniqueness theorems were put forward for problems with spatial data supports of the form $\{ (y,z)\colon y\in \Sigma,\, z\in O_{\varepsilon}(y) \}$, $\{ (y,z)\colon y,\, z\in \Sigma \}$ (of dimension at least 4). Here $\Sigma$ is a closed surface surrounding $\overline{D}$. In [26], a uniqueness theorem for non-overdetermined phaseless inverse problem was established if (4.1) involves a unique spatially distributed sounding source. This result was obtained under quite stringent assumptions about the form and smoothness of the sought-for function $n(x)$.

In § 5, some corollaries to Theorems 2, 3 will be discussed.

§ 5. Axial symmetry of inhomogeneities

Note that the expressions in the denominator of the integrand in equations (3.10), (4.10) are invariant under orthogonal transformations of the variables. With a suitable arrangement of sets of sources and detectors, this property in combination with Theorems 2, 3 can be used for establishing symmetry properties of the inhomogeneity given analogous symmetry in the input data of the inverse problems $\{ Y, Z\}$, $[Y, Z]$. Let us illustrate this for the axial symmetry property.

We first consider the problem $\{ Y, Z\}$ equivalent to the integral equation (3.10). For definiteness, we will assume that

$$ \begin{equation} \begin{gathered} \, D = O_1(0), \qquad Y=\{ x\in \mathbb{R}^3\colon x=(0,0,x_3),\, x_3 \in (\chi_1, \chi_2) \}, \quad 1<\chi_1< \chi_2, \nonumber \\ Z=\{ x\in \mathbb{R}^3\colon (x_1, x_2, \kappa),\, x_1^2+x_2^2 < r^2 \}, \qquad r>0,\quad \kappa<-1. \end{gathered} \end{equation} \tag{5.1} $$
Here $Z$ is an open subset of the plane $\{ x\in \mathbb{R}^3\colon x_3=\kappa \}$ that does not intersect $\overline{D}$, and is a uniqueness set.

Let $R_{\varphi}$ be the matrix of rotation of $\mathbb{R}^3$ by an angle $\varphi\in [-\pi,\pi)$ about the $x_3$-axis. Suppose that

$$ \begin{equation} u(R_{\varphi}z,t;y)=u(z,t;y) \quad \forall \, (y,z)\in Y\times Z,\quad t>0,\quad \varphi\in [-\pi,\pi). \end{equation} \tag{5.2} $$
We have $|y-R_{\varphi}z|=|y-z|$ for all $(y,z)\in Y\times Z$, and hence, from (3.11), (3.12) it follows that under (5.2)
$$ \begin{equation} f_0 (y,R_{\varphi}z)= f_0 (y,z) \quad \forall \, (y,z)\in Y\times Z,\quad \varphi\in [-\pi,\pi). \end{equation} \tag{5.3} $$
Next, $R_{-\varphi}(D)=D$ and $\operatorname{det}R_{\varphi}=1$, and hence, changing the variable $x=R_{\varphi} x'$ in the left-hand side of (3.10) and using (5.3), we obtain
$$ \begin{equation} \int_D \frac{\xi(R_{\varphi} x')\, dx'}{|R_{\varphi} x'-y||R_{\varphi} x'-z|}=\int_D \frac{\xi(R_{\varphi} x')\, dx'}{|x'-y||x'-R_{-\varphi} z|}=f_0(y,z)=f_0(y,R_{-\varphi}z) \end{equation} \tag{5.4} $$
for all $(y,z)\in Y \times Z$. Here, we used the equalities
$$ \begin{equation*} \begin{gathered} \, |R_{\varphi}x-y|=|R_{\varphi} (x-y)|=|x-y|, \quad |R_{\varphi}x-z|= |R_{\varphi}(x-R_{-\varphi}z)|=|x-R_{-\varphi}z|, \\ x\in D, \qquad y\in Y, \qquad z\in Z, \qquad \varphi\in [-\pi,\pi), \end{gathered} \end{equation*} \notag $$
which hold because the matrix $R_{\varphi}$ is orthogonal.

Since, for a fixed $\varphi$, the range of all points $\widetilde{z}=R_{-\varphi}z$ is the whole set $Z$ as $z$ varies over $Z$, from (5.4) we get

$$ \begin{equation} \int_D \frac{\xi(R_{\varphi} x')\, dx'}{|x'-y||x'-\widetilde{z}|} =f_0(y,\widetilde{z}), \qquad (y,\widetilde{z}) \in Y \times Z. \end{equation} \tag{5.5} $$
A combination of (5.5) and (3.10) shows that equation (3.10), along with $\xi=\xi(x)$, has the solution $\xi=\xi(R_{\varphi}x)$. From (3.3) it follows that the function $c=c(x)$ is also a solution (in addition to $c(R_{\varphi}x)$) of problem $\{ Y, Z\}$. The following result is now a consequence of Theorem 2 since the angle $\varphi\in [-\pi,\pi)$ is arbitrary.

Theorem 4. Let in problem $\{ Y,Z \}$ or $\{ Z,Y \}$, the sets $Y$, $Z$ are chosen according to (5.1) and let condition (5.2) be met. Then the solution $c(x)$ of this problem is axially symmetric, that is,

$$ \begin{equation} c(R_{\varphi}x)=c(x), \qquad \varphi\in [-\pi,\pi). \end{equation} \tag{5.6} $$

Remark 1. A similar argument shows that if (5.2) holds for the rotation angles $\varphi=\varphi_k=\pi k/N$, $-N \leqslant k \leqslant N-1$, $N\geqslant 2$, then $c(R_{\varphi_k} x)=c(x)$, $-N \leqslant k \leqslant N-1$.

Remark 2. It is easily checked that (5.2) holds under condition (5.6). This result is immediate from equations (3.1) with $x=R_{\varphi} x'$ and since problem (3.1) is uniquely solvable. Theorem 4 can be looked upon as a converse result to that given in this remark.

As a well-know analogue of Theorem 4, we mention S. N. Karp’s theorem [24], § 5.1, § 7.1, and its generalizations [31]. However, these results are concerned with spherical symmetry of the scattering amplitude $A(\eta,\eta')$ qua a function of two vectors $\eta$, $\eta'$ from the unit sphere $S$. Hence, testing the corresponding symmetry condition $A(U\eta, U\eta')=A(\eta,\eta')$ for all $\eta,\eta' \in S$ and all orthogonal $(3\times 3)$-matrices $U$ with $\operatorname{det}U=1$ calls for a solution of the overdetermined scattering problem with data in the four-dimensional manifold $S \times S$. The hypotheses of Theorem 4 can be tested in solving the inverse problems $\{ Y,Z \}$, $\{ Z, Y \}$ with three-dimensional data support $Y \times Z$. At the same time, in Theorem 4 and Remark 1, one deals with narrower transformation groups.

Let us now consider the inverse problem $[Y,Z]$ in which $Y$, $Z$, $D$ are chosen according to (5.1). It is easily checked that, under the condition

$$ \begin{equation*} n(R_{\varphi}x)=n(x), \qquad \varphi\in [-\pi,\pi), \end{equation*} \notag $$
the solution of problem (4.1), (4.2) features a similar symmetry, that is, for all $\varphi\in [-\pi,\pi)$.
$$ \begin{equation*} u(R_{\varphi}x,y;k)=u(x,y;k), \qquad x\neq y. \end{equation*} \notag $$
For a proof, it suffices to put $x=R_{\varphi}\widetilde{x}$ in (4.3) and change $x'=R_{\varphi} x''$ in the integral.

Conversely, let the equality

$$ \begin{equation} |u(R_{\varphi}z,y;k)|=|u(z,y;k)|\quad \forall \, (y,z)\in Y\times Z,\quad \varphi\in [-\pi,\pi), \end{equation} \tag{5.7} $$
hold. Since the functions of $k$ in both sides of (5.7) are analytic with respect to $k>-\delta$, this equality can be extended by analyticity to the entire range $k\geqslant 0$. Hence, in view of (4.9), (4.10), we have
$$ \begin{equation*} f_2(R_{\varphi}z,y)=f_2(z,y), \qquad (y,z)\in Y\times Z. \end{equation*} \notag $$

Arguing as in the proof of Theorem 4, we get the following result.

Theorem 5. Let in problem $[ Y,Z ]$ or $[ Z,Y ]$, the sets $Y$, $Z$ are chosen according to (5.1) and let condition (5.7) be met. Then the solution $n(x)$ of the corresponding inverse problem satisfies

$$ \begin{equation*} n(R_{\varphi}x)=n(x), \qquad \varphi \in [-\pi,\pi). \end{equation*} \notag $$

A result similar to Remark 1 holds under the conditions of Theorem 5.

Remark 3. Transforming the integral in (2.1), using the inversion as a change of variables, and taking into account that harmonicity is invariant under the Kelvin transform, it can be shown that Theorem 1 and the subsequent results are also true if $Y$ lies in bounded component of the set $L\setminus \overline{D}$.

The author is grateful to the referee for calling his attention to the last remark.


Bibliography

1. A. P. Calderón, “On an inverse boundary value problem”, Seminar on numerical analysis and its applications to continuum physics (Rio de Janeiro 1980), Soc. Brasil. Mat., Rio de Janeiro, 1980, 65–73  mathscinet  zmath
2. V. Isakov, Inverse problems for partial differential equations, Appl. Math. Sci., 127, 2nd ed., Springer, New York, 2006  crossref  mathscinet  zmath
3. C. Kenig and M. Salo, “Recent progress in the Calderón problem with partial data”, Inverse problems and applications, Contemp. Math., 615, Amer. Math. Soc., Providence, RI, 2014, 193–222  crossref  mathscinet  zmath
4. M. Yu. Kokurin, “On the completeness of products of harmonic functions and the uniqueness of the solution of the inverse acoustic sounding problem”, Mat. Zametki, 104:5 (2018), 708–716  mathnet  crossref  mathscinet  zmath; English transl. Math. Notes, 104:5 (2018), 689–695  crossref
5. M. Yu. Kokurin, “Completeness of the asymmetric products of solutions of a second-order elliptic equation and the uniqueness of the solution of an inverse problem for the wave equation”, Differ. Uravn., 57:2 (2021), 255–264  mathscinet  zmath; English transl. Differ. Equ., 57:2 (2021), 241–250  crossref
6. V. P. Mikhailov, Partial differential equations, Nauka, Moscow, 1976  mathscinet  zmath; English transl. Moscow, Mir, 1978  mathscinet  zmath
7. A. F. Nikiforov and V. B. Uvarov, Special functions of mathematical physics. A unified introduction with applications, Nauka, Moscow, 1978  mathscinet; English transl. Birkhäuser, Basel, 1988  crossref  mathscinet  zmath
8. D. A. Varshalovich, A. N. Moskalev, and V. K. Khersonskiĭ, Quantum theory of angular momentum. Irreducible tensors, spherical harmonics, vector coupling coefficients, $3nj$ symbols, Nauka, Leningrad, 1975; English transl. World Sci. Publ., Teaneck, NJ, 1988  crossref  mathscinet  zmath
9. E. M. Stein and G. Weiss, Introduction to Fourier analysis on Euclidean spaces, Princeton Math. Ser., 32, Princeton Univ. Press, Princeton, NJ, 1971  mathscinet  zmath; Russian transl. Mir, Moscow, 1974  zmath
10. S. M. Nikol'skiĭ, Approximation of functions of several variables and imbedding theorems, 2nd ed., Nauka, Moscow, 1977  mathscinet  zmath; English transl. of 1st ed. Grundlehren Math. Wiss., 205, Springer-Verlag, New York–Heidelberg, 1975  crossref  mathscinet  zmath
11. I. P. Natanson, Theory of functions of a real variable, 3rd ed., Nauka, Moscow, 1974  mathscinet; English transl. of 1st ed. v. 1, 2, Frederick Ungar Publishing Co., New York, 1955, 1961  mathscinet  mathscinet  zmath
12. L. D. Landau and E. M. Lifshitz, Course of theoretical physics, v. 3, Quantum mechanics: non-relativistic theory, 6th ed., Fizmatlit, Moscow, 2008; English transl. of 1st ed. Addison-Wesley Series in Advanced Physics, Pergamon Press, London–Paris; Addison-Wesley, Reading, MA, 1958  mathscinet  zmath
13. N. I. Ahiezer, Lectures in the theory of approximation, 2nd ed., Nauka, Moscow, 1965 (Russian)  mathscinet  zmath; German transl. N. I. Achieser, Vorlesungen über Approximationstheorie, Math. Lehrbücher und Monogr., II, Akademie, Berlin, 1967  mathscinet  zmath
14. A. B. Bakushinskii, A. I. Kozlov, and M. Yu. Kokurin, “On some inverse problem for a three-dimensional wave equation”, Zh. Vychisl. Mat. Mat. Fiz., 43:8 (2003), 1201–1209  mathnet  mathscinet  zmath; English transl. Comput. Math. Math. Phys., 43:8 (2003), 1149–1158
15. M. M. Lavrent'ev, “On an inverse problem for the wave equation”, Dokl. Akad. Nauk SSSR, 157:3 (1964), 520–521  mathnet  mathscinet  zmath; English transl. Soviet Math. Dokl., 5 (1964), 970–972
16. M. M. Lavrent'ev, “A class of inverse problems for differential equations”, Dokl. Akad. Nauk SSSR, 160:1 (1965), 32–35  mathnet  mathscinet  zmath; English transl. Soviet Math. Dokl., 6 (1965), 29–32
17. B. Vainberg, Asymptotic methods in equations of mathematical physics, Moscow Univ. Press, Moscow, 1982 (Russian)  mathscinet  zmath; English transl. Gordon and Breach, New York, 1989  mathscinet  zmath
18. V. G. Romanov, “On smoothness of a fundamental solution to a second order hyperbolic equation”, Sibirsk. Mat. Zh., 50:4 (2009), 883–889  mathnet  mathscinet  zmath; English transl. Siberian Math. J., 50:4 (2009), 700–705  crossref
19. M. M. Lavrent'ev, V. G. Romanov, and S. P. Shishatskiĭ, Nauka, Moscow, 1980  mathscinet  zmath; English transl. Ill-posed problems of mathematical physics and analysis, Transl. Math. Monogr., 64, Amer. Math. Soc., Providence, RI, 1986  mathscinet  zmath
20. A. G. Ramm, Multidimensional inverse scattering problems, Pitman Monogr. Surveys Pure Appl. Math., 51, Longman Scientific & Technical, Harlow; John Wiley & Sons, New York, 1992  mathscinet  zmath; Russian transl. Mir, Moscow, 1994  zmath
21. A. L. Bukhgeĭm, G. V. Dyatlov, V. B. Kardakov, and E. V. Tantserev, “Uniqueness in one inverse problem for the elasticity system”, Sibirsk. Mat. Zh., 45:4 (2004), 747–757  mathnet  mathscinet  zmath; English transl. Siberian Math. J., 45:4 (2004), 618–627  crossref
22. M. Yu. Kokurin, “On a multidimensional integral equation with data supported by low-dimensional analytic manifolds”, J. Inverse Ill-Posed Probl., 21:1 (2013), 125–140  crossref  mathscinet  zmath
23. M. V. Klibanov, Jingzhi Li, and Wenlong Zhang, “Linear Lavrent'ev integral equation for the numerical solution of a nonlinear coefficient inverse problem”, SIAM J. Appl. Math., 81:5 (2021), 1954–1978  crossref  mathscinet  zmath
24. D. Colton and R. Kress, Inverse acoustic and electromagnetic scattering theory, Appl. Math. Sci., 93, 2nd ed., Springer-Verlag, Berlin, 1998  crossref  mathscinet  zmath
25. M. V. Klibanov, “Phaseless inverse scattering problems in three dimensions”, SIAM J. Appl. Math., 74:2 (2014), 392–410  crossref  mathscinet  zmath
26. M. V. Klibanov, “Uniqueness of two phaseless non-overdetermined inverse acoustic problems in 3-d”, Appl. Anal., 93:6 (2014), 1135–1149  crossref  mathscinet  zmath
27. M. V. Klibanov and V. G. Romanov, “Reconstruction procedures for two inverse scattering problems without the phase information”, SIAM J. Appl. Math., 76:1 (2016), 178–196  crossref  mathscinet  zmath
28. V. G. Romanov, “The problem of recovering the permittivity coefficient from the modulus of the scattered electromagnetic field”, Sibirsk. Mat. Zh., 58:4 (2017), 916–924  mathnet  crossref  mathscinet  zmath; English transl. Siberian Math. J., 58:4 (2017), 711–717  crossref
29. V. G. Romanov, “Phaseless inverse problems that use wave interference”, Sibirsk. Mat. Zh., 59:3 (2018), 626–638  mathnet  crossref  mathscinet  zmath; English transl. Siberian Math. J., 59:3 (2018), 494–504  crossref
30. M. A. Krasnosel'skii, P. P. Zabreyko, E. I. Pustylnik, and P. E. Sobolevski, Integral operators in spaces of summable functions, Nauka, Moscow, 1966  mathscinet  zmath; English transl. Monographs Textbooks Mech. Solids Fluids: Mech. Anal., Noordhoff International Publishing, Leiden, 1976  mathscinet  zmath
31. A. G. Ramm, “Symmetry properties of scattering amplitudes and applications to inverse problems”, J. Math. Anal. Appl., 156:2 (1991), 333–340  crossref  mathscinet  zmath

Citation: M. Yu. Kokurin, “Completeness of asymmetric products of harmonic functions and uniqueness of the solution to the Lavrent'ev equation in inverse wave sounding problems”, Izv. RAN. Ser. Mat., 86:6 (2022), 101–122; Izv. Math., 86:6 (2022), 1123–1142
Citation in format AMSBIB
\Bibitem{Kok22}
\by M.~Yu.~Kokurin
\paper Completeness of asymmetric products of harmonic functions and uniqueness of the solution
to the Lavrent'ev equation in inverse wave sounding problems
\jour Izv. RAN. Ser. Mat.
\yr 2022
\vol 86
\issue 6
\pages 101--122
\mathnet{http://mi.mathnet.ru/im9253}
\crossref{https://doi.org/10.4213/im9253}
\mathscinet{http://mathscinet.ams.org/mathscinet-getitem?mr=4582548}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2022IzMat..86.1123K}
\transl
\jour Izv. Math.
\yr 2022
\vol 86
\issue 6
\pages 1123--1142
\crossref{https://doi.org/10.4213/im9253e}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=000992259900004}
\scopus{https://www.scopus.com/record/display.url?origin=inward&eid=2-s2.0-85165325517}
Linking options:
  • https://www.mathnet.ru/eng/im9253
  • https://doi.org/10.4213/im9253e
  • https://www.mathnet.ru/eng/im/v86/i6/p101
  • This publication is cited in the following 2 articles:
    Citing articles in Google Scholar: Russian citations, English citations
    Related articles in Google Scholar: Russian articles, English articles
    Известия Российской академии наук. Серия математическая Izvestiya: Mathematics
    Statistics & downloads:
    Abstract page:236
    Russian version PDF:23
    English version PDF:51
    Russian version HTML:136
    English version HTML:54
    References:41
    First page:12
     
      Contact us:
     Terms of Use  Registration to the website  Logotypes © Steklov Mathematical Institute RAS, 2024