Loading [MathJax]/extensions/TeX/mathchoice.js
Special Issues

On recent progress of single-realization recoveries of random Schrödinger systems

  • We consider the recovery of some statistical quantities by using the near-field or far-field data in quantum scattering generated under a single realization of the randomness. We survey the recent main progress in the literature and point out the similarity among the existing results. The methodologies in the reformulation of the forward problems are also investigated. We consider two separate cases of using the near-field and far-field data, and discuss the key ideas of obtaining some crucial asymptotic estimates. We pay special attention on the use of the theory of pseudodifferential operators and microlocal analysis needed in the proofs.

    Citation: Shiqi Ma. On recent progress of single-realization recoveries of random Schrödinger systems[J]. Electronic Research Archive, 2021, 29(3): 2391-2415. doi: 10.3934/era.2020121

    Related Papers:

    [1] Shiqi Ma . On recent progress of single-realization recoveries of random Schrödinger systems. Electronic Research Archive, 2021, 29(3): 2391-2415. doi: 10.3934/era.2020121
    [2] Xinlin Cao, Huaian Diao, Jinhong Li . Some recent progress on inverse scattering problems within general polyhedral geometry. Electronic Research Archive, 2021, 29(1): 1753-1782. doi: 10.3934/era.2020090
    [3] Xiaoping Fang, Youjun Deng, Wing-Yan Tsui, Zaiyun Zhang . On simultaneous recovery of sources/obstacles and surrounding mediums by boundary measurements. Electronic Research Archive, 2020, 28(3): 1239-1255. doi: 10.3934/era.2020068
    [4] Yan Chang, Yukun Guo . Simultaneous recovery of an obstacle and its excitation sources from near-field scattering data. Electronic Research Archive, 2022, 30(4): 1296-1321. doi: 10.3934/era.2022068
    [5] Deyue Zhang, Yukun Guo . Some recent developments in the unique determinations in phaseless inverse acoustic scattering theory. Electronic Research Archive, 2021, 29(2): 2149-2165. doi: 10.3934/era.2020110
    [6] Koung Hee Leem, Jun Liu, George Pelekanos . A regularized eigenmatrix method for unstructured sparse recovery. Electronic Research Archive, 2024, 32(7): 4365-4377. doi: 10.3934/era.2024196
    [7] Messoud Efendiev, Vitali Vougalter . Linear and nonlinear non-Fredholm operators and their applications. Electronic Research Archive, 2022, 30(2): 515-534. doi: 10.3934/era.2022027
    [8] Won-Ki Seo . Fredholm inversion around a singularity: Application to autoregressive time series in Banach space. Electronic Research Archive, 2023, 31(8): 4925-4950. doi: 10.3934/era.2023252
    [9] Yanjiao Li, Yue Zhang . Dynamic behavior on a multi-time scale eco-epidemic model with stochastic disturbances. Electronic Research Archive, 2025, 33(3): 1667-1692. doi: 10.3934/era.2025078
    [10] Zhaoyan Meng, Shuting Lyu, Mengqing Zhang, Xining Li, Qimin Zhang . Sufficient and necessary conditions of near-optimal controls for a stochastic listeriosis model with spatial diffusion. Electronic Research Archive, 2024, 32(5): 3059-3091. doi: 10.3934/era.2024140
  • We consider the recovery of some statistical quantities by using the near-field or far-field data in quantum scattering generated under a single realization of the randomness. We survey the recent main progress in the literature and point out the similarity among the existing results. The methodologies in the reformulation of the forward problems are also investigated. We consider two separate cases of using the near-field and far-field data, and discuss the key ideas of obtaining some crucial asymptotic estimates. We pay special attention on the use of the theory of pseudodifferential operators and microlocal analysis needed in the proofs.



    In this paper, we mainly focus on the random inverse problems associated with the following time-harmonic Schrödinger system

    (ΔE+potential)u(x)=source,xRn, (1.1)

    where E is the energy level, n is the dimension, and "source" and the "potential" in (1.1) shall be specified later. In some cases we may impose incident waves to the system in order to obtain more useful information, thus

    u(x)=αuin(x)+usc(x) (1.2)

    where α takes the value of either 0 or 1 corresponding to impose or suppress the incident wave, respectively. The corresponding data are thus called passive or active measurements. Moreover, we shall impose the Sommerfeld radiation condition [10]

    limrr(uscriEusc)=0,r:=|x|, (1.3)

    that characterizes the outgoing nature of the scattered field usc. The system (1.1)-(1.3) describes the quantum scattering [13,14] associated with a source and a potential at the energy level E. Later we follow the convention to use k:=E to signify the frequency at which the system is acting on.

    Under different assumptions of the potential and source, of the dimension, and of the incident wave, the regularity of the Schrödinger system (1.1)-(1.3) behaves differently and calls for different techniques for the recovery procedure. The randomness of the Schrödinger system (1.1)-(1.3) can present either in the potential, or in the source, or in both. In this paper we shall investigate all of these three cases, survey the results in the literature and give details of part of the proofs.

    There are rich literature on the inverse scattering problem using either passive or active measurements as data. For a fixed potential, the recovery of the deterministic unknown source of the system is called the inverse source problem. For the theoretical analysis and computational methods of the inverse source problems, readers may refer to [3,4,5,9,31,34] and references therein. The simultaneous recovery of the deterministic unknown source and potential are also studied in the literature. In [17,26], the authors considered the simultaneous recovery of an unknown source and its surrounding medium parameter. This type of inverse problems also arises in the deterministic magnetic anomaly detections using geomagnetic monitoring [11,12] with passive measurements. While [11,12,17,26] focus on deterministic setting with passive measurements, the works [2,6,7,18,19,20,27,33] pay attention to random settings. We are particularly interested in the case with a single realization of the random sample. The single-realization recovery has been studied in the literature. In this paper we mainly focus on [8,18,19,20,21,22,23,24,25].

    In [18,19], Lassas et. al. considered the inverse scattering problem for the two-dimensional random Schrödinger system (Δk2q(x,ω))u(x,k,ω)=δy, xR2 which is incited by point sources uin(x)=i4H(1)0(k|xy|); the H(1)0 is the Hankel function for the first kind, and the origin y of this source are located outside the support of the potential. The potential q(x,ω) is a micro-locally isotropic generalized Gaussian field (migr field) with compact support. The definition of the migr field can be found in Definition 1.1. They introduced the so-called rough strength μ(x), which is the informative part of the principal symbol μ(x)|ξ|m of the covariance operator. The m in μ(x)|ξ|m is the rough order of the random potential. The main result in their work states that under a single measurement of the random field inside a measurement domain, the rough strength can be recovered.

    In 2019, Caro et. al. [8] considered an inverse scattering problem for an n-dimensional (n2) random Schrödinger system (Δk2q(x,ω))u(x,k,ω)=0, xRn with incident wave being the plane wave, i.e. u is incited by the point sources uin(x)=eikdx; d is the incident direction. Again, the potential q is assumed to be a migr field with compact support. The main result is as follows: they used the backscattering far-field pattern and recovered the rough strength μ(x) almost surely, under a single realization of the randomness.

    In [20], Li, et. al. studied the case where the potential is zero and the source is migr field. In [24] Li, et. al. studied the same setting but with the energy level E replaced by (k2+iσk) where the σ is the attenuation parameter. The random source term considered is constructed as a migr field. The system has been changed to Helmholtz system in [24] but the underlying equation is uniform with the Schrödinger's equation. The authors studied the regularity of the random source and gave the well-posedness of the direct problem. Then they represented the solution as the convolution between the fundamental solution and the random source. By truncating the fundamental solution, they indicated that the rough strength can be recovered by utilizing the correspondingly truncated solution. Further, the authors used calculus of symbols to recover the rough strength.

    Then in [23], Li, et. al. further extended their study to Maxwell's equation. The recovery procedure in these three works share the same idea–the leading order term in the Bonn expansion gives the recovery of the desired statistics while the higher order terms converge to zero. The proof of these converges involve the utilization of Fourier integral operator. We shall give detailed explanations in Section 3.

    In [21], the authors consider the direct and inverse scatterings for (1.1)–(1.3) with a deterministic potential and a random source. The random source is a generalized Gaussian random field with local mean value function and local variance function, which are assumed to be bounded and compactly supported. The well-posedness of the direct scattering has been formulated in some weighted L2 space. Then the inverse scattering is studied and a recovery formula of the variance function is obtained, and the uniqueness recovery of the potential is given. The authors used both passive and active measurements to recover the unknowns. The passive measurements refer to the scattering data generated only by the unknown source (α is set to be 0 in (1.2)); active measurements refer to the scattering data generated by both the source and the incident wave (α is set to be 1 in (1.2)). To recover the variance function, only the passive measurements are needed, while the unique recovery of the potential needs active measurements.

    In [25], the authors extended the work [21] to the case where the source is a migr field. The direct scattering problem is formulated in a similar manner as in [21], while the technique used in the inverse scattering problem differs from that of [21]. In order to analyze the asymptotics of higher order terms in the Bonn expansion corresponding to the migr fields, stationary phase lemma and pseudodifferential operator are utilized.

    Then the authors extended the work [25] to the case where both the potential and the source are random (of migr type), and the extended result is presented in [22]. The results between [21] and [22,25] have two major differences. First, in [21] the random part of the source is assumed to be a Gaussian white noise, while in [22] the potential and the source are assumed to be migr fields. The migr field can fit larger range of randomness by tuning its rough order and rough strength. Second, in [22] both the source and potential are random, while in [25] the potential is assumed to be deterministic. These two facts make [22] much more challenging than that in [25]. The techniques used in the estimates of higher order terms in [22] are pseudodifferential operators and microlocal analysis and we shall give a detailed treatment in Section 4.

    Although the techniques used in [21,22,25] are different, the recovery formulae fall into the same pattern. The thesis [28] partially collected these three works[21,22,25] and readers may refer to the thesis for a more coherent discussion on this topic.

    In this paper we mainly pay attention to two types of random model, the Gaussian white noise and the migr field. The Gaussian white noise is well-known and readers may refer to [21,Section 2.1] for more details. Here we give a brief introduction to the migr field. We assume f to be a generalized Gaussian random distribution of the microlocally isotropic type (cf. Definition 1.1). It means that f(,ω) is a random distribution and the mapping

    ωΩ  f(,ω),φC,φS(Rn),

    is a Gaussian random variable whose probabilistic measure depends on the test function φ. Here and also in what follows, S(Rn) stands for the Schwartz space. Since both f(,ω),φ and f(,ω),ψ are random variables for φ, ψS(Rn), from a statistical point of view, the covariance between these two random variables,

    Eω(¯f(,ω)E(f(,ω)),φf(,ω)E(f(,ω)),ψ), (1.4)

    can be understood as the covariance of f. Here Eω means to take expectation on the random variable ω. Hence, formula (1.4) defines an operator Cf,

    Cf:φS(Rn)  Cf(φ)S(Rn),

    in a way that Cf(φ):ψS(Rn)  (Cf(φ))(ψ)C where

    (Cf(φ))(ψ):=Eω(¯f(,ω)E(f(,ω)),φf(,ω)E(f(,ω)),ψ).

    The operator Cf is called the covariance operator of f.

    Definition 1.1 (Migr field). A generalized Gaussian random distribution f on Rn is called microlocally isotropic with rough order m and rough strength μ(x) in a bounded domain D, if the following conditions hold:

    1. the expectation E(f) is in Cc(Rn) with suppE(f)D;

    2. f is supported in D a.s. (almost surely);

    3. the covariance operator Cf is a classical pseudodifferential operator of order m;

    4. Cf has a principal symbol of the form μ(x)|ξ|m with μCc(Rn;R), suppμD and μ(x)0 for all xRn.

    We call a microlocally isotropic Gaussian random distribution as an migr field.

    For the case where both the source and the potential are deterministic and are L functions with compact supports, the well-posedness of the direct problem of system (1.1)–(1.3) is known; see, e.g., [10,13,29]. Moreover, there holds the following asymptotic expansion of the outgoing radiating field usc as |x|+,

    usc(x)=eik|x||x|(n1)/2u(ˆx,k,d)+o(|x|(n1)/2),xRn.

    u(ˆx,k,d) is referred to as the far-field pattern, which encodes information of the potential and the source. ˆx:=x/|x| and d in u(ˆx,k,d) are unit vectors and they respectively stand for the observation direction and the impinging direction of the incident wave. When d=ˆx, u(ˆx,k,ˆx) is called the backscattering far-field pattern. We shall see very soon that both the near-field usc and the far-field u can be used to achieve the recovery.

    In (1.1), let us denote the source term as f and the potential term as q. In our study, both the source f and the potential q are assumed to be compactly supported. We shall treat [8,18,19,20,21,22,23,24,25] in more details. In [8,18,19], q is assumed to be a migr field while f is either zero or point a point source, i.e. δy(x). In [20,23,24], q is assumed to be zero and f is assumed to be a migr field. In [21], q is assumed to be unknown and deterministic and f is assumed to be a Gaussian white noise, while in [22,25], q is assumed to be deterministic or migr type and f is assumed to be a migr field.

    In [18,19] the authors considered the inverse scattering problem for the two-dimensional random Schrödinger system (Δk2q(x,ω))u(x,k,ω)=δy(x)(xR2) which is incited by point sources uin(x)=i4H(1)0(k|xy|); the H(1)0 is the Hankel function for the first kind, and the origin y of this source is located in U. The potential q(x,ω) is a migr field with compact support D and ¯U¯D=. The main result is presented as follows (cf. [19,Theorem 7.1]).

    Theorem 1.1. In [18,19], for x,yU the limit

    R(x,y)=limK+1K1K1k2+m|usc(x,y,k,ω)|2dk

    holds almost surely where

    R(x,x):=126+mπ2R2μq(z)|xz|2dz,xU.

    and the μq is the rough strength and m is the rough order of q.

    In [8], the authors considered (Δk2q(x,ω))u(x,k,ω)=0, xRn with incident plane wave uin(x)=eikds. The potential q is assumed to be a migr field with compact support. The main result (cf. [8,Corollary 4.4]) is as follows.

    Theorem 1.2. In [8], the limit

    ˆμq(2τˆx)limK+1K2KKkmu(ˆx,ˆx,k)¯u(ˆx,ˆx,k+τ)dk,ˆxS2, τ>0.

    holds almost surely.

    We note that the near-field data are used in [18,19], while in [8], the authors used the far-field data.

    Part of the results in [20] and [23,24] are similar to each other and we only survey the first result in [20]. In [20] the authors studied the Helmholtz equation (Δk2)u(x)=f where f is a source of migr type. Note that the potential equals zero. The main result (cf. [20,Theorem 3.9]) is similar to Theorem 1.1.

    Theorem 1.3. In [20], the limit

    μf(z)|xz|dzlimK+1K1K1k1+m|usc(x,k,ω)|2dk,xU,

    holds almost surely.

    In [21], the authors considered direct and inverse scattering for (1.1)–(1.3) with an unknown deterministic potential and a Gaussian noise source of the form σ(x)˙Bx(ω), where σ(x) is the variance and ˙Bx(ω) is the Gaussian white noise. The main result (cf. [21,Lemma 4.3]) is

    Theorem 1.4. In [21], the identity

    ^σ2(x)=42πlimj+1Kj2KjKj¯u(ˆx,k,ω)u(ˆx,k+τ,ω)dk.

    holds almost surely.

    The paper [25] extended the work [21] to the case where the source is a migr field f with μf as its rough strength and m as its rough order. For notational convenience, we shall use {Kj}P(t) to signify a sequence {Kj}jN satisfying KjCjt(jN) for some fixed constant C>0. Throughout the rest of the paper, γ stands for a fixed positive real number. The main result (cf. [25,Theorem 4.3]) is presented below.

    Theorem 1.5. In [25], assume 2<m<3 and let m=max{2/3,(3m)1/2}. Assume that {Kj}P(m+γ). Then Ω0Ω:P(Ω0)=0, Ω0 depending only on {Kj}jN, such that for any ωΩΩ0, there exists SωR3:|Sω|=0, it holds that for τR+ and ˆxS2 satisfying τˆxR3Sω,

    ˆμ(τˆx)=42πlimj+1Kj2KjKjkm¯u(ˆx,k,ω)u(ˆx,k+τ,ω)dk,

    holds for τR+ and ˆxS2 satisfying τˆxR3Sω.

    Then in [22] the authors further extended the work [25] to the case where both the potential q and the source f are random of migr type. The f (resp. q) is assumed to be supported in the domain Df (resp. Dq). In what follows, we assume that there is a positive distance between the convex hulls of the supports of f and q, i.e.,

    dist(CH(Df),CH(Dq)):=inf{|xy|;xCH(Df),yCH(Dq)}>0, (1.5)

    where CH means taking the convex hull of a domain. Therefore, one can find a plane which separates Df and Dq. In order to simplify the exposition, we assume that Df and Dq are convex domains and hence CH(Df)=Df and CH(Dq)=Dq. Moreover, we let n denote the unit normal vector of the aforementioned plane that separates Df and Dq, pointing from the half-space containing Df into the half-space containing Dq. Then the result of this work (cf. [22,Theorems 1.1 and 1.2]) is as follows.

    Theorem 1.6. In [22], suppose that f and q in system (1.1)-(1.3) are migr fields of order mf and mq, respectively, satisfying

    2<mf<4, mf<5mq11.

    Assume that (1.5) is satisfied and n is defined as above. Then, independent of μq, μf can be uniquely recovered almost surely and the recovering formula of μf is given by

    ˆμf(τˆx)={limK+42πK2KKkmf¯u(ˆx,k,ω)u(ˆx,k+τ,ω)dk,ˆxn0, ¯ˆμf(τˆx),ˆxn<0, (1.6)

    where τ0 and u(ˆx,k,ω)Mf(ω):={u(ˆx,k,ω);ˆxS2,kR+}.

    When mq<mf, μq can be uniquely recovered almost surely by the data set Mq(ω) for a fixed ωΩ. Moreover, the recovering formula is given by

    ˆμq(τˆx)={limK+42πK2KKkmq¯u(ˆx,k,ˆx,ω)u(ˆx,k+τ2,ˆx,ω)dk, ˆxn0, ¯ˆμf(τˆx),ˆxn<0, (1.7)

    where τ0 and u(ˆx,k,ˆx,ω)Mq(ω):={u(ˆx,k,ˆx,ω);ˆxS2,kR+}.

    Remark 1.1. In Theorem 1.6, the data sets Mf(ω) and Mf(ω) correspond to the case where the incident wave is passive and active, respectively. Readers may refer to [22,Section 1] for more details.

    Readers should note that the recovery formulae in Theorems 1.1–1.6 only use a single realization of the randomness; the terms on the left-hand-sides are independent of the random sample ω, while these on the right-hand-sides are limits of terms depending on ω. This feature is also described as "statistically stable" in the literature. The key ingredient of making this single-realization recovery possible is ergodicity; on the right-hand-sides of these recoveries formulae in Theorems 1.1–1.6, the probabilistic expectation operation are replaced by the average in the frequency variable and then taking to the infinity of the frequency variable. Theorems 1.1 and 1.3 utilize the near-field data to achieve the recovery, while Theorem 1.2 and 1.4–1.6 use the far-field data. Due to this difference, the corresponding techniques required in the proofs are also different. We shall present these techniques separately in Sections 3 and 4.

    The rest of this paper is organized as follows. In Section 2, we first give some preliminaries and present the well-posedness of the direct problems. In Section 3, we give the sketch of the proofs in [8,18,19,20,23,24]. Section 4 is devoted to the details of the works [22,25]. We conclude the paper in Section 5 with some remarks and open problems.

    Due to the presence of the randomness, the regularity of the potential and/or the source may be too bad to fall into the scenarios of standard PDEs techniques. In this section, we show some details used in reformulating the direct problems of (1.1)-(1.3) in a proper sense. Before that, we first present some preliminaries as well as some facts related to the migr field for the subsequent use.

    For convenient reference and self- containedness, we first present some preliminary and auxiliary results. In this paper we mainly focus on the two- and three-dimensional cases. Nevertheless, some of the results derived also hold for higher dimensions and in those cases, we choose to present the results in the general dimension n3 since they might be useful in other studies. Here we follow closely [22].

    Throughout the paper, we write L(A,B) to denote the set of all the bounded linear mappings from a normed vector space A to a normed vector space B. For any mapping KL(A,B), we denote its operator norm as KL(A,B). We also use C and its variants, such as CD, CD,f, to denote some generic constants whose particular values may change line by line. For two quantities, we write PQ to signify PCQ and PQ to signify ˜CQPCQ, for some generic positive constants C and ˜C. We write "almost everywhere" as "a.e." and "almost surely" as "a.s." for short. We use |S| to denote the Lebesgue measure of any Lebesgue-measurable set S.

    The Fourier transform and inverse Fourier transform of a function φ are respectively defined as

    Fφ(ξ)=ˆφ(ξ):=(2π)n/2eixξφ(x)dx,F1φ(ξ):=(2π)n/2eixξφ(x)dx.

    Set

    Φ(x,y)=Φk(x,y):=eik|xy|4π|xy|,xR3{y}.

    Φk is the outgoing fundamental solution, centered at y, to the differential operator Δk2. Define the resolvent operator Rk,

    (Rkφ)(x):=R3Φk(x,y)φ(y)dy,xR3, (2.1)

    where φ can be any measurable function on R3 as long as (2.1) is well-defined for almost all x in R3.

    Write x:=(1+|x|2)1/2 for xRn, n1. We introduce the following weighted Lp-norm and the corresponding function space over Rn for any δR,

    φLpδ(Rn):= δφ()Lp(Rn)=(Rnxpδ|φ|pdx)1p,Lpδ(Rn):= {φL1loc(Rn);φLpδ(Rn)<+}. (2.2)

    We also define Lpδ(S) for any subset S in Rn by replacing Rn in (2.2) with S. In what follows, we may write L2δ(Rn) as L2δ for short without ambiguities. Let I be the identity operator and define

    fHs,pδ(Rn):=(IΔ)s/2fLpδ(Rn), Hs,pδ(Rn)={fS;fHs,pδ(Rn)<+},

    where S stands for the dual space of the Schwartz space S(Rn). The space Hs,2δ(Rn) is abbreviated as Hsδ(Rn), and Hs,p0(Rn) is abbreviated as Hs,p(Rn). It can be verified that

    fHsδ(Rn)=sˆf()Hδ(Rn). (2.3)

    Let m(,+). We define Sm to be the set of all functions σ(x,ξ)C(Rn,Rn;C) such that for any two multi-indices α and β, there is a positive constant Cα,β, depending on α and β only, for which

    |(DαxDβξσ)(x,ξ)|Cα,β(1+|ξ|)m|β|,x,ξRn.

    We call any function σ in mRSm a symbol. A principal symbol of σ is an equivalent class [σ]={˜σSm;σ˜σSm1}. In what follows, we may use one representative ˜σ in [σ] to represent the equivalent class [σ]. Let σ be a symbol. Then the pseudo-differential operator T, defined on S(Rn) and associated with σ, is defined by

    (Tσφ)(x):=(2π)n/2Rneixξσ(x,ξ)ˆφ(ξ)dξ =(2π)nRn×Rnei(xy)ξσ(x,ξ)φ(y)dydξ,φS(Rn).

    Recall Definition 1.1. Lemma 2.1 below shows how the rough order of a migr field is related to its Sobolev regularity.

    Lemma 2.1. Let h be an migr distribution of rough order m in Dh. Then, hHs,p(Rn) almost surely for any 1<p<+ and s>(nm)/2.

    Proof of Lemma 2.1. See [8,Proposition 2.4].

    By the Schwartz kernel theorem (see [15,Theorem 5.2.1]), there exists a kernel Kh(x,y) with suppKhDh×Dh such that

    (Chφ)(ψ)=Eω(¯h(,ω),φh(,ω),ψ)=Kh(x,y)φ(x)ψ(y)dxdy, (2.4)

    for all φ, ψS(Rn). It is easy to verify that Kh(x,y)=¯Kh(y,x). Denote the symbol of Ch as ch, then it can be verified (see [8]) that the equalities

    {Kh(x,y)=(2π)nei(xy)ξch(x,ξ)dξ,(2.5a)ch(x,ξ)=eiξ(xy)Kh(x,y)dx,(2.5b)

    hold in the distributional sense, and the integrals in (2.5) shall be understood as oscillatory integrals. {Despite} the fact that h usually is not a function, intuitively speaking, however, it is helpful to keep in mind the following correspondence,

    Kh(x,y)Eω(¯h(x,ω)h(y,ω)).

    One way to study the direct problem of (1.1)-(1.3) is to transform it into the Lippmann-Schwinger equation, and then use the Bonn expansion to define a solution. To that end, the estimate of the operator norm of the resolvent Rk is crucial. Among different types of the estimates in the literature, one of them is known as Agmon's estimate (cf. [13,§29]). Reformulating (1.1) into the Lippmann-Schwinger equation formally (cf. [10]), we obtain

    (IRkq)usc=αRkquinRkf.

    We demonstrate two lemmas dealing with the lack of regularity when utilizing Agmon's estimates. Lemma 2.2 (cf. [25,Lemma 2.2]) shows the resolvent can take a migr field as an input without any trouble, while Lemma 2.3 (cf. [22,Theorem 2.1]) gives a variation of Agmon's estimate to fit our own problem settings.

    Lemma 2.2. Assume f is a migr field with rough order m and suppfDf almost surely, then we have RkfL21/2ϵ for any ϵ>0 almost surely.

    Proof. We split Rkf into two parts, Rk(Ef) and Rk(fEf). [21,Lemma 2.1] gives Rk(Ef)L21/2ϵ. For Rk(fEf), by using (2.4), (2.5) and (2.1), one can compute

    E(Rk(fEf)(,ω)2L21/2ϵ)=R3x12ϵE(¯fEf,Φk,xfEf,Φk,x)dx=R3x12ϵCfΦk,x,Φk,xdxx12ϵDf(DfI(y,z)eik|xz||xz||yz|2dz)eik|xy||xy|dydx, (2.6)

    where cf(y,ξ) is the symbol of the covariance operator Cf and

    I(y,z):=R3|yz|2ei(yz)ξcf(y,ξ)dξ.

    When y=z, we know I(y,z)=0 because the integrand is zero. Thanks to the condition m>2, when yz we have

    |I(y,z)|=|3j=1R3ei(yz)ξ(2ξjcf)(y,ξ)dξ|3j=1R3Cjξm2dξC0<+, (2.7)

    for some constant C0 independent of y and z. Note that if Df is bounded, then for j=1,2 we have

    Df|xy|jdyCf,jxj,xR3, (2.8)

    for some constant Cf,j depending only on f,j and the dimension. The notation x in (2.8) stands for (1+|x|2)1/2 and readers may note the difference between the and the , appeared in (2.1). With the help of (2.7) and (2.8) and Hölder's inequality, we can continue (2.6) as

    E(Rk(fEf)(,ω)2L21/2ϵ)x12ϵ(Df×Df(|xz||yz|2|xy|)1dzdy)dxx12ϵCfx2dxCf<+,

    which gives

    E(Rk(fEf)(,ω)2L21/2ϵ)Cf<+. (2.9)

    By using the Hölder inequality applied to the probability measure, we obtain from (2.9) that

    ERk(fEf)L21/2ϵ[E(Rk(fEf)2L21/2ϵ)]1/2C1/2f<+, (2.10)

    for some constant Cf independent of k. The formula (2.10) gives that Rk(fEf)L21/2ϵ almost surely, and hence RkfL21/2ϵ almost surely.

    The proof is complete.

    Lemma 2.3. For any 0<s<1/2 and ϵ>0, when k>2,

    RkφHs1/2ϵ(R3)Cϵ,sk(12s)φHs1/2+ϵ(R3),φHs1/2+ϵ(R3).

    Proof. We adopt the concept of Limiting absorption principle to first show desired results on a family of operator Rk,τ controlled by a parameter τ, and then show that Rk,τ converges in a proper sense as τ approaches zero. We sketch out the key steps in the proof and readers may refer to the proof of [22,Theorem 2.1] for complete details.

    Define an operator

    Rk,τφ(x):=(2π)3/2R3eixξˆφ(ξ)|ξ|2k2iτdξ, (2.11)

    where τR+. Fix a function χ satisfying

    {χCc(Rn),0χ1,χ(x)=1 when |x|1,χ(x)=0 when |x|2. (2.12)

    Write Rψ(x):=ψ(x). We have

     (Rk,τφ,ψ)L2(R3)=  R3Rk,τφ(x)¯ψ(x)dx=R3F{Rk,τφ}(ξ)F{R¯ψ}(ξ)dξ=  0(1χ2(rk))r2k2iτdr|ξ|=rˆφ(ξ)^R¯ψ(ξ)dS(ξ) +0r1/pr2χ2(rk)r2k2iτdr×S2[k12pˆφ(kω)][k12p^R¯ψ(kω)]dS(ω) +0r1/pr2χ2(rk)r2k2iτdrS2{[r12pˆφ(rω)][r12p^R¯ψ(rω)][k12pˆφ(kω)][k12p^R¯ψ(kω)]}dS(ω)=: I1(τ)+I2(τ)+I3(τ). (2.13)

    Here we divide (Rk,τφ,ψ)L2(R3) into three parts in order to deal with the singularity happened in the integral when |ξ| is close to k. The integral in I1 has avoided this singularity by the cutoff function χ. The singularity in I2 is only contained in the integration w.r.t. r, and it can be shown that by using Cauchy's integral theorem and choosing a proper integral path w.r.t. r, the norm of the denominator τ2k2iτ can always be bounded below by k, e.g. |τ2k2iτ|k. The singularity in I3 is compensated by the difference [] inside the integration S2[]dS(ω). In the following, we only show how to deal with I2.

    Now we estimate I1(τ). By Young's inequality abap/p+bq/q, for a,b>0,p,q>1,1/p+1/q=1 we have

    (p1/pq1/q)a1/pb1/qa+b. (2.14)

    Note that |rk|>1 in the support of the function 1χ2(rk) and |^R¯ψ(ξ)|=|ˆψ(ξ)|, one can compute

    |I1(τ)|01χ2(rk)1p1/pq1/q(r+1)1/p(k1)1/qdr|ξ|=r|ˆφ(ξ)||ˆψ(ξ)|dS(ξ)(by (2.14))Cpk1/p1φH1/(2p)δ(R3)ψH1/(2p)δ(R3), (2.15)

    where 1<p<+ and δ>0 and the Cp is independent of τ.

    We next estimate I2(τ). One has

    I2(τ)=S2[k12pˆφ(kω)][k12p^R¯ψ(kω)]0r1pr2χ2(rk)drr2k2iτdS(ω). (2.16)

    It can be shown that, by choosing a fixed τ0(0,1) carefully, we can show that the denominator pτ(r):=r2k2iτ could satisfy

    |pτ(r)|τ0kand|r|k, r{r;2|rk|τ0}Γk,τ0,τ(0,τ0), (2.17)

    where Γk,τ0:={rC;|rk|=τ0,r0}. It is obvious that the purpose of (2.17) is to use Cauchy's integral theorem. By combining (2.17) with Cauchy's integral theorem, we can continue (2.16) as

    |I2(τ)||ξ|=kξ12p|ˆφ(ξ)|ξ12p|ˆψ(ξ)|({rR+;2|rk|τ0}r1p(r/k)2τ0kdr)dS(ξ)   +|ξ|=kξ12p|ˆφ(ξ)|ξ12p|ˆψ(ξ)|(Γk,τ0(1+|r|2)12p(|r|/k)2τ0kdr)dS(ξ)Cτ0|ξ|=kξ12p|ˆφ(ξ)|ξ12p|ˆψ(ξ)|(Γk,τ0{rR+;2|rk|τ0}k1/pτ0kdr)dS(ξ)   +Cτ0|ξ|=kξ12p|ˆφ(ξ)|ξ12p|ˆψ(ξ)|(Γk,τ0k1/pτ0kdr)dS(ξ)Cτ0k1/p1(|ξ|=k|ξ12pˆh(ξ)|2dS(ξ))12(|ξ|=k|ξ12pˆψ(ξ)|2dS(ξ))12Cτ0,ϵk1/p1φH1/(2p)1/2+ϵ(R3)ψH1/(2p)1/2+ϵ(R3), (2.18)

    where the constant Cτ0,ϵ is independent of τ. Here, in deriving the last inequality in (2.18), we have made use of (2.3).

    Finally, we estimate I3(τ). Denote F(rω)=Fr(ω):=r1/(2p)ˆφ(rω) and G(rω)=Gr(ω):=r1/(2p)^Rˉψ(rω). One can compute

    |I3(τ)|0r1/pχ2(rk)|r2k2|FrL2(S2r)(r2S2|GrGk|2dS(ω))12dr   +0r1/pχ2(rk)|r2k2|(r2S2|FrFk|2dS(ω))12(rk)2GkL2(S2k)dr, (2.19)

    where S2r signifies the central sphere of radius r. Combining [13,Remark 13.1 and (13.28)] and (2.3) and (2.14), we can continue (2.19) as

    |I3(τ)|Cα,ϵ0r1/pχ2(rk)|rk|(r+k)FH1/2+ϵ(R3)|rk|αGH1/2+ϵ(R3)drCα,ϵ,p0r1/pχ2(rk)|rk|1α(r+1)1/p(k1)11/pdrFH1/2+ϵ(R3)GH1/2+ϵ(R3)Cα,ϵ,pk1/p1φH1/(2p)1/2+ϵ(R3)ψH1/(2p)1/2+ϵ(R3), (2.20)

    where the ϵ can be any positive real number and the α satisfies 0<α<ϵ, and the constant Cα,ϵ,p is independent of τ.

    Combining (2.13), (2.15), (2.18) and (2.20), we arrive at

    |(Rk,τφ,ψ)L2(R3)||I1(τ)|+|I2(τ)|+|I3(τ)|Ck1/p1φH1/(2p)1/2+ϵ(R3)ψH1/(2p)1/2+ϵ(R3),

    which implies that

    Rk,τφH1/(2p)1/2ϵ(R3)Ck1/p1φH1/(2p)1/2+ϵ(R3) (2.21)

    for some constant C independent of τ.

    Next we investigate the limiting case limτ0+Rk,τφ. Following similar steps when dealing with I1, I2 and I3, it can be shown that for any ˜τ>0, we have

    |Ij(τ1)Ij(τ2)|˜τβk1/p1φH1/(2p)1/2+ϵ(R3)ψH1/(2p)1/2+ϵ(R3),(j=1,2,3)

    holds for τ1,τ2(0,˜τ). Therefore, we can conclude

    Rk,τ1φRk,τ2φH1/(2p)1/2ϵ(R3)˜τφH1/(2p)1/2+ϵ(R3),τ1,τ2(0,˜τ),

    and thus Rk,˜τφ converges and

    lim˜τ0+Rk,˜τφ=Rkφ in H1/(2p)1/2ϵ(R3). (2.22)

    Hence from (2.21) and (2.22) we conclude that

    RkφH1/(2p)1/2ϵ(R3)Cϵ,pk(11/p)φH1/(2p)1/2+ϵ(R3)

    holds for any 1<p<+ and any ϵ>0.

    The proof is complete.

    With the help of Lemmas 2.2 and 2.3, the direct problems can be reformulated. Readers may refer to [25,Theorem 2.1], [22,Theorem 2.3], [19,Theorem 4.3], [20,Theorem 3.3], and [24,Theorem 3.3] as examples of how to formulate the direct problems, and we omit the details here.

    In this section we consider key steps in the works [8,18,19,20,23,24]. Lemma 3.3 is crucial in the key steps of the works, and its proof relies on Lemmas 3.1 and 3.2. We shall first investigate these useful lemmas.

    Lemma 3.1 is a standard result in the field of oscillatory integral and microlocal analysis.

    Lemma 3.1. Assume α and β are multi-indexes, then the following identities hold in the oscillatory integral sense,

    Rnx×Rnξeixξdxdξ=(2π)n, (3.1)
    Rnx×Rnξeixξxαξβdxdξ=(2π)ni|α|α!δαβ. (3.2)

    Here δαβ equals to 1 when α=β and equals to 0 otherwise.

    Proof. The integral in (3.1) should be understood as oscillatory integral. Fix a cutoff function χCc(Rn) with χ(0)=1, we can compute

    Rnx×Rnξeixξdxdξ=limϵ0+eixξχ(ϵx)χ(ϵξ)dxdξ=(2π)n/2limϵ0+χ(ϵ2ξ)ˆχ(ξ)dξ. (3.3)

    Denote M=supRnχ. We have |χ(ϵ2ξ)|M<. Note that χCc(Rn), so ˆχ is rapidly decaying, thus ˆχ(ξ) is Lebesgue integrable. Therefore, we can see that ˆχ(ξ)χ(ϵ2ξ) is dominated by a Lebesgue integrable function. Thus by using Lebesgue Dominated Convergence Theorem, we can continue (3.3) as

    Rnx×Rnξeixξdxdξ=(2π)n/2ˆχ(ξ)dξ=(2π)nχ(0)=(2π)n.

    We arrive at (3.1).

    To show (3.2), we first show that

    (2π)neiyηyαηβdydη=(2π)neiyηDαη(ηβ)dydη, (3.4)

    where Dηj:=1iηj. Both the LHS and RHS in (3.4) should be understood as a oscillatory integral. Thus fix some χD(Rn) such that χ(x)1 when |x|1, we have

    (3.5)

    As goes to zero, we have

    Because , . Therefore, we have

    (3.6)

    Combining (3.5) and (3.6), we arrive at

    We proved (3.4).

    Then, for multi-indexes and , if there exists such that , say, , then and so

    When , we have

    We have arrived at (3.2).

    We also need [16,Lemma 18.2.1] and we present a proof below.

    Lemma 3.2. If and is defined by the oscillatory integral

    then there exists such that

    and has the asymptotic expansion

    Remark 3.1. Note if near , e.g. for some and some cutoff function satisfying near the origin, then Lemma 3.2 implies that .

    Proof. The is the Fourier transform of with some constants, i.e.

    Then we can have

    By adopting the way used in [1,§I.8.1] in computing the oscillatory integral, we can easily show that , and this can be seen by the fact that

    so .

    The idea of the proof is to expand in terms of and by Taylor expansion

    and to use Lemma 3.1. We have

    (3.7)

    Note that the constraint in (3.7) comes from the fact that when . Moreover, the constraint "" gives

    Now we show that each remainder term in (3.7) is controlled by . Denote with underlining assumptions and , and we have

    where and is as in [1,§I.8.1]. Here we only show how the second term in the equation above is controlled by . The computation is as follows,

    thus if we take to be large enough such that , we can have

    This shows . Using the same procedure, we can show , and hence

    The proof is complete.

    We also need [16,Lemma 18.2.9] and we present a proof below.

    Lemma 3.3. Assume that and

    and a diffeomorphism preserving the hyperplane . The is -dimensional while is -dimensional. Assume and the pull-back is -smooth in , then there exists such that can be represented as

    and

    where and signify the transpose and transpose with inverse of a matrix, respectively.

    Remark 3.2. The condition " and is -smooth in " is indispensable.

    Proof. Because preserves the hyperplane , there exists a matrix-valued function such that , where the dot operation "" here signifies the matrix multiplication. According to Lemma 3.2, there exist such that Hence we have

    According to Remark 3.1, we could continue

    where with in a neighborhood such that the matrix is invertible in , and with . Using Lemma 3.2, we obtain where

    Note that satisfies so

    The proof is complete.

    Finally, we need Lemma 3.4.

    Lemma 3.4. For any stochastic process satisfying

    (3.8)

    it holds that

    (3.9)

    Proof. Check [22,Lemma 4.1].

    Lemma 3.4 turns the justification of the ergodicity into the asymptotic analysis of the expectation of related terms.

    With the help of Lemma 3.4, the most difficult part of the work [18,19,20,23,24] boils down to the estimate of the integral

    (3.10)

    where and . Readers may refer to [19,(30)-(31)], [20,(3.21) and (3.24)], [24,(4.2) and (2.1)] as well as [23,Theorems 3.1 and 3.3] as examples.

    One wonders the decaying rate of in terms of and , and after we got the decaying rate, we substitute this estimate into (3.8). If decays fast enough in terms of and/or , the corresponding integral in (3.8) will be finite and we can obtain some asymptotic ergodicity like (3.9). This is the principal idea in [18,19,20,23,24].

    Proposition 3.1. Assume is defined as in (3.10) and with is a symbol. Then for there exists constants such that

    holds uniformly for , .

    Proof. Denote , then and is the phase function. We have

    (3.11)

    We note that the part of the second term in (3.11) is always positive and the first term equals to zero when . Also, the function will be singular when . Therefore, the situation near the hyperplane is crucial for the behavior of regarding the decaying rate in terms of , . Therefore, we are willing to do a change of variables inside the integral (3.10) such that the hyperplane can be featured by a single variable, i.e. for some variable . To be specific, we choose the change of variables where

    The pull-back of under is

    (3.12)

    Second, in order to make the phase function more easy to handle, we are also willing to do another change of variables such that can be represented in the form of inner products, i.e. for some and depending on , , , , and . One of the choices is , and where

    (3.13)

    We comment that under (3.13), the phase function will only depend on and , and the choice of and is inessential as long as the change of variables is a diffeomorphism. Hence we omit the precise definitions of and and readers may refer to [18,19,20,23,24] for more details. Another thing to note is the map preserves , i.e. . By Lemma 3.3, there exists a symbol such that the pull-back of under is

    (3.14)

    By using Lemma 3.3, we can express by , and , which involves some detailed computations. Note that we only need the leading term of so the computations wouldn't be too complicated.

    The relationship (3.14) also gives

    and hence we can do the change of variables in (3.10) to obtain

    (3.15)

    Here we need the help of Lemma 3.2 to deal with the term: there exists a symbol such that

    (3.16)

    The computation of the leading term of is straight forward,

    Combining (3.15) and (3.16), we arrive at

    Now we can see is decaying at the rate of for arbitrary .

    We would like to comment that the estimation of is difficult due to the presence of the norm inside the phase function . However, the designs of and in the arguments above are so peculiar that the estimate of is possible.

    In this section we consider the key steps in the works [22,25]. In [22,25], the authors use far-field data to achieve the recovery, and this makes the derivations different from what has been discussed in Section 3. A different methodology is required to obtain accurate estimate of the decaying rate. Lemmas 4.1–4.3 plays key roles in the derivation. Before stepping into the key steps in the derivation, we shall first investigate some useful lemmas.

    First, let us recall the notion of the fractional Laplacian [30] of order in (),

    (4.1)

    where the integration is defined as an oscillatory integral. When , (4.1) can be understood as a usual Lebesgue integral if one integrates w.r.t. first and then integrates w.r.t. . By duality arguments, the fractional Laplacian can be generalized to act on wider range of functions and distributions (cf. [32]). It can be verified that the fractional Laplacian is self-adjoint.

    In the following two lemmas, we present the results in a more general form where the space dimension can be arbitrary but greater than 2, though only the case shall be used subsequently.

    Lemma 4.1. For any , we have

    in the distributional sense.

    Proof. Check [22,Lemma 3.1].

    Lemma 4.2. For any and , we have

    Proof. Check [22,Corollary 3.1].

    In the sequel, we denote .

    Lemma 4.3. Assume is a bounded domain in . For such that and , and for , there exists a constant independent of and such that

    Proof. Check [25,Lemma 3.5].

    In this subsection we restrict ourselves to . One of the key difficulty in [22] is to obtain an asymptotics about a integral

    (4.2)

    in terms of , where , and with , satisfying the requirement in Theorem 1.6, is a short notation for , and and two convex domains and satisfying (1.5). Recall the definition of the unit normal vector after (1.5). We introduce two differential operators with -smooth coefficients as follows,

    where . The operator depends on because does. Due to the fact that while , the operator is well-defined. It can be verified there is a positive lower bound of for all . It can also be verified that

    In what follows, we shall use and its variants, such as , etc., to represent some generic smooth scalar/vector functions, within or , whose particular definition may change line by line. By using integration by parts, one can compute

    (4.3)

    where the integral domain is bounded and

    and (resp. ) is the -th component of the vector (resp. ).

    Here we only show how to estimate and skip the details regarding , , and ; readers may refer to the proof of [22,Lemma 3.3] for details. For the case where , we have

    (4.4)

    Similarly, we can have

    (4.5)

    But for , if we mimic the derivation (4.4), then

    (4.6)

    Note that and thus is not absolutely integrable in . If we further differentiate the term in (4.6) by and then transfer the operator onto by using integration by parts, we would arrive at

    The term is absolutely integrable now, but the term is not integrable at the hyperplane in . To circumvent this dilemma, the fractional Laplacian can be applied as follows. By using Lemma 4.1 and 4.2, we can continue (4.6) as

    (4.7)

    where the number is chosen to satisfy , and the existence of such a number is guaranteed by noting that . Therefore, we have

    Thanks to the condition (4.8a), we can continue (4.7) as

    (4.9)

    Using similar arguments, we can also conclude that .

    Combining (4.3), (4.5) and (4.9), we arrive at

    (4.10)

    for some sufficiently large but bounded domain satisfying . Note that the integral (4.10) should be understood as a singular integral because of the presence of the singularities occurring when and . By (4.10) and (4.8b), we can finally conclude as be large enough.

    In this subsection we restrict ourselves to . We note that in (4.2), the domains and are assumed to be separated by two convex hulls. This condition is relaxed in [25] and the corresponding details in the proof is also modified. One of the key difficulty in [25] is to obtain an asymptotics about a integral

    (4.11)

    where is the kernel of the covariance operator of the migr field (cf. (2.4)), and is defined in the beginning of Section 2.1. From (4.11) we have

    (4.12)

    Define two differential operators

    It can be verified that

    Hence, noting that the integrand is compactly supported in and by using integration by part, we can continue (4.12) as

    (4.13)

    where are indices running from 1 to 3, and

    Because of the condition (cf. Theorem 1.5), we can find a number satisfying the inequalities . Therefore, we have

    By using Lemmas 4.1 and 4.2, these quantities , and can be estimated as follows:

    (4.15)

    The last inequality in (4.15) makes use of the fact (4.14a). Similarly, by first using fractional Laplacian and then using first-order differential operator on , we can have

    (4.16)
    (4.17)

    where the constant is independent of the indices , . Combining (4.13), (4.15), (4.16) and (4.17), we can rewrite (4.13) as

    (4.18)

    Denote . Then we apply Lemma 4.3 to estimate as follows,

    (4.19)

    Note that in (4.19) we used Lemma 4.3 twice. Similarly,

    (4.20)

    Recall that . By (4.18), (4.19) and (4.20) we arrive at

    We have reviewed the recoveries of some statistics by using the near-field data as well as far-field data generated under a single realization of the randomness. In this paper we mainly focus on time-harmonic Schrödinger systems. One of the possible ways to extend the current works is to study the Helmholtz systems. It would be also interesting to conduct the work in the time domain. Moreover, the stability of the recovering procedure is also worth of investigation.



    [1] S. Alinhac and P. Gérard, Pseudo-Differential Operators and the Nash-Moser Theorem, volume 82 of Graduate Studies in Mathematics, American Mathematical Society, Providence, RI, 2007. Translated from the 1991 French original by Stephen S. Wilson. doi: 10.1090/gsm/082
    [2] Inverse random source scattering problems in several dimensions. SIAM/ASA J. Uncertain. Quantif. (2016) 4: 1263-1287.
    [3] A multi-frequency inverse source problem. J. Differential Equations (2010) 249: 3443-3465.
    [4] Nonradiating sources and transmission eigenfunctions vanish at corners and edges. SIAM J. Math. Anal. (2018) 50: 6255-6270.
    [5] E. Blåsten and H. Liu, Scattering by curvatures, radiationless sources, transmission eigenfunctions and inverse scattering problems, arXiv: 1808.01425, 2018.
    [6] Super-resolution in time-reversal acoustics. J. Acoust. Soc. Am. (2002) 111: 230-248.
    [7] Adaptive interferometric imaging in clutter and optimal illumination. Inverse Prob. (2006) 22: 1405-1436.
    [8] Inverse scattering for a random potential. Anal. Appl. (2019) 17: 513-567.
    [9] The quasi-reversibility method for thermoacoustic tomography in a heterogeneous medium. SIAM J. Sci. Comput. (2007/08) 30: 1-23.
    [10] D. Colton and R. Kress, Inverse Acoustic and Electromagnetic Scattering Theory, volume 93 of Applied Mathematical Sciences, Springer, Cham, [2019] ⓒ2019. Fourth edition of [MR1183732]. doi: 10.1007/978-3-030-30351-8
    [11] On identifying magnetized anomalies using geomagnetic monitoring. Arch. Ration. Mech. Anal. (2019) 231: 153-187.
    [12] On identifying magnetized anomalies using geomagnetic monitoring within a magnetohydrodynamic model. Arch. Ration. Mech. Anal. (2020) 235: 691-721.
    [13] G. Eskin, Lectures on Linear Partial Differential Equations, volume 123. American Mathematical Society, 2011. doi: 10.1090/gsm/123
    [14] (2016) Introduction to Quantum Mechanics. Cambridge: Cambridge Univ. Press.
    [15] L. Hörmander, The Analysis of Linear Partial Differential Operators I. Distribution Theory and Fourier Analysis, Springer, Berlin, second edition edition, 1990. doi: 10.1007/978-3-642-61497-2
    [16] L. Hörmander, The Analysis of Linear Partial Differential Operators III. Pseudo-Differential Operators, Springer, Berlin, 1994 edition edition, 2007. doi: 10.1007/978-3-540-49938-1
    [17] C. Knox and A. Moradifam, Determining both the source of a wave and its speed in a medium from boundary measurements, Inverse Prob., 36 (2020), 025002, 15 pp. doi: 10.1088/1361-6420/ab53fc
    [18] M. Lassas, L. Päivärinta and E. Saksman, Inverse problem for a random potential, In Partial Differential Equations and Inverse Problems, volume 362 of Contemp. Math., pages 277–288. Amer. Math. Soc., Providence, RI, 2004. doi: 10.1090/conm/362/06618
    [19] Inverse scattering problem for a two dimensional random potential. Comm. Math. Phys. (2008) 279: 669-703.
    [20] Inverse random source problems for time-harmonic acoustic and elastic waves. Comm. Partial Differential Equations (2020) 45: 1335-1380.
    [21] Determining a random Schrödinger equation with unknown source and potential. SIAM J. Math. Anal. (2019) 51: 3465-3491.
    [22] J. Li, H. Liu and S. Ma, Determining a random Schrödinger operator: both potential and source are random, Comm. Math. Phys., 2020. doi: 10.1007/s00220-020-03889-9
    [23] P. Li and X. Wang, An inverse random source problem for maxwell's equations, arXiv: 2002.08732, 2020.
    [24] P. Li and X. Wang, Inverse random source scattering for the Helmholtz equation with attenuation, arXiv: 1911.11189, 2019.
    [25] H. Liu and S. Ma, Single-realization recovery of a random Schrödinger equation with unknown source and potential, arXiv: 2005.04984, 2020.
    [26] H. Liu and G. Uhlmann, Determining both sound speed and internal source in thermo-and photo-acoustic tomography, Inverse Prob., 31 (2015), 105005, 10 pp. doi: 10.1088/0266-5611/31/10/105005
    [27] Global uniqueness for an inverse stochastic hyperbolic problem with three unknowns. Comm. Pure Appl. Math. (2015) 68: 948-963.
    [28] S. Ma, Determination of random Schrödinger operators, Open Access Theses and Dissertations, 671, Hong Kong Baptist University, 2019.
    [29] (2000) Strongly Elliptic Systems and Boundary Integral Equations. Cambridgen University Press.
    [30] C. Pozrikidis, The Fractional Laplacian, Chapman & Hall/CRC, New York, 2016. doi: 10.1201/b19666
    [31] X. Wang, Y. Guo, D. Zhang and H. Liu, Fourier method for recovering acoustic sources from multi-frequency far-field data, Inverse Prob., 33 (2017), 035001, 18 pp. doi: 10.1088/1361-6420/aa573c
    [32] M. W. Wong, An Introduction to Pseudo-Differential Operators, volume 6 of Series on Analysis, Applications and Computation, World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, third edition, 2014. doi: 10.1142/9074
    [33] G. Yuan, Determination of two kinds of sources simultaneously for a stochastic wave equation, Inverse Prob., 31 (2015), 085003, 13 pp. doi: 10.1088/0266-5611/31/8/085003
    [34] D. Zhang and Y. Guo, Fourier method for solving the multi-frequency inverse source problem for the Helmholtz equation, Inverse Prob., 31 (2015), 035007, 30 pp. doi: 10.1088/0266-5611/31/3/035007
  • This article has been cited by:

    1. Jianliang Li, Peijun Li, Xu Wang, Inverse Random Potential Scattering for Elastic Waves, 2023, 21, 1540-3459, 426, 10.1137/22M1497183
    2. Qi Lü, Xu Zhang, Inverse problems for stochastic partial differential equations: Some progresses and open problems, 2023, 0, 2155-3289, 0, 10.3934/naco.2023014
  • Reader Comments
  • © 2021 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2753) PDF downloads(220) Cited by(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog