ROM inversion of monostatic data lifted to full MIMO.

V. Druskin111Worcester Polytechnic Institute, Department of Mathematical Sciences, Stratton Hall, 100 Institute Road, Worcester MA, 01609 ([email protected])    S. Moskow222Department of Mathematics, Drexel University, Korman Center, 3141 Chestnut Street, Philadelphia, PA 19104 ([email protected]) and M. Zaslavsky333Southern Methodist University, Department of Mathematics, Clements Hall, 3100 Dyer st, Dallas, TX 75205 ([email protected])
Abstract

The Lippmann–Schwinger–Lanczos (LSL) algorithm has recently been shown to provide an efficient tool for imaging and direct inversion of synthetic aperture radar data in multi-scattering environments [17], where the data set is limited to the monostatic, a.k.a. single input/single output (SISO) measurements. The approach is based on constructing data-driven estimates of internal fields via a reduced-order model (ROM) framework and then plugging them into the Lippmann-Schwinger integral equation. However, the approximations of the internal solutions may have more error due to missing the off diagonal elements of the multiple input/multiple output (MIMO) matrix valued transfer function. This, in turn, may result in multiple echoes in the image. Here we present a ROM-based data completion algorithm to mitigate this problem. First, we apply the LSL algorithm to the SISO data as in [17] to obtain approximate reconstructions as well as the estimate of internal field. Next, we use these estimates to calculate a forward Lippmann-Schwinger integral to populate the missing off-diagonal data (the lifting step). Finally, to update the reconstructions, we solve the Lippmann-Schwinger equation using the original SISO data, where the internal fields are constructed from the lifted MIMO data. The steps of obtaining the approximate reconstructions and internal fields and populating the missing MIMO data entries can be repeated for complex models to improve the images even further. Efficiency of the proposed approach is demonstrated on 2D and 2.5D numerical examples, where we see reconstructions are improved substantially.

1 Introduction

Although the Born approximation and linear sampling methods [11, 10] are methods of choice for synthetic aperture radar (SAR) due to their speed and versatility, nonlinear scattering is known to produce significant artifacts. Approaches to deal with multiple scattering include optimization [21], improved linearizations using some knowledge of an inhomogeneous background [18, 27], statistical and machine learning methods [22], iterative time reversal and matching filtering [20], and reweighting [12]. Also in seismic imaging are full wave inversion [26], boundary control [19], and inverse series methods [23]. Many of the above approaches require prior knowledge, the solution of full scale forward problems or large training sets.

Reduced order model (ROM) approaches have been shown to provide accurate images for several inverse impedance, scattering and diffusion problems [3, 5, 14, 13, 6, 7, 4, 2, 15, 9, 16, 8]. In the recent work [17], we applied the Lippmann-Schwinger-Lanczos algorithm [15] to models of synthetic aperture radar. Most of the ROM-based imaging approaches strongly rely on having a complete square MIMO dataset accessible, as seen in the schematic (1)

(...............)matrixabsentabsentabsentabsentabsentmissing-subexpressionabsentabsentabsentabsentmissing-subexpressionmissing-subexpressionabsentabsentabsentmissing-subexpressionmissing-subexpressionmissing-subexpressionabsentabsentmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionabsent\begin{pmatrix}.&.&.&.&.\\ \ &.&.&.&.\\ \ &\ &.&.&.\\ \ &\ &\ &.&.\\ \ &\ &\ &\ &.\\ \end{pmatrix}( start_ARG start_ROW start_CELL . end_CELL start_CELL . end_CELL start_CELL . end_CELL start_CELL . end_CELL start_CELL . end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL . end_CELL start_CELL . end_CELL start_CELL . end_CELL start_CELL . end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL . end_CELL start_CELL . end_CELL start_CELL . end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL . end_CELL start_CELL . end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL . end_CELL end_ROW end_ARG ) (1)

where the column and row numbers correspond respectively to the indices of the receivers (outputs) and transmitters (inputs). The other half of the transfer matrix extends by reciprocity. In SAR, as well as in conventional (non-array) medical ultrasound, profiling is done by mono-static transmitter/receivers, and images are constructed by combining multiple single-input/single output (SISO) data. The data in this case is visualized in (2),

(.....)matrixabsentmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionabsentmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionabsentmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionabsentmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionabsent\begin{pmatrix}.&&&&\\ \ &.&&&\\ \ &\ &.&&\\ \ &\ &\ &.&\\ \ &\ &\ &\ &.\\ \end{pmatrix}( start_ARG start_ROW start_CELL . end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL . end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL . end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL . end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL . end_CELL end_ROW end_ARG ) (2)

where we have only the diagonal. The nature of the Lippmann-Schwinger-Lanczos (LSL) algorithm allowed us to extend the ROM approach naturally to this case in [17]. LSL for monostatic data [17] contains two major steps:

  • Step (i) Use the individual SISO ROMs to generate internal fields

  • Step (ii) Use these internal fields to linearize the Lippmann-Schwinger equation and solve for the profile.

Using this method we were able to substantially reduce the echos compared to the Born approximation, with minimal increase to the computational cost. However, due to the lack of the off diagonal data, we did not have the full strength of the MIMO ROM, and so the data generated internal fields were less accurate. In this work we propose to use the SISO ROMs to lift the data to improve the data generated internal fields. We start by first implementing LSL for monostatic data. Once we have approximate internal fields and profile, we do a forward application of the Lippmann-Schwinger integral to approximate the off-diagonals. That is, we complete the data so that it is of the form (3)

(.....)matrixabsentabsentabsentabsentabsent\begin{pmatrix}.&\circ&\circ&\circ&\circ\\ \circ&.&\circ&\circ&\circ\\ \circ&\circ&.&\circ&\circ\\ \circ&\circ&\circ&.&\circ\\ \circ&\circ&\circ&\circ&.\\ \end{pmatrix}( start_ARG start_ROW start_CELL . end_CELL start_CELL ∘ end_CELL start_CELL ∘ end_CELL start_CELL ∘ end_CELL start_CELL ∘ end_CELL end_ROW start_ROW start_CELL ∘ end_CELL start_CELL . end_CELL start_CELL ∘ end_CELL start_CELL ∘ end_CELL start_CELL ∘ end_CELL end_ROW start_ROW start_CELL ∘ end_CELL start_CELL ∘ end_CELL start_CELL . end_CELL start_CELL ∘ end_CELL start_CELL ∘ end_CELL end_ROW start_ROW start_CELL ∘ end_CELL start_CELL ∘ end_CELL start_CELL ∘ end_CELL start_CELL . end_CELL start_CELL ∘ end_CELL end_ROW start_ROW start_CELL ∘ end_CELL start_CELL ∘ end_CELL start_CELL ∘ end_CELL start_CELL ∘ end_CELL start_CELL . end_CELL end_ROW end_ARG ) (3)

where the symbol \circ represents approximate, or lifted, data. With a full (yet approximate) transfer matrix in hand, we can use a MIMO ROM to construct improved internal fields, for a new Step (i). These improved fields are then used in the Lippmann Schwinger equation in Step (ii). Evidently, at Step (ii), we use only the true diagonal data, since the lifted off diagonals will not provide any new information.

We note that the lifted data matrix (3) does not have the correct structure apriori, that is to say, the resulting mass matrix may not be symmetric positive definite. Regularization will therefore be necessary at this step separately, in addition to the usual regularization required for the inversion of the Lippmann-Schwinger equation. We also emphasize that lifting doesn’t provide additional data for the inverse problem; it is used solely to improve the ROM’s performance in generating the internal fields.

We now make a remark about ROM based inversion methods. In general, a complete ROM is a reduced version of the full scale PDE in the form of a Galerkin system, containing both a mass matrix M𝑀Mitalic_M (the Grammian used here) and a stiffness matrix. Historically, the complete reduced order model was needed to do inversion (see for example [14]). It was later found that in the time domain the stiffness matrix was not needed for the LSL approach, hence in this work we discuss only M𝑀Mitalic_M, and refer to this as the ROM.

This paper is organized as follows. In Section 2 we state precisely the model and inverse problem and give an overview of the entire process. Sections 3,4, and 5 contain detailed descriptions of the steps, which are monostatic LSL, data completion, and MIMO generation of internal fields, respectively. In Section 6 we present numerical experiments demonstrating the improved performance. We present a brief conclusion in Section 7. As a notational note, throughout the manuscript, we use boldface to denote data generated approximations while the regular script denotes actual values.

2 Statement of the inverse problem and algorithm overview

We assume the following wave model problem for a domain ΩnΩsuperscript𝑛\Omega\subset\mathbb{R}^{n}roman_Ω ⊂ blackboard_R start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT

utt+Au=0inΩ×[0,)subscript𝑢𝑡𝑡𝐴𝑢0inΩ0u_{tt}+Au=0\ \mbox{in}\ \Omega\times[0,\infty)italic_u start_POSTSUBSCRIPT italic_t italic_t end_POSTSUBSCRIPT + italic_A italic_u = 0 in roman_Ω × [ 0 , ∞ ) (4)

with initial conditions

u(t=0)𝑢𝑡0\displaystyle u(t=0)italic_u ( italic_t = 0 ) =\displaystyle== giinΩsubscript𝑔𝑖inΩ\displaystyle g_{i}\ \mbox{in}\ \Omegaitalic_g start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT in roman_Ω (5)
ut(t=0)subscript𝑢𝑡𝑡0\displaystyle u_{t}(t=0)italic_u start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_t = 0 ) =\displaystyle== 0inΩ0inΩ\displaystyle 0\ \mbox{in}\ \Omega0 in roman_Ω (6)

where {gi}subscript𝑔𝑖\{g_{i}\}{ italic_g start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } for i=1,K𝑖1𝐾i=1,\ldots Kitalic_i = 1 , … italic_K are Dirichlet initial conditions representing localized sources near the boundary,***The formulation for the homogeneous wave equation (4) with an inhomogeneous initial condition (5) can be equivalently derived from the more conventional radar formulation of an inhomogeneous wave equation with homogeneous initial conditions. We refer to [14] for details. and we assume homogeneous Neumann boundary conditions on the spatial boundary ΩΩ\partial\Omega∂ roman_Ω. The operator A𝐴Aitalic_A is taken to be the negative Laplacian plus a scattering term:

A=Δ+q𝐴Δ𝑞A=-\Delta+qitalic_A = - roman_Δ + italic_q (7)

where q(x)0𝑞𝑥0q(x)\geq 0italic_q ( italic_x ) ≥ 0 is our unknown potential, not necessarily small, but with compact support. The initial data is assumed to be localized at source i𝑖iitalic_i, and the corresponding exact forward solutions to (4-5) are

u(i)(x,t)=cos(At)gi(x),superscript𝑢𝑖𝑥𝑡𝐴𝑡subscript𝑔𝑖𝑥u^{(i)}(x,t)=\cos{(\sqrt{A}t)}g_{i}(x),italic_u start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_x , italic_t ) = roman_cos ( square-root start_ARG italic_A end_ARG italic_t ) italic_g start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_x ) , (8)

where the square root and cosine are defined via the spectral theorem. These solutions are assumed to be unknown, except near the receiver. The data here is monostatic, that is, we measure the data from source i𝑖iitalic_i only back at the receiver collocated with the source. The data is measured at the 2n12𝑛12n-12 italic_n - 1 evenly spaced time steps t=kτ𝑡𝑘𝜏t=k\tauitalic_t = italic_k italic_τ for k=0,,2n2𝑘02𝑛2k=0,\dots,2n-2italic_k = 0 , … , 2 italic_n - 2 and is modeled by

Fii(kτ)superscript𝐹𝑖𝑖𝑘𝜏\displaystyle F^{ii}(k\tau)italic_F start_POSTSUPERSCRIPT italic_i italic_i end_POSTSUPERSCRIPT ( italic_k italic_τ ) =\displaystyle== Ωgi(x)u(i)(x,kτ)𝑑xsubscriptΩsubscript𝑔𝑖𝑥superscript𝑢𝑖𝑥𝑘𝜏differential-d𝑥\displaystyle\int_{\Omega}g_{i}(x)u^{(i)}(x,k\tau)dx∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT italic_g start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_x ) italic_u start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_x , italic_k italic_τ ) italic_d italic_x (9)
=\displaystyle== Ωgi(x)cos(Akτ)gi(x)𝑑x.subscriptΩsubscript𝑔𝑖𝑥𝐴𝑘𝜏subscript𝑔𝑖𝑥differential-d𝑥\displaystyle\int_{\Omega}g_{i}(x)\cos{(\sqrt{A}k\tau)}g_{i}(x)dx.∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT italic_g start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_x ) roman_cos ( square-root start_ARG italic_A end_ARG italic_k italic_τ ) italic_g start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_x ) italic_d italic_x .

The inverse problem is as follows: Given

{Fii(kτ)} for k=0,,2n2;i=1,K,formulae-sequencesuperscript𝐹𝑖𝑖𝑘𝜏 for 𝑘02𝑛2𝑖1𝐾\{F^{ii}(k\tau)\}\ \mbox{ for }\ k=0,\dots,2n-2;\ \ i=1,\ldots K,{ italic_F start_POSTSUPERSCRIPT italic_i italic_i end_POSTSUPERSCRIPT ( italic_k italic_τ ) } for italic_k = 0 , … , 2 italic_n - 2 ; italic_i = 1 , … italic_K ,

reconstruct q𝑞qitalic_q. We will also need to assume we have background fields {u(i),0}superscript𝑢𝑖0\{u^{(i),0}\}{ italic_u start_POSTSUPERSCRIPT ( italic_i ) , 0 end_POSTSUPERSCRIPT }, solutions to (4-5) corresponding to q=0𝑞0q=0italic_q = 0. Denote the full K×K𝐾𝐾K\times Kitalic_K × italic_K background transfer function by F0(t)subscript𝐹0𝑡F_{0}(t)italic_F start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_t ) and background solution anti-derivatives by

w(i),0(x,t)=0tu(i),0(x,s)𝑑s.superscript𝑤𝑖0𝑥𝑡subscriptsuperscript𝑡0superscript𝑢𝑖0𝑥𝑠differential-d𝑠w^{(i),0}(x,t)=\int^{t}_{0}{u^{(i),0}(x,s)ds}.italic_w start_POSTSUPERSCRIPT ( italic_i ) , 0 end_POSTSUPERSCRIPT ( italic_x , italic_t ) = ∫ start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_u start_POSTSUPERSCRIPT ( italic_i ) , 0 end_POSTSUPERSCRIPT ( italic_x , italic_s ) italic_d italic_s .

Note that {w(i),0}superscript𝑤𝑖0\{w^{(i),0}\}{ italic_w start_POSTSUPERSCRIPT ( italic_i ) , 0 end_POSTSUPERSCRIPT } satisfy (4) for q=0𝑞0q=0italic_q = 0 with homogeneous Neumann boundary conditions and Neumann initial conditions

w(i),0(t=0)superscript𝑤𝑖0𝑡0\displaystyle w^{(i),0}(t=0)italic_w start_POSTSUPERSCRIPT ( italic_i ) , 0 end_POSTSUPERSCRIPT ( italic_t = 0 ) =\displaystyle== 0inΩ0inΩ\displaystyle 0\ \mbox{in}\ \Omega0 in roman_Ω (10)
wt(i),0(t=0)subscriptsuperscript𝑤𝑖0𝑡𝑡0\displaystyle w^{(i),0}_{t}(t=0)italic_w start_POSTSUPERSCRIPT ( italic_i ) , 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_t = 0 ) =\displaystyle== giinΩsubscript𝑔𝑖inΩ\displaystyle g_{i}\ \mbox{in}\ \Omegaitalic_g start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT in roman_Ω (11)

Throughout the process we will use the following form of the Lippmann-Schwinger equation

F0ij(t)Fij(t)=0tΩw(j),0(x,tτ)u(i)(x,τ)q(x)𝑑x𝑑τ.subscriptsuperscript𝐹𝑖𝑗0𝑡superscript𝐹𝑖𝑗𝑡superscriptsubscript0𝑡subscriptΩsuperscript𝑤𝑗0𝑥𝑡𝜏superscript𝑢𝑖𝑥𝜏𝑞𝑥differential-d𝑥differential-d𝜏F^{ij}_{0}(t)-F^{ij}(t)=\int_{0}^{t}\int_{\Omega}w^{(j),0}(x,t-\tau)u^{(i)}(x,% \tau)q(x)dxd\tau.italic_F start_POSTSUPERSCRIPT italic_i italic_j end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_t ) - italic_F start_POSTSUPERSCRIPT italic_i italic_j end_POSTSUPERSCRIPT ( italic_t ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT italic_w start_POSTSUPERSCRIPT ( italic_j ) , 0 end_POSTSUPERSCRIPT ( italic_x , italic_t - italic_τ ) italic_u start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_x , italic_τ ) italic_q ( italic_x ) italic_d italic_x italic_d italic_τ . (12)

We remind the reader that the inverse Born method is to replace the unknown field u(i)superscript𝑢𝑖u^{(i)}italic_u start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT in (12) with the corresponding background solution u(i),0superscript𝑢𝑖0u^{(i),0}italic_u start_POSTSUPERSCRIPT ( italic_i ) , 0 end_POSTSUPERSCRIPT, and solve for q𝑞qitalic_q. The LSL method is to instead use the ROM-based data generated internal fields, which we denote 𝐮𝐮{\bf u}bold_u. As one would expect, these data generated fields would be much more accurate were we to have the full MIMO data {Fij}superscript𝐹𝑖𝑗\{F^{ij}\}{ italic_F start_POSTSUPERSCRIPT italic_i italic_j end_POSTSUPERSCRIPT } for i,j=1,Kformulae-sequence𝑖𝑗1𝐾i,j=1,\dots Kitalic_i , italic_j = 1 , … italic_K instead of the SISO only data {Fii}superscript𝐹𝑖𝑖\{F^{ii}\}{ italic_F start_POSTSUPERSCRIPT italic_i italic_i end_POSTSUPERSCRIPT }. This is precisely the motivation for this work. We propose the following algorithm:

Algorithm 1 (Data completion for ROM-based internal solutions)

Given monostatic data {Fii(kτ)}superscript𝐹𝑖𝑖𝑘𝜏\{F^{ii}(k\tau)\}{ italic_F start_POSTSUPERSCRIPT italic_i italic_i end_POSTSUPERSCRIPT ( italic_k italic_τ ) } for k=0,,2n2𝑘02𝑛2k=0,\dots,2n-2italic_k = 0 , … , 2 italic_n - 2 for sources i=1,,K𝑖1𝐾i=1,\ldots,Kitalic_i = 1 , … , italic_K:

  1. 1.

    SISO step: Use LSL for true SISO data F𝐹Fitalic_F to generate approximate 𝐪𝐪{\bf q}bold_q and internal fields 𝐮(i)superscript𝐮𝑖{\bf u}^{(i)}bold_u start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT.

  2. 2.

    Lifting step: given approximate 𝐪𝐪{\bf q}bold_q and internal solutions 𝐮(i)superscript𝐮𝑖{\bf u}^{(i)}bold_u start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT, compute 𝐅𝐅{\bf F}bold_F for all inaccessible pairs (i;j)𝑖𝑗(i;j)( italic_i ; italic_j )

  3. 3.

    MIMO step: given a full 𝐅𝐅{\bf F}bold_F and its mass matrix 𝐌𝐌{\bf M}bold_M, compute new internal fields 𝐮(i)superscript𝐮𝑖{\bf u}^{(i)}bold_u start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT and new approximate 𝐪𝐪{\bf q}bold_q from Lippmann-Schwinger.

  4. 4.

    Exit or return to lifting step 2.

3 Step 1: The LSL algorithm for SISO data

This step is precisely to perform the algorithm from [17]. We include a short description here, see [17] and also [15],[16] for more details. We assume we are given only the diagonal response

Fjj(kτ)=Ωgj(x)cos(Akτ)gj(x)𝑑x,superscript𝐹𝑗𝑗𝑘𝜏subscriptΩsubscript𝑔𝑗𝑥𝐴𝑘𝜏subscript𝑔𝑗𝑥differential-d𝑥F^{jj}(k\tau)=\int_{\Omega}g_{j}(x)\cos{(\sqrt{A}k\tau)}g_{j}(x)dx,italic_F start_POSTSUPERSCRIPT italic_j italic_j end_POSTSUPERSCRIPT ( italic_k italic_τ ) = ∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT italic_g start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_x ) roman_cos ( square-root start_ARG italic_A end_ARG italic_k italic_τ ) italic_g start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_x ) italic_d italic_x , (13)

for j=1,,K𝑗1𝐾j=1,\dots,Kitalic_j = 1 , … , italic_K, k=0,,2n1𝑘02𝑛1k=0,\ldots,2n-1italic_k = 0 , … , 2 italic_n - 1. For each source, we will compute a mass matrix and corresponding approximate internal snapshots. Let u(j)(x,t)superscript𝑢𝑗𝑥𝑡u^{(j)}(x,t)italic_u start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT ( italic_x , italic_t ) be the true internal fields corresponding to source gjsubscript𝑔𝑗g_{j}italic_g start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, and consider the snapshots

uk(j)=u(j)(kτ,x)subscriptsuperscript𝑢𝑗𝑘superscript𝑢𝑗𝑘𝜏𝑥u^{(j)}_{k}=u^{(j)}(k\tau,x)italic_u start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_u start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT ( italic_k italic_τ , italic_x )

for k=0,,2n1𝑘02𝑛1k=0,\ldots,2n-1italic_k = 0 , … , 2 italic_n - 1. The n×n𝑛𝑛n\times nitalic_n × italic_n mass matrix corresponding to this source is defined by

Mkl(j)=Ωuk(j)ul(j)𝑑xsubscriptsuperscript𝑀𝑗𝑘𝑙subscriptΩsubscriptsuperscript𝑢𝑗𝑘subscriptsuperscript𝑢𝑗𝑙differential-d𝑥M^{(j)}_{kl}=\int_{\Omega}u^{(j)}_{k}u^{(j)}_{l}dxitalic_M start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT = ∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT italic_u start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_u start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT italic_d italic_x (14)

for k,l=0,,n1formulae-sequence𝑘𝑙0𝑛1k,l=0,\ldots,n-1italic_k , italic_l = 0 , … , italic_n - 1. Given the expression (8) for the exact solution, we have that,

Mkl(j)=Ωgj(x)cos(Akτ)cos(Alτ)gj(x)𝑑x.subscriptsuperscript𝑀𝑗𝑘𝑙subscriptΩsubscript𝑔𝑗𝑥𝐴𝑘𝜏𝐴𝑙𝜏subscript𝑔𝑗𝑥differential-d𝑥M^{(j)}_{kl}=\int_{\Omega}g_{j}(x)\cos{(\sqrt{A}k\tau)}\cos{(\sqrt{A}l\tau)}g_% {j}(x)dx.italic_M start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT = ∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT italic_g start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_x ) roman_cos ( square-root start_ARG italic_A end_ARG italic_k italic_τ ) roman_cos ( square-root start_ARG italic_A end_ARG italic_l italic_τ ) italic_g start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_x ) italic_d italic_x . (15)

The steps are as follows.

  1. 1.

    Obtain the mass matrices from the data: Compute

    Mkl(j)=12(Fjj((kl)τ)+Fjj((k+l)τ)).subscriptsuperscript𝑀𝑗𝑘𝑙12superscript𝐹𝑗𝑗𝑘𝑙𝜏superscript𝐹𝑗𝑗𝑘𝑙𝜏M^{(j)}_{kl}=\frac{1}{2}\left(F^{jj}((k-l)\tau)+F^{jj}((k+l)\tau)\right).italic_M start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_F start_POSTSUPERSCRIPT italic_j italic_j end_POSTSUPERSCRIPT ( ( italic_k - italic_l ) italic_τ ) + italic_F start_POSTSUPERSCRIPT italic_j italic_j end_POSTSUPERSCRIPT ( ( italic_k + italic_l ) italic_τ ) ) . (16)

    This formula follows from (13), (15), and the cosine angle sum formula.

  2. 2.

    Orthogonalize the snapshots: Compute the Cholesky decompositions (note each M(j)superscript𝑀𝑗M^{(j)}italic_M start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT is positive definite),

    M(j)=(U(j))U(j)superscript𝑀𝑗superscriptsuperscript𝑈𝑗topsuperscript𝑈𝑗M^{(j)}=(U^{(j)})^{\top}{U^{(j)}}italic_M start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT = ( italic_U start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT italic_U start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT

    where each U(j)superscript𝑈𝑗U^{(j)}italic_U start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT is upper triangular, for j=1,,m𝑗1𝑚j=1,\ldots,mitalic_j = 1 , … , italic_m. This allows us to define the sequentially orthogonalized snapshots

    v(j)=u(j)(U(j))1,superscript𝑣𝑗superscript𝑢𝑗superscriptsuperscript𝑈𝑗1\vec{v}^{(j)}=\vec{u}^{(j)}(U^{(j)})^{-1},over→ start_ARG italic_v end_ARG start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT = over→ start_ARG italic_u end_ARG start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT ( italic_U start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT , (17)

    where u(j)superscript𝑢𝑗\vec{u}^{(j)}over→ start_ARG italic_u end_ARG start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT is the column vector of the true (still unknown) snapshots.

  3. 3.

    Repeat for the background medium: Compute the background snapshots {u(j),0}superscript𝑢𝑗0\{u^{(j),0}\}{ italic_u start_POSTSUPERSCRIPT ( italic_j ) , 0 end_POSTSUPERSCRIPT }, the mass matrices

    Mkl(j),0=Ωuk(j),0ul(j),0𝑑x,subscriptsuperscript𝑀𝑗0𝑘𝑙subscriptΩsubscriptsuperscript𝑢𝑗0𝑘subscriptsuperscript𝑢𝑗0𝑙differential-d𝑥M^{(j),0}_{kl}=\int_{\Omega}u^{(j),0}_{k}u^{(j),0}_{l}dx,italic_M start_POSTSUPERSCRIPT ( italic_j ) , 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT = ∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT italic_u start_POSTSUPERSCRIPT ( italic_j ) , 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_u start_POSTSUPERSCRIPT ( italic_j ) , 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT italic_d italic_x , (18)

    their Cholesky decompositions

    M(j),0=(U(j),0)U(j),0,superscript𝑀𝑗0superscriptsuperscript𝑈𝑗0topsuperscript𝑈𝑗0M^{(j),0}=(U^{(j),0})^{\top}U^{(j),0},italic_M start_POSTSUPERSCRIPT ( italic_j ) , 0 end_POSTSUPERSCRIPT = ( italic_U start_POSTSUPERSCRIPT ( italic_j ) , 0 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT italic_U start_POSTSUPERSCRIPT ( italic_j ) , 0 end_POSTSUPERSCRIPT ,

    and the orthogonalized background snapshots,

    v(j),0=u(j),0(U(j),0)1.superscript𝑣𝑗0superscript𝑢𝑗0superscriptsuperscript𝑈𝑗01\vec{v}^{(j),0}=\vec{u}^{(j),0}(U^{(j),0})^{-1}.over→ start_ARG italic_v end_ARG start_POSTSUPERSCRIPT ( italic_j ) , 0 end_POSTSUPERSCRIPT = over→ start_ARG italic_u end_ARG start_POSTSUPERSCRIPT ( italic_j ) , 0 end_POSTSUPERSCRIPT ( italic_U start_POSTSUPERSCRIPT ( italic_j ) , 0 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT . (19)
  4. 4.

    Compute data generated internal snapshots: In this step we replace the unknown orthogonalized snapshots v(j)superscript𝑣𝑗\vec{v}^{(j)}over→ start_ARG italic_v end_ARG start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT with the orthogonalized background snapshots v(j),0superscript𝑣𝑗0\vec{v}^{(j),0}over→ start_ARG italic_v end_ARG start_POSTSUPERSCRIPT ( italic_j ) , 0 end_POSTSUPERSCRIPT. They should be close in the sense of spherical averages; see [17] Section 4.2 for a discussion and Figure 7.7 in Section 7 here for a numerical example comparing internal solutions. From (17) and (19), we have our approximations to the internal snapshots:

    𝐮(j)superscript𝐮𝑗\displaystyle\vec{\bf{u}}^{(j)}over→ start_ARG bold_u end_ARG start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT =\displaystyle== v(j),0U(j)superscript𝑣𝑗0superscript𝑈𝑗\displaystyle\vec{v}^{(j),0}U^{(j)}over→ start_ARG italic_v end_ARG start_POSTSUPERSCRIPT ( italic_j ) , 0 end_POSTSUPERSCRIPT italic_U start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT
    =\displaystyle== u(j),0(U(j),0)1U(j).superscript𝑢𝑗0superscriptsuperscript𝑈𝑗01superscript𝑈𝑗\displaystyle\vec{u}^{(j),0}(U^{(j),0})^{-1}U^{(j)}.over→ start_ARG italic_u end_ARG start_POSTSUPERSCRIPT ( italic_j ) , 0 end_POSTSUPERSCRIPT ( italic_U start_POSTSUPERSCRIPT ( italic_j ) , 0 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_U start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT .
  5. 5.

    Invert the Lippmann-Schwinger equation: In order to obtain an estimate for the unknown q𝑞qitalic_q, we now use the Lippmann-Schwinger equation. For each time step kτ𝑘𝜏k\tauitalic_k italic_τ, k=0,,n1𝑘0𝑛1k=0,\dots,n-1italic_k = 0 , … , italic_n - 1, and for each source j=1,,m𝑗1𝑚j=1,\ldots,mitalic_j = 1 , … , italic_m, we have that

    F0jj(kτ)Fjj(kτ)=0kτΩw(j),0(x,kτt)u(j)(x,t)q(x)𝑑x𝑑t.subscriptsuperscript𝐹𝑗𝑗0𝑘𝜏superscript𝐹𝑗𝑗𝑘𝜏superscriptsubscript0𝑘𝜏subscriptΩsuperscript𝑤𝑗0𝑥𝑘𝜏𝑡superscript𝑢𝑗𝑥𝑡𝑞𝑥differential-d𝑥differential-d𝑡F^{jj}_{0}(k\tau)-F^{jj}(k\tau)=\int_{0}^{k\tau}\int_{\Omega}w^{(j),0}(x,k\tau% -t)u^{(j)}(x,t)q(x)dxdt.italic_F start_POSTSUPERSCRIPT italic_j italic_j end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_k italic_τ ) - italic_F start_POSTSUPERSCRIPT italic_j italic_j end_POSTSUPERSCRIPT ( italic_k italic_τ ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k italic_τ end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT italic_w start_POSTSUPERSCRIPT ( italic_j ) , 0 end_POSTSUPERSCRIPT ( italic_x , italic_k italic_τ - italic_t ) italic_u start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT ( italic_x , italic_t ) italic_q ( italic_x ) italic_d italic_x italic_d italic_t . (21)

    The Lippmann-Schwinger-Lanczos method is to replace u(j)superscript𝑢𝑗u^{(j)}italic_u start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT in the above by its data generated approximation 𝐮(j)superscript𝐮𝑗\vec{\bf{u}}^{(j)}over→ start_ARG bold_u end_ARG start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT from (4). Let 𝐮(j)(x,t)superscript𝐮𝑗𝑥𝑡{\bf u}^{(j)}(x,t)bold_u start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT ( italic_x , italic_t ) be a time interpolant of 𝐮(j)superscript𝐮𝑗\vec{\bf{u}}^{(j)}over→ start_ARG bold_u end_ARG start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT, and we solve

    F0jj(kτ)Fjj(kτ)=0kτΩw(j),0(x,kτt)𝐮(j)(x,t)q(x)𝑑x𝑑t,subscriptsuperscript𝐹𝑗𝑗0𝑘𝜏superscript𝐹𝑗𝑗𝑘𝜏superscriptsubscript0𝑘𝜏subscriptΩsuperscript𝑤𝑗0𝑥𝑘𝜏𝑡superscript𝐮𝑗𝑥𝑡𝑞𝑥differential-d𝑥differential-d𝑡F^{jj}_{0}(k\tau)-F^{jj}(k\tau)=\int_{0}^{k\tau}\int_{\Omega}w^{(j),0}(x,k\tau% -t){\bf{u}}^{(j)}(x,t)q(x)dxdt,italic_F start_POSTSUPERSCRIPT italic_j italic_j end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_k italic_τ ) - italic_F start_POSTSUPERSCRIPT italic_j italic_j end_POSTSUPERSCRIPT ( italic_k italic_τ ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k italic_τ end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT italic_w start_POSTSUPERSCRIPT ( italic_j ) , 0 end_POSTSUPERSCRIPT ( italic_x , italic_k italic_τ - italic_t ) bold_u start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT ( italic_x , italic_t ) italic_q ( italic_x ) italic_d italic_x italic_d italic_t , (22)

    for j=1,,K𝑗1𝐾j=1,\ldots,Kitalic_j = 1 , … , italic_K , k=0,,n1𝑘0𝑛1k=0,\ldots,n-1italic_k = 0 , … , italic_n - 1, yielding nm𝑛𝑚nmitalic_n italic_m equations to be inverted to find 𝐪𝐪{\bf q}bold_q, an approximation to q𝑞qitalic_q.

4 Step 2: Data completion and its mass matrix

In this step we assume we are given an approximation to the internal snapshots 𝐮(j)(x,t)superscript𝐮𝑗𝑥𝑡{\bf u}^{(j)}(x,t)bold_u start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT ( italic_x , italic_t ) for k=0,n1𝑘0𝑛1k=0,\ldots n-1italic_k = 0 , … italic_n - 1, j=1,K𝑗1𝐾j=1,\ldots Kitalic_j = 1 , … italic_K, along with an approximation 𝐪𝐪\bf{q}bold_q to the unknown q𝑞qitalic_q. We now use (12) to approximate the off diagonals of F𝐹Fitalic_F.

  1. 1.

    Data lifting: For ij𝑖𝑗i\neq jitalic_i ≠ italic_j, compute approximations 𝐅ijsuperscript𝐅𝑖𝑗{\bf F}^{ij}bold_F start_POSTSUPERSCRIPT italic_i italic_j end_POSTSUPERSCRIPT to Fijsuperscript𝐹𝑖𝑗F^{ij}italic_F start_POSTSUPERSCRIPT italic_i italic_j end_POSTSUPERSCRIPT

    𝐅ij(kτ):=F0ij(kτ)0kτΩw(j),0(x,kτt)𝐮(i)(x,t)𝐪(x)𝑑x𝑑t.assignsuperscript𝐅𝑖𝑗𝑘𝜏subscriptsuperscript𝐹𝑖𝑗0𝑘𝜏superscriptsubscript0𝑘𝜏subscriptΩsuperscript𝑤𝑗0𝑥𝑘𝜏𝑡superscript𝐮𝑖𝑥𝑡𝐪𝑥differential-d𝑥differential-d𝑡{\bf F}^{ij}(k\tau):=F^{ij}_{0}(k\tau)-\int_{0}^{k\tau}\int_{\Omega}w^{(j),0}(% x,k\tau-t){\bf u}^{(i)}(x,t){\bf q}(x)dxdt.bold_F start_POSTSUPERSCRIPT italic_i italic_j end_POSTSUPERSCRIPT ( italic_k italic_τ ) := italic_F start_POSTSUPERSCRIPT italic_i italic_j end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_k italic_τ ) - ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k italic_τ end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT italic_w start_POSTSUPERSCRIPT ( italic_j ) , 0 end_POSTSUPERSCRIPT ( italic_x , italic_k italic_τ - italic_t ) bold_u start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_x , italic_t ) bold_q ( italic_x ) italic_d italic_x italic_d italic_t . (23)

    Define the diagonals

    𝐅ii=Fiisuperscript𝐅𝑖𝑖superscript𝐹𝑖𝑖{\bf F}^{ii}=F^{ii}bold_F start_POSTSUPERSCRIPT italic_i italic_i end_POSTSUPERSCRIPT = italic_F start_POSTSUPERSCRIPT italic_i italic_i end_POSTSUPERSCRIPT (24)

    by the actual data values. What we have now is the lifted data

    {𝐅ij(kτ)} for k=0,,n1i,j=1,K.formulae-sequencesuperscript𝐅𝑖𝑗𝑘𝜏 for 𝑘0𝑛1𝑖𝑗1𝐾\{{\bf F}^{ij}(k\tau)\}\ \mbox{ for }\ k=0,\dots,n-1\ \ i,j=1,\ldots K.{ bold_F start_POSTSUPERSCRIPT italic_i italic_j end_POSTSUPERSCRIPT ( italic_k italic_τ ) } for italic_k = 0 , … , italic_n - 1 italic_i , italic_j = 1 , … italic_K .

    Note that the number of time steps is reduced by half compared to the original data (or the data from the previous lifting iterate).

  2. 2.

    MIMO mass matrix: Since 𝐅𝐅{\bf F}bold_F is now a full K×K𝐾𝐾K\times Kitalic_K × italic_K matrix, we use the block version of (16) (see for example [8]) to compute the MIMO mass matrix

    𝐌kl=12(𝐅((kl)τ)+𝐅((k+l)τ))subscript𝐌𝑘𝑙12𝐅𝑘𝑙𝜏𝐅𝑘𝑙𝜏{\bf M}_{kl}=\frac{1}{2}\left({\bf F}((k-l)\tau)+{\bf F}((k+l)\tau)\right)bold_M start_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( bold_F ( ( italic_k - italic_l ) italic_τ ) + bold_F ( ( italic_k + italic_l ) italic_τ ) ) (25)

    for k,l=0,,(n1)/2formulae-sequence𝑘𝑙0𝑛12k,l=0,\ldots,\lfloor(n-1)/2\rflooritalic_k , italic_l = 0 , … , ⌊ ( italic_n - 1 ) / 2 ⌋. Here 𝐌klsubscript𝐌𝑘𝑙{\bf M}_{kl}bold_M start_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT represents the kl𝑘𝑙klitalic_k italic_lth K×K𝐾𝐾K\times Kitalic_K × italic_K block of 𝐌𝐌{\bf M}bold_M.

  3. 3.

    Regularization of 𝐌𝐌{\bf M}bold_M: Since 𝐌𝐌{\bf M}bold_M is not a true mass matrix, it will likely not be symmetric positive definite, so regularization is required. We regularize by taking the symmetric part and thresholding: all eigenvalues below a chosen small ϵ0>0subscriptitalic-ϵ00\epsilon_{0}>0italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT > 0 are set to ϵ0subscriptitalic-ϵ0\epsilon_{0}italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. We note the result here is not very sensitive to the choice of ϵ0subscriptitalic-ϵ0\epsilon_{0}italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, provided that ϵ0subscriptitalic-ϵ0\epsilon_{0}italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is smaller than the minimum positive eigenvalue λmin+subscriptsuperscript𝜆𝑚𝑖𝑛\lambda^{+}_{min}italic_λ start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m italic_i italic_n end_POSTSUBSCRIPT of 𝐌𝐌{\bf M}bold_M. In our experiments we pick ϵ0=1012λmax+λmin+subscriptitalic-ϵ0superscript1012subscriptsuperscript𝜆𝑚𝑎𝑥subscriptsuperscript𝜆𝑚𝑖𝑛\epsilon_{0}=\sqrt{10^{-12}\lambda^{+}_{max}\lambda^{+}_{min}}italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = square-root start_ARG 10 start_POSTSUPERSCRIPT - 12 end_POSTSUPERSCRIPT italic_λ start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT italic_λ start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m italic_i italic_n end_POSTSUBSCRIPT end_ARG where λmax+subscriptsuperscript𝜆𝑚𝑎𝑥\lambda^{+}_{max}italic_λ start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT is the maximum positive eigenvalue of 𝐌𝐌{\bf M}bold_M. We note that this ϵ0subscriptitalic-ϵ0\epsilon_{0}italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT corresponds to the geometric mean of λmin+subscriptsuperscript𝜆𝑚𝑖𝑛\lambda^{+}_{min}italic_λ start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m italic_i italic_n end_POSTSUBSCRIPT and the minimum value 1012λmax+superscript1012subscriptsuperscript𝜆𝑚𝑎𝑥10^{-12}\lambda^{+}_{max}10 start_POSTSUPERSCRIPT - 12 end_POSTSUPERSCRIPT italic_λ start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT for stable linear algebraic operations with modified 𝐌𝐌{\bf M}bold_M in machine double precision. We abuse notation and continue to denote this regularized mass matrix by 𝐌𝐌{\bf M}bold_M.

5 Step 3: The MIMO LSL algorithm for completed data

In this step we assume we have a positive definite mass matrix 𝐌𝐌{\bf M}bold_M computed from a full 𝐅𝐅{\bf F}bold_F, where 𝐌𝐌{\bf M}bold_M is corresponding to time steps k=0,,N𝑘0𝑁k=0,\ldots,Nitalic_k = 0 , … , italic_N. The first time that we perform this step we will have N=(n1)/2𝑁𝑛12N=\lfloor(n-1)/2\rflooritalic_N = ⌊ ( italic_n - 1 ) / 2 ⌋; subsequent steps will be halved in size.

  1. 1.

    Orthogonalize the snapshots: We do a block Cholesky (with K×K𝐾𝐾K\times Kitalic_K × italic_K blocks) decomposition,

    𝐌=UU𝐌superscript𝑈top𝑈{\bf M}=U^{\top}Ubold_M = italic_U start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT italic_U

    where U𝑈Uitalic_U is block upper triangular. We note that there is some ambiguity in the choice of the blocks; we choose them so that all resulting orthogonalized functions

    v=uU1.𝑣𝑢superscript𝑈1\vec{v}=\vec{u}U^{-1}.over→ start_ARG italic_v end_ARG = over→ start_ARG italic_u end_ARG italic_U start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT .

    are all mutually orthogonal.

  2. 2.

    Repeat for the background medium: Compute the background MIMO mass matrix corresponding to the same time steps, its Cholesky decomposition

    M0=(U0)U0,superscript𝑀0superscriptsuperscript𝑈0topsuperscript𝑈0M^{0}=(U^{0})^{\top}U^{0},italic_M start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = ( italic_U start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT italic_U start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ,

    and the orthogonalized background snapshots

    v0=u0(U0)1.superscript𝑣0superscript𝑢0superscriptsuperscript𝑈01\vec{v}^{0}=\vec{u}^{0}(U^{0})^{-1}.over→ start_ARG italic_v end_ARG start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = over→ start_ARG italic_u end_ARG start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ( italic_U start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT .
  3. 3.

    Compute new data generated internal snapshots:

    𝐮=u0(U0)1U.𝐮superscript𝑢0superscriptsuperscript𝑈01𝑈\vec{\bf{u}}=\vec{u}^{0}(U^{0})^{-1}U.over→ start_ARG bold_u end_ARG = over→ start_ARG italic_u end_ARG start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ( italic_U start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_U .

    An important point is that these internal fields have been improved substantially compared to those generated in Step 1.

  4. 4.

    Invert the Lippman-Schwinger equation: Let 𝐮(j)(x,t)superscript𝐮𝑗𝑥𝑡{\bf u}^{(j)}(x,t)bold_u start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT ( italic_x , italic_t ) be a time interpolant of the fields in 𝐮𝐮\vec{\bf{u}}over→ start_ARG bold_u end_ARG corresponding to the j𝑗jitalic_jth source, and solve the Lippmann-Schwinger equation

    F0jj(kτ)Fjj(kτ)=0kτΩw(j),0(x,kτt)𝐮(j)(x,t)q(x)𝑑x𝑑tsubscriptsuperscript𝐹𝑗𝑗0𝑘𝜏superscript𝐹𝑗𝑗𝑘𝜏superscriptsubscript0𝑘𝜏subscriptΩsuperscript𝑤𝑗0𝑥𝑘𝜏𝑡superscript𝐮𝑗𝑥𝑡𝑞𝑥differential-d𝑥differential-d𝑡F^{jj}_{0}(k\tau)-F^{jj}(k\tau)=\int_{0}^{k\tau}\int_{\Omega}w^{(j),0}(x,k\tau% -t){\bf{u}}^{(j)}(x,t)q(x)dxdtitalic_F start_POSTSUPERSCRIPT italic_j italic_j end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_k italic_τ ) - italic_F start_POSTSUPERSCRIPT italic_j italic_j end_POSTSUPERSCRIPT ( italic_k italic_τ ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k italic_τ end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT italic_w start_POSTSUPERSCRIPT ( italic_j ) , 0 end_POSTSUPERSCRIPT ( italic_x , italic_k italic_τ - italic_t ) bold_u start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT ( italic_x , italic_t ) italic_q ( italic_x ) italic_d italic_x italic_d italic_t (26)

    to obtain a new approximation 𝐪𝐪{\bf q}bold_q to the unknown q𝑞qitalic_q. In this Lippmann-Schwinger step we use only the true SISO data, since our lifted off diagonals do not add any new information; they were used only to improve the internal fields.

6 Step 4: loop for iterations

We note that our experiments suggest that whether or not we need to proceed with iterations and return to step 2 depends on the number of targets to resolve. Our observations show that the number of iterations should be one less than the maximum number of objects with internal multiples. In particular, for the problems with two targets we will consider section 7, no loop is needed. However, there is benefit from having an extra iteration for the three-objects problem (see our last numerical example). We can vaguely explain this as follows: the Born solution (as well as LSL without completion) successfully resolve the closest target only, while the images of deeper ones are polluted by multiple echoes. Then, every iteration of LSL with completion cleans up these multiples from the deeper and deeper objects one by one. In this sense there is a similarity between our approach and the Inverse Born Series [24], however, a more precise connection needs to be explored further.

7 Numerical Experiments

Refer to caption
Figure 1: True model for imaging of two targets in homogeneous background. Red crosses show the locations of SAR sources/receivers.
Refer to caption
Figure 2: While the background solution (top middle) doesn’t capture any reflections of the true solution (top left), the LSL solution without data completion captures them in spherical averages only (top right). The data completion step improved the internal solution reconstruction (bottom left) significantly. The second iteration (bottom right) sharpened the profile, though all of the reflections were already captured during the first iteration.
True Born
Refer to caption Refer to caption
LSL w/o completion LSL with completion
Refer to caption Refer to caption
2nd iteration of data completion fo LSL
Refer to caption
Figure 3: 2D example. LSL solution without data completion (bottom left) improves upon the Born solution (top right) . One step of LSL data completion (bottom right) allowed us to improve the image significantly. The second iteration didn’t improve the image for this simple model.
True Born
Refer to caption Refer to caption
LSL w/o completion LSL with completion
Refer to caption Refer to caption
LSL with completion for noisy data
Refer to caption
Figure 4: 2D example: Red crosses in the true model (top left) show the locations of SAR sources/receivers. LSL solution without data completion (middle left) improves upon the Born solution (top right) . One step of LSL data completion (middle right) allowed us to resolve the hidden object. Adding 5% noise didn’t worsen the image (bottom) significantly.
True Born
Refer to caption Refer to caption
LSL w/o completion LSL with completion
Refer to caption Refer to caption
Figure 5: 2.5D example: Red crosses in the true model (top left) show the locations of SAR sources/receivers. LSL solution without data completion (bottom left) , Born solution (top right). One step of LSL data completion (bottom right) yields a dramatic improvement in the image.

In our first numerical experiment, we considered the imaging of two elongated targets in a homogeneous background (see Fig. 1). To obtain synthetic data, we discretized 4 in Ω=[0;200]×[0;100]Ω02000100\Omega=[0;200]\times[0;100]roman_Ω = [ 0 ; 200 ] × [ 0 ; 100 ] using finite-differences on a 400×200400200400\times 200400 × 200 2D grid. The obtained discrete problem was solved for 27 positions of radar that were emitting a non-modulated Gaussian pulse (radar positions are marked by crosses in Fig. 1). To solve the Lippmann-Schwinger equation, we discretized it using quadrature on a 200×100200100200\times 100200 × 100 grid. First, we examine how data completion improves the accuracy of the internal fields. In Fig. 2 we see the true internal solution (top left) and the background solution (top middle), which is used for the Born approximation in (12). The ROM-based SISO internal solution without data completion is shown in the top right. While the background solution doesn’t have any reflections, the internal solution we obtained without data completion captures the reflections in the sense of spherical averages. The internal solution computed after data completion is shown in the bottom left. As one can observe, MIMO data completion improves the accuracy of the internal solution reconstruction significantly. We also note that the second iteration made the profile sharper (see bottom right), however, all of the multiples were already captured during the first iteration.

In Figure 3 we compare reconstructions for this first example, where we see the Born image (top right) and LSL images without and with data completion (middle left and right respectively). We note that the ROM step had to be performed in a structure-preserving way in order to obtain a stable internal solution. The data completion for this first example results in a significantly sharper image of two targets, with almost no ghost reflections. As was expected, the second iteration (bottom plot) didn’t yield any improvement for this two target model. We note that in each of the reconstructions we chose the SVD truncation level which was best for each separately, which is why the frequency of the artifacts appears higher for the LSL with completion. In all cases the artifacts seen in these pictures could possibly be further diminished by using sparse or other regularization techniques.

For our second example we considered another two dimensional target (see top left plot in Figure 4) with 60 sources and receivers located at the top and emitting a non-modulated Gaussian pulse. We used an 800×300800300800\times 300800 × 300 finite-difference grid for discretization of the forward problem in Ω=[0;400]×[0;150]Ω04000150\Omega=[0;400]\times[0;150]roman_Ω = [ 0 ; 400 ] × [ 0 ; 150 ] to generate synthetic data. Then, data-driven internal solutions were obtained and we solved the Lippmann–Schwinger equation using the quadrature on a 400×150400150400\times 150400 × 150 grid. Similar to the previous example, the data completion (middle right) really sharpens the image compared to Born (top right) and LSL without completion (middle left), and in this case allows us to resolve the image inside of the box. In this experiment we added the result with 5% relative random noise (see bottom plot) to show a low sensitivity of the approach to the measurement errors.

For our next example, we considered a 2.5D problem, where our waves are fully three dimensional but the medium is assumed to be homogeneous in the transversal direction (see top left in Fig. 5). The forward problem was discretized on a 200×200×200200200200200\times 200\times 200200 × 200 × 200 grid in a 3D domain Ω=[0;100]3Ωsuperscript01003\Omega=[0;100]^{3}roman_Ω = [ 0 ; 100 ] start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT. Similar to the first example, we used 27 sources/receivers with a non-modulated Gaussian pulse and solved the discrete problems for them to obtain synthetic data. We then computed the ROM-based approximants of the internal solutions and plugged them in into the Lippmann-Schwinger equation, which was then solved for a 2D image via quadratures on a 100×100100100100\times 100100 × 100 grid. We see the reconstructions in Figure 5. Born (top right) and LSL without completion (bottom left) both show significant ghost reflections, but these are greatly reduced with one step of LSL with data completion (bottom right).

For all of the examples above, adding an extra iteration of data completion in LSL will not improve the image. Indeed, as we noted before, iterations for LSL with data completion are beneficial when there are more than two objects that produce internal multiples. To illustrate this, we considered the 2D problem with 3 objects shown in Fig. 6. Our modeling and inversion setups were precisely the same as in the second example, i.e. an 800×300800300800\times 300800 × 300 grid in Ω=[0;400]×[0;150]Ω04000150\Omega=[0;400]\times[0;150]roman_Ω = [ 0 ; 400 ] × [ 0 ; 150 ], non-modulated Gaussian pulses emitted by 60 sources/receivers and quadrature for the Lippmann-Schwinger equation on a 400×150400150400\times 150400 × 150 grid. In Fig. 7 we show how the extra iteration (bottom right) of LSL with data completion improved the internal solution (bottom left), though both are significantly closer to the exact solution (top left) than Born (top middle) and LSL without completion (top right). This improved internal solution helped to reveal the deepest object in the model (see bottom right plot in Fig.8), which was neither visible in the Born image (top left), nor in the LSL images without completion (top right) and with one iteration of completion (bottom left).

Refer to caption
Figure 6: True model for imaging of three targets in homogeneous background. Red crosses show the locations of SAR sources/receivers.
Refer to caption
Figure 7: Similar to the problem with 2 objects, the background solution (top middle) doesn’t capture any reflections of the true solution (top left) while the LSL solution without data completion captures them in spherical averages only (top right). The data completion step improved the internal solution reconstruction (bottom left) significantly, and the second iteration corrected for the missing internal multiples even further (bottom right).
Born LSL w/o completion
Refer to caption Refer to caption
1 iteration of LSL with completion 2 iterations of LSL with completion
Refer to caption Refer to caption
Figure 8: 2D example with 3 objects: LSL solution without data completion (top right) improves upon the Born solution (top left), however the image is still polluted by multiple echoes. One step of LSL data completion (bottom left) allowed us to improve the image significantly, but the deepest object only became visible after 1 extra iteration of data completion in LSL (bottom right).

8 Conclusions

In the ROM based Lippmann-Schwinger-Lanczos method, one constructs data driven internal solutions, and uses them in the Lippmann-Schwinger integral. When the given data is monostatic, as is the case for SAR, these internal solutions are only accurate up to spherical averages. We saw here that adding the step of ROM based data completion can dramatically improve the accuracy of the internal solutions, leading to sharpened reconstructions. Moreover, additional iterations of data completion in LSL can improve images further in the case of complex models with more than two objects producing internal multiples.

Many practically important problems have more complicated data matrix patterns than shown in (1-2). For example, in the case of tomographic ultrasound, the data acquisition may be block diagonal. The method we described in this work can be extended many of those cases as well [25]. The LSL approach can also be extended in some cases to variable wave speed problems, in particular for small slow speed perturbations [1], or more generally by combining with iteration [8]. We plan to investigate extension of this data-completion approach to these and more general setups in future work.

Acknowledgments The authors are grateful to Liliana Borcea, Alex Mamonov and Jörn Zimmerling for productive discussions that inspired this research. V. Druskin was partially supported by AFOSR grants FA 955020-1-0079, FA9550-20-1-0079, and NSF grant DMS-2110773. M. Zaslavsky was partially supported by AFOSR grant FA9550-20-1-0079. S. Moskow was partially supported by NSF grants DMS-2008441 and DMS-2308200.

References

  • [1] J. Baker, E. Cherkaev, V. Druskin, S. Moskow, and M. Zaslavsky. Doubly regularized lippmann-schwinger-lanczos algorithm for inverse scattering problems in the frequency domain. submitted.
  • [2] L. Borcea, V. Druskin, A. Mamonov, M. Zaslavsky, and J. Zimmerling. Reduced order model approach to inverse scattering. SIAM Journal on Imaging Sciences, 13(2):685–723, 2020.
  • [3] Liliana Borcea, Vladimir Druskin, Fernando Guevara Vasquez, and Alexander V. Mamonov. Resistor network approaches to electrical impedance tomography. Inverse Problems and Applications: Inside Out II, Math. Sci. Res. Inst. Publ, 60:55–118, 2011.
  • [4] Liliana Borcea, Vladimir Druskin, Alexander V. Mamonov, Shari Moskow, and Mikhail Zaslavsky. Reduced order models for spectral domain inversion: embedding into the continuous problem and generation of internal data. Inverse Problems, 36(5), 2020.
  • [5] Liliana Borcea, Vladimir Druskin, Alexander V. Mamonov, and Mikhail Zaslavsky. A model reduction approach to numerical inversion for a parabolic partial differential equation. Inverse Problems, 30(12):125011, 2014.
  • [6] Liliana Borcea, Vladimir Druskin, Alexander V Mamonov, and Mikhail Zaslavsky. Untangling nonlinearity in inverse scattering with data-driven reduced order models. Inverse Problems, 34(6):065008, 2018.
  • [7] Liliana Borcea, Vladimir Druskin, Alexander V. Mamonov, and Mikhail Zaslavsky. Robust nonlinear processing of active array data in inverse scattering via truncated reduced order models. Journal of Computational Physics, 381:1–26, 2019.
  • [8] Liliana Borcea, Josselin Garnier, Alexander V. Mamonov, and Jörn Zimmerling. Waveform inversion with a data driven estimate of the internal wave. SIAM Journal on Imaging Sciences, 16(1):280–312, 2023.
  • [9] Liliana Borcea, Josselin Garnier, Alexander V. Mamonov, and Jörn T. Zimmerling. Reduced order model approach for imaging with waves. ArXiv, abs/2108.01609, 2021.
  • [10] Matthew J. Burfeindt and Hatim F. Alqadah. Phase-encoded linear sampling method imaging of 3d targets from circular synthetic aperture data. In 2024 IEEE Radar Conference (RadarConf24), pages 1–6, 2024.
  • [11] Fioralba Cakoni, David Colton, and Peter Monk. The Linear Sampling Method in Inverse Electromagnetic Scattering. Society for Industrial and Applied Mathematics, 2011.
  • [12] Margaret Cheney, Matthew J Burfeindt, and NAVAL RESEARCH LAB WASHINGTON DCColorado State University. A procedure for suppressing multiple scattering. 2021.
  • [13] V. Druskin, A.V. Mamonov, and M. Zaslavsky. A nonlinear method for imaging with acoustic waves via reduced order model backprojection. SIAM Journal on Imaging Sciences, 11(1):164–196, 2018.
  • [14] Vladimir Druskin, Alexander V. Mamonov, Andrew E. Thaler, and Mikhail Zaslavsky. Direct, nonlinear inversion algorithm for hyperbolic problems via projection-based model reduction. SIAM Journal on Imaging Sciences, 9(2):684–747, 2016.
  • [15] Vladimir Druskin, Shari Moskow, and Mikhail Zaslavsky. Lippmann-schwinger-lanczos algorithm for inverse scattering problems. Inverse Problems, apr 2021.
  • [16] Vladimir Druskin, Shari Moskow, and Mikhail Zaslavsky. On extension of the data driven rom inverse scattering framework to partially nonreciprocal arrays. Inverse Problems, 8(38), 2022.
  • [17] Vladimir Druskin, Shari Moskow, and Mikhail Zaslavsky. Reduced order modeling inversion of monostatic data in a multi-scattering environment. SIAM Journal on Imaging Sciences, 17(1), 2024.
  • [18] Mikhail Gilman and Semyon Tsynkov. A mathematical model for sar imaging beyond the first born approximation. SIAM Journal on Imaging Sciences, 8(1):186–225, 2015.
  • [19] Paul Kepley, Lauri Oksanen, and Maarten V. de Hoop. Generating virtual interior point source traveltimes and redatuming using boundary control. Seg Technical Program Expanded Abstracts, 2016.
  • [20] Jerry T. Kim, Eric L. Mokole, and Margaret Cheney. Concept of iterative time-reversal radar (itrr). In 2023 IEEE Conference on Antenna Measurements and Applications (CAMA), pages 411–416, 2023.
  • [21] Michael V. Klibanov, Alexey V. Smirnov, Vo A. Khoa, Anders J. Sullivan, and Lam H. Nguyen. Through-the-wall nonlinear sar imaging. IEEE Transactions on Geoscience and Remote Sensing, 59(9):7475–7486, 2021.
  • [22] John Lagergren, Kevin Flores, Mikhail Gilman, and Semyon Tsynkov. Deep learning approach to the detection of scattering delay in radar images. Journal of Statistical Theory and Practice, 15(1):14, 2021.
  • [23] Alison E. Malcolm, Maarten V. de Hoop, and Henri Calandra. Identification of image artifacts from internal multiples. Geophysics, 72, 2007.
  • [24] Shari Moskow and John C. Schotland. 12. Inverse Born series. The Radon Transform: The First 100 Years and Beyond, edited by Ronny Ramlau and Otmar Scherzer, pages 273–296. De Gruyter, Berlin, Boston, 2019.
  • [25] J. D. Roberts. Linear model reduction and solution of the algebraic Riccati equation by use of the sign function. Internat. J. Control, 32(4):677–687, 1980. First issued as report CUED/B-Control/TR13, Department of Engineering, University of Cambridge, 1971.
  • [26] Jean Virieux. An introduction to full waveform inversion. 2016.
  • [27] Francis Watson, Daniel Andre, and William R B Lionheart. Resolving full-wave through-wall transmission effects in multi-static synthetic aperture radar. Inverse Problems, 2024.