Publication Cover
Applicable Analysis
An International Journal
Volume 103, 2024 - Issue 6
358
Views
1
CrossRef citations to date
0
Altmetric
Research Article

An inverse source problem for linearly anisotropic radiative sources in absorbing and scattering medium

&
Pages 1149-1164 | Received 10 Dec 2022, Accepted 01 Jul 2023, Published online: 12 Jul 2023

Abstract

We consider in a two dimensional absorbing and scattering medium, an inverse source problem in the stationary radiative transport, where the source is linearly anisotropic. The medium has an anisotropic scattering property that is neither negligible nor large enough for the diffusion approximation to hold. The attenuating and scattering properties of the medium are assumed known. For scattering kernels of finite Fourier content in the angular variable, we show how to recover the anisotropic radiative sources from boundary measurements. The approach is based on the Cauchy problem for a Beltrami-like equation associated with A-analytic maps. As an application, we determine necessary and sufficient conditions for the data coming from two different sources to be mistaken for each other.

2010 Mathematics Subject Classifications:

1. Introduction

In this work, we consider an inverse source problem for stationary radiative transfer (transport) [Citation1,Citation2], in a two-dimensional bounded, strictly convex domain ΩR2, with boundary Γ. The stationary radiative transport models the linear transport of particles through a medium and includes absorption and scattering phenomena. In the steady state case, when generated solely by a linearly anisotropic source f inside Ω, the density u(z,θ) of particles at z traveling in the direction θ solves the stationary radiative transport boundary value problem (1) θu(z,θ)+a(z,θ)u(z,θ)S1k(z,θ,θ)u(z,θ)dθ=f(z,θ),(z,θ)Ω×S1,u|Γ=0.(1) In boundary value problem (Equation1), the function a(z,θ) is the medium capability of absorption per unit path-length at z moving in the direction θ called the attenuation coefficient, the function k(z,θ,θ) is the scattering coefficient which accounts for particles from an arbitrary direction θ which scatter in the direction θ at a point z, and Γ:={(ζ,θ)Γ×S1:ν(ζ)θ<0} is the incoming unit tangent sub-bundle of the boundary, with ν(ζ) being the outer unit normal at ζΓ. The attenuation and scattering coefficients are assumed known real valued functions. The boundary condition in (Equation1) indicates that no radiation is coming from outside the domain. Throughout, the measure dθ on the unit sphere S1 is normalized to S1dθ=1.

The (forward) boundary value problem (Equation1) is known to be well-posed under various assumptions, e.g in [Citation3–6], with a general result in [Citation7] showing that, for an open and dense set of coefficients aC2(Ω¯×S1) and kC2(Ω¯×S1×S1), the boundary value problem (Equation1) has a unique solution uL2(Ω×S1) for any fL2(Ω×S1). In [Citation8], it is shown that for attenuation merely once differentiable, aC1(Ω¯×S1) and kC2(Ω¯×S1×S1), the boundary value problem (Equation1) has a unique solution uLp(Ω×S1) for any fLp(Ω×S1), p>1. Moreover, uniqueness result of the forward problem (Equation1) are also establish in weighted Lp spaces in [Citation9], and in [Citation10,Citation11] using Carleman estimates.

In our reconstruction method here, some of our arguments require solutions uC2,μ(Ω¯×S1), 12<μ<1. We revisit the arguments in [Citation7,Citation8] and show that such a regularity can be achieved for sources fW3,p(Ω×S1), p>4; see Theorem A.2 (iii) below.

For a given medium, i.e. a and k both known, we consider the inverse problem of determining the linear anisotropic source f=f0+θF; in particular, recovering the isotropic scalar field f0 and the vector field F from measurements gf0,F of exiting radiation on Γ, (2) u|Γ+=gf0,F,(2) where Γ+:={(z,θ)Γ×S1:ν(z)θ>0} is the outgoing unit tangent sub-bundle of the boundary.

For anisotropic sources the problem has non-uniqueness [Citation4,Citation12–14]. One of our main result, Theorem 3.1 shows that from boundary measurement data gf0,F, one can only recover the part of the linear anisotropic source f=f0+θF; in particular, only the solenoidal part Fs of the vector field F is recovered inside the domain. However, in Theorem 3.2, if one know apriori that the source F is divergence-free, then from the data gf0,F, one can recover both isotropic field f0 and the vector field F inside the domain. Moreover, instead of apriori information of the divergence-free source F, if one has the additional data gf0,0 information along with the data gf0,F, then in Theorem 3.3, one can recover both sources f0 and F under subcritical assumption of the medium. One of the main crux in our reconstruction method is the observation that any finite Fourier content in the angular variable of the scattering kernel splits the problem into an infinite system of non-scattering case and a boundary value problem for a finite elliptic system. The role of the finite Fourier content has been independently recognized in [Citation15,Citation16].

The inverse source problem above has applications in medical imaging: In a non-scattering (k = 0) and non-attenuating (a = 0) medium the problem is mathematically equivalent to the one occurring in classical computerized X-ray tomography (e.g. [Citation17,Citation18]). In the absorbing non-scattering medium, such a problem (with only isotropic source f=f0), appears in Positron/Single Photon Emission Tomography [Citation18,Citation19], and f=θF with f0=0, appears in Doppler Tomography [Citation18–20]. For applications in scattering media the inverse source problem formulated here is the two dimensional version of the corresponding three dimensional problem occurring in imaging techniques such as Bioluminescence tomography and Optical Molecular Imaging, see [Citation21–23] and references therein.

In this work, except for the results in the appendix, the attenuation coefficient are assumed isotropic a=a(z), and that the scattering kernel k(z,θ,θ)=k(z,θθ) depends polynomially on the angle between the directions. Moreover, the functions a, k and the source f are assumed real valued.

In Section 2, we recall some basic properties of A-analytic theory, and in Section 3 we provide the reconstruction method for the full (part) of the linearly anisotropic source. Our approach is based on the Cauchy problem for a Beltrami-like equation associated with A-analytic maps in the sense of Bukhgeim [Citation17]. The A-analytic approach developed in [Citation17] treats the non-attenuating case, and the absorbing but non-scattering case is treated in [Citation24]. The original idea of Bukhgeim from the absorbing non-scattering media [Citation17,Citation24] to the absorbing and scattering media has been extended in [Citation8,Citation15]. In here we extend the results in [Citation8,Citation15] to linear anisotropic sources.

In Section 4, the method used will explain when the data coming from two different linear anisotropic field sources can be mistaken for each other.

In the appendix, we revisit the arguments in [Citation7,Citation8] and remark on the existence and regularity of the forward boundary value problem. The results in the appendix consider both attenuation coefficient and scattering kernel in general setting.

2. Ingredients from A-analytic theory

In this section, we briefly introduce the properties of A-analytic maps needed later, and introduce notation. We recall some of the existing results and concepts used in our reconstruction method.

For 0<μ<1, p = 1, 2, we consider the Banach spaces: (3) Cμ(Γ;l1):={g=g0,g1,g2,:supξΓg(ξ)l1+supξ,ηΓξηg(ξ)g(η)l1|ξη|μ<},Yμ(Γ):={g:supξΓj=0j2|gj(ξ)|<,andsupξ,ηΓξηj=0j|gj(ξ)gj(η)||ξη|μ<},(3) where j=(1+|j|2)1/2. Similarly, we consider Cμ(Ω¯;l1), and Cμ(Ω¯;l).

For z=x1+ix2, we consider the Cauchy-Riemann operators ¯=(x1+ix2)/2,=(x1ix2)/2. A sequence valued map Ωzv(z):=v0(z),v1(z),v2(z), in C(Ω¯;l)C1(Ω;l) is called L2-analytic (in the sense of Bukhgeim [Citation17]), if (4) ¯v(z)+L2v(z)=0,zΩ,(4) where L is the left shift operator Lv0,v1,v2,=v1,v2,, and L2=LL.

Bukhgeim's original theory [Citation17] shows that solutions of (Equation4), satisfy a Cauchy-like integral formula, (5) v(z)=B[v|Γ](z),zΩ,(5) where B is the Bukhgeim-Cauchy operator acting on v|Γ. We use the formula in [Citation25], where B is defined component-wise for n0 by (6) (Bv)n(z):=12πiΓvn(ζ)ζzdζ+12πiΓ{ζzdζ¯ζ¯z¯}j=1vn2j(ζ)(ζ¯z¯ζz)j,zΩ.(6) Similar to the analytic maps, the traces of L2-analytic maps on the boundary must satisfy some constraints, which can be expressed in terms of a corresponding Hilbert-like transform introduced in [Citation26]. More precisely, the Bukhgeim-Hilbert transform H is defined component-wise for n0 by (7) (Hg)n(z)=1πΓgn(ζ)ζzdζ+1πΓ{ζzdζ¯ζ¯z¯}j=1gn2j(ζ)(ζ¯z¯ζz)j,zΓ,(7) and we refer to [Citation26] for its mapping properties.

Another ingredient, in addition to L2-analytic maps, consists in the one-to-one relation between solutions u:=u0,u1,u2, satisfying (8) ¯u+L2u+aLu=0,(8) and the L2-analytic map v=v0,v1,v2, satisfying (Equation4), via a special function h, see [Citation27, Lemma 4.2] for details. The function h is defined as h(z,θ):=0a(z+tθ)dt12(IiH)Ra(zθ,θ), where Ra(s,θ)=a(sθ+tθ)dt is the Radon transform of the attenuation a, and Hh(s)=1πh(t)stdt is the classical Hilbert transform [Citation28]. The function h has vanishing negative Fourier modes yielding the expansions eh(z,θ):=k=0αk(z)ei,eh(z,θ):=k=0βk(z)ei, for (z,θ)Ω¯×S1. Using the Fourier coefficients of e±h, define the sequence valued maps Ω¯zβ(z):=β0(z),β1(z),,Ω¯zα(z):=α0(z),α1(z),,,and define the operators e±G component-wise for each n0, by (9) (eGu)n=(αu)n=k=0αkunk,and(eGu)n=(βu)n=k=0βkunk.(9) Note the commutating property [e±G,L]=0.

Lemma 2.1

[Citation29, Lemma 4.2]

Let aC1,μ(Ω¯), μ>1/2, and e±G be operators as defined in (Equation9).

  1. If uC1(Ω,l1) solves ¯u+L2u+aLu=0, then v=eGuC1(Ω,l1) solves ¯v+L2v=0.

  2. Conversely, if vC1(Ω,l1) solves ¯v+L2v=0, then u=eGvC1(Ω,l1) solves ¯u+L2u+aLu=0.

3. Reconstruction of a sufficiently smooth linearly anisotropic source

For an isotropic real valued vector field F and real map f0, recall the boundary value problem (Equation1): (10) θu(z,θ)+a(z)u(z,θ)S1k(z,θθ)u(z,θ)dθ=f0(z)+θF(z)f(z,θ),(z,θ)Ω×S1,u|Γ=0,(10) with an isotropic attenuation a=a(z), and with the scattering kernel k(z,θ,θ)=k(z,θθ) depending polynomially on the angle between the directions, (11) k(z,cosθ)=k0(z)+2n=1Mkn(z)cos(),(11) for some fixed integer M1. Note that, since k(z,cosθ) is both real valued and even in θ, the coefficient kn is the (n)th Fourier coefficient of k(z,cos()). Moreover kn is real valued, and kn(z)=kn(z)=12πππk(z,cosθ)einθdθ.

For the real vector field F=F1,F2, let (12) f1:=12(F1+iF2),(12) and for θ=(cosθ,sinθ)S1, a calculation shows that the linear anisotropic source (13) f(z,θ)=f0(z)+θF(z)=f0(z)+f1(z)¯eiθ+f1(z)eiθ.(13) We assume that the coefficients a,k0,k1,,kMC3(Ω¯) are such that the forward problem (Equation10) has a unique solution uLp(Ω×S1) for any fLp(Ω×S1), p>1, see Theorem A.1. Moreover, we assume also an unknown source of a priori regularity fW3,p(Ω¯;R), p>4, and by Theorem A.2 part (iii), the solution uW3,p(Ω×S1),p>4. Furthermore, the functions a, k and source f are assumed real valued, so that the solution u is also real valued.

Let u(z,θ)=un(z)einθ be the formal Fourier series representation of the solution of (Equation10) in the angular variable θ=(cosθ,sinθ). Since u is real valued, the Fourier modes {un} occurs in complex-conjugate pairs un=un¯, and the angular dependence is completely determined by the sequence of its nonpositive Fourier modes (14) Ωzu(z):=u0(z),u1(z),u2(z),.(14) For the derivatives ,¯ in the spatial variable, the advection operator θ=eiθ¯+eiθ. By identifying the Fourier coefficients of the same order, Equation (Equation10) reduces to the system: (15) ¯u1(z)+u1(z)+[a(z)k0(z)]u0(z)=f0(z),(15) (16) ¯u0(z)+u2(z)+[a(z)k1(z)]u1(z)=f1(z),(16) (17) ¯un(z)+un2(z)+[a(z)kn1(z)]un1(z)=0,1nM1,(17) (18) ¯un(z)+un2(z)+a(z)un1(z)=0,nM,(18) where f1 as in (Equation12).

By Hodge decomposition [Citation13], any vector field F=F1,F2H1(Ω;R2) decomposes into a gradient field and a divergence-free (solenoidal) field : (19) F=φ+Fs,φ|∂Ω=0,divFs=0,(19) where φH02(Ω;R) and Fs=F1s,F2sHdiv1(Ω;R2):={FsH1(Ω;R2):divFs=0}.

Note that for f1 in (Equation12), we have (20) 4f1=divF+icurlF.(20) Using 4¯f1=Δf1, we have ΔF1=x1divFx2curlF, and ΔF2=x2divF+x1curlF. Moreover, for f1s=(F1s+iF2s)/2, the Hodge decomposition (Equation19) can be rewritten as (21) f1=¯φ+f1s,φ|∂Ω=0,Re(f1s)=0.(21) The following result show that from the knowledge of boundary data, one can only recover the part of the linear anisotropic source f; in particular, only the solenoidal part Fs of the vector field source F can be recovered inside Ω.

Theorem 3.1

Let ΩR2 be a strictly convex bounded domain, and Γ be its boundary. Consider the boundary value problem (Equation10) for some known real valued a,k0,k1,,kMC3(Ω¯) such that (Equation10) is well-posed. If scalar and vector field sources f0 and F are real valued, W3,p(Ω;R) and W3,p(Ω;R2)-regular, respectively, with p>4, then the data gf0,F defined in (Equation2), uniquely determine the solenoidal part Fs in Ω. Moreover, uu0 is also uniquely determined in Ω, where u is the solution of (Equation10) and u0 is the zeroth Fourier mode of u in the angular variable.

Proof.

Let u be the solution of the boundary value problem (Equation10) and let u=u0,u1,u2, be the sequence valued map of its non-positive Fourier modes. Since the scalar field f0W3,p(Ω;R),p>4, and isotropic vector field FW3,p(Ω;R2),p>4, then the anisotropic source f=f0+θF belong to W3,p(Ω×S1) for p>4. By applying Theorem A.2 (iii), we have uW3,p(Ω×S1),p>4. Moreover, by the Sobolev embedding [Citation30], W3,p(Ω×S1)C2,μ(Ω¯×S1) with μ=12p>12, we have uC2,μ(Ω¯×S1), and thus, by [Citation26, Proposition 4.1 (ii)], the sequence valued map uYμ(Γ).

Since FW3,p(Ω;R2),p>4, then by compact imbedding of Sobolev spaces [Citation30], FH1(Ω;R2). By Hodge decomposition (Equation19), field F=φ+Fs, with φ|Γ=0, and divFs=0.

We note from (Equation18) that the shifted sequence valued map LMu=uM,uM1,uM2, solves (22) ¯LMu(z)+L2LMu(z)+a(z)LM+1u(z)=0,zΩ.(22) Let v:=eGLMu. By Lemma 2.1, the system (Equation22) becomes ¯v+L2v=0, i.e v is L2 analytic.

By (Equation2), the data u|Γ+=g determines the sequence valued map LMu on Γ. By Proposition Equation9 (iii), and the convolution formula (Equation9), traces LMu|Γ determines the traces vYμ(Γ) on Γ.

Since v|Γ is the boundary value of an L2-analytic function in Ω, then [Citation26, Theorem 3.2 (i)] yields (23) [I+iH]v|Γ=0,(23) where H is the Bukhgeim-Hilbert transform in (Equation7).

From v on Γ, we use the Bukhgeim-Cauchy Integral formula (Equation6) to construct the sequence valued map v inside Ω: (24) v(z)=B[v|Γ](z),zΩ,(24) By [Citation29, Proposition 2.3] and [Citation26, Theorem 3.2 (ii)], the constructed sequence valued map vC1,μ(Ω;l1)Cμ(Ω¯;l1)C2(Ω;l) is L2-analytic in Ω.

From the convolution formula (Equation9), we construct the sequence valued map (25) LMu:=eGv.(25) Thus, determining un inside Ω for nM. In particular, we recover modes uM1,uMC2(Ω).

Recall that the modes u1,u2,,uM,uM1 satisfy (26a) ¯uM+j=uM+j2[(akM+j1)uM+j1],1jM1,(26a) (26b) uM+j|Γ=gM+j.(26b) By applying 4 to (Equation26a), the mode uM+1 (for j=1) is then the solution to the Dirichlet problem for the Poisson equation (27a) ΔuM+1=42uM14[(akM)uM],(27a) (27b) uM+1|Γ=gM+1,(27b) where the right hand side of (Equation27a) is known. We solve repeatedly (Equation27a) for j=2,,M1 in (Equation26a), to recover the modes (28) uM+1,uM+2,,u1,inΩ.(28) From determined LMu=uM,uM1,uM2, in (Equation25) and modes uM+1,uM+2,,u1 in (Equation28), the sequence Lu=u1,u2, is determined in Ω. Thus uu0 is determined in Ω.

Since u0,u1,u2C2(Ω), we can take 4 on both sides of Equation (Equation16) to get (29) Δu0+42u2+4([ak1]u1)=4f1=divF+icurlF,(29) where in the last equality we use (Equation20).

Moreover, since u0 is real valued and divF=Δφ, by equating the real part in (Equation29) yields the boundary value problem: (30a) Δ(u0φ)=4Re[2u2+([ak1]u1)],(30a) (30b) (u0φ)|Γ=g0,(30b) where the right hand side of (Equation30a) is known. Thus, real valued function (u0φ) is recovered in Ω, by solving the Dirichlet problem for the above Poisson Equation (Equation30a).

Even though u0 is not determined, the function (u0φ) is uniquely determined in Ω. Moreover, modes u1 and u2 are also uniquely determined in Ω. Furthermore, using expression of f1 from (Equation16) and f1s=f1¯φ from (Equation21), we define (31) f1s:=¯(u0(z)φ(z))+u2(z)+[a(z)k1(z)]u1(z),zΩ,(31) with f1s satisfying Re(f1s)=0.

Thus, the solenoidal part Fs=2Ref1s,2Imf1s, of the vector field F is recovered in Ω.

If we know apriori that the vector field F is incompressible (i.e divergenceless), then we can reconstruct both scalar field source f0 and vector field source F in Ω.

Theorem 3.2

Let ΩR2 be a strictly convex bounded domain, and Γ be its boundary. Consider the boundary value problem (Equation10) for some known real valued a,k0,k1,,kMC3(Ω¯) such that (Equation10) is well-posed. If the unknown scalar field source f0 and divergenceless vector field sources F are real valued, W3,p(Ω;R) and W3,p(Ω;R2)-regular, respectively, with p>4, then the data gf0,F defined in (Equation2) uniquely determine both f0 and F in Ω.

Proof.

Let u be the solution of the boundary value problem (Equation10) and let u=u0,u1,u2, be the sequence valued map of its non-positive Fourier modes, Since the isotropic scalar and vector field f0W3,p(Ω;R), and FW3,p(Ω;R2) respectively for p>4, then the anistropic source f=f0+θFW3,p(Ω×S1) and by applying Theorem A.2 (iii), we have uW3,p(Ω×S1). By the Sobolev embedding [Citation30], W3,p(Ω×S1)C2,μ(Ω¯×S1) with μ=12p>12, we have uC2,μ(Ω¯×S1), and thus, by [Citation26, Proposition 4.1 (ii)], uYμ(Γ).

Since FW3,p(Ω;R2),p>4, then by compact imbedding of Sobolev spaces [Citation30], FH1(Ω;R2). By Hodge decomposition (Equation19), field F=φ+Fs, with φ|Γ=0, and divFs=0.

If we know apriori that the vector field F is incompressible (i.e divergenceless F=0). Then φ=divF=0 and φ|∂Ω=0 implies φ0 inside Ω. Thus, vector field F=Fs inside Ω.

By Theorem 3.1, the data u|Γ+=gf0,F uniquely determine the solenoidal field Fs=F in Ω by Equation (Equation31) with φ0, and the sequence valued map Lu=u1,u2, in Ω. Moreover, the real valued mode u0 is also then recovered (with φ0 ) in Ω, by solving the Dirichlet problem for the Poisson Equation (Equation30a).

Thus, from modes u1 and u0, the scalar field f0 is also recovered in Ω by (32) f0:=2Re[u1]+[ak0]u0.(32)

In the radiative transport literature, the attenuation coefficient a=σa+σs, where σa represents pure loss due to absorption and σs(z)=12π02πk(z,θ)dθ=k0(z) is the isotropic part of scattering kernel. We consider the subcritical region: (33) σa:=ak0δ>0,forsomepositiveconstantδ.(33)

Remark 3.1

In addition to the hypothesis to Theorem 3.1, if we assume that coefficients a,k0 satisfies (Equation33), then in the region {zΩ:f0(z)=0}, one can recover explicitly the entire vector field F=2Ref1,2Imf1. Indeed, Equation (Equation15) gives u0=2Re(u1)/σa and, following (Equation16), the vector field F can be recovered by the formula (34) f1=u2+[ak1]u12¯(Re(u1)σa).(34)

Next, we show that one can also determine both scalar field f0 and vector field F, if one has the additional data gf0,0 (or g0,F) information, instead of F being incompressible as in Theorem 3.2.

Theorem 3.3

Let ΩR2 be a strictly convex bounded domain, and Γ be its boundary. Consider the boundary value problem (Equation10) for some known real valued a,k0,k1,,kMC3(Ω¯) such that (Equation10) is well-posed. If the unknown scalar field source f0 and vector field source F are real valued, W3,p(Ω;R) and W3,p(Ω;R2)-regular, respectively, with p>4, and coefficients a,k0 satisfying (Equation33), then the data gf0,F and gf0,0 defined in (Equation2) uniquely determine both f0 and F in Ω.

Proof.

Let u be the solution of the boundary value problem (Equation10) and let u=u0,u1,u2, be the sequence valued map of its non-positive Fourier modes. Since the scalar field f0W3,p(Ω;R),p>4, and isotropic vector field FW3,p(Ω;R2),p>4, then the anisotropic source f=f0+θF belong to W3,p(Ω×S1) for p>4. By applying Theorem A.2 (iii), we have uW3,p(Ω×S1),p>4. Moreover, by the Sobolev embedding, W3,p(Ω×S1)C2,μ(Ω¯×S1) with μ=12p>12, we have uC2,μ(Ω¯×S1), and thus, by [Citation26, Proposition 4.1 (ii)], the sequence valued map uYμ(Γ).

We consider the boundary value problems (35) θv+avKv=f0,subjecttov|Γ=0,v|Γ+=gf0,0,and(35) (36) θw+awKw=θF,subjecttow|Γ=0,w|Γ+=g~:=gf0,Fgf0,0.(36) Then u = v + w satisfy the boundary value problem (Equation10).

We consider first the boundary value problem (Equation35), and reconstruct the scalar field f0 from the given boundary data gf0,0 as follows.

If nZvn(z)ei is the Fourier series expansion in the angular variable θ of a solution v of boundary value problem (Equation35), then, by identifying the Fourier modes of the same order, (Equation35) reduces to the system: (37) ¯v1¯(z)+v1(z)+[a(z)k0(z)]v0(z)=f0(z),(37) (38) ¯vn(z)+vn2(z)+[a(z)kn1(z)]vn1(z)=0,0nM1,(38) (39) ¯vn(z)+vn2(z)+a(z)vn1(z)=0,nM.(39) Let v=v0,v1,v2, be the sequence valued map of its non-positive Fourier modes. By Theorem 3.1, the data gf0,0, uniquely determine the sequence Lv=v1,v2, in Ω. Moreover, as (Equation38) holds also for n = 0 (f1=0 in this case), the mode v0 is also determined in Ω by solving the Dirichlet problem for the Poisson equation (40a) Δv0=42v24[(ak1)v1],(40a) (40b) v0|Γ=g0,(40b) where the right hand side of (Equation40a) is known. Thus, using modes v0 and v1, the isotropic scalar source f0 is recovered in Ω by (41) f0(z):=2Re(v1(z))+(a(z)k0(z))v0(z),zΩ.(41) Next, we consider the boundary value problem (Equation36), and reconstruct the vector field F from the given boundary data g~=gf0,Fgf0,0 as follows.

If nZwn(z)ei is the Fourier series expansion in the angular variable θ of a solution w of the boundary value problem (Equation36), then (Equation36) reduces to the system: (42) ¯w1¯(z)+w1(z)+[a(z)k0(z)]w0(z)=0,(42) (43) ¯w0(z)+w2(z)+[a(z)k1(z)]w1(z)=(F1(z)+iF2(z))/2,(43) (44) ¯wn(z)+wn2(z)+[a(z)kn1(z)]wn1(z)=0,1nM1,(44) (45) ¯wn(z)+wn2(z)+a(z)wn1(z)=0,nM.(45) Let w=w0,w1,w2, be the sequence valued map of its non-positive Fourier modes. By Theorem 3.1, the data g~=gf0,Fgf0,0, uniquely determine the sequence Lw=w1,w2, in Ω.

Using the subcriticality condition (Equation33): σa=ak0>0, we define via (Equation42): (46) w0(z):=2Rew1(z)a(z)k0(z)=2Rew1(z)σa(z),zΩ.(46) Using determined modes w1,w2 in Lw and mode w0 from (Equation42), the real valued vector field F=2Ref1,2Imf1 is recovered in Ω by (47) f1(z):=¯w0(z)+w2(z)+[a(z)k1(z)]w1(z),zΩ.(47)

4. When can the data coming from two sources be mistaken for each other ?

In this section, we show when the data coming from two different linear anisotropic field sources can be mistaken for each other.

In Theorem 4.1 below the data are assuming the same attenuation a and scattering coefficient k.

Theorem 4.1

  1. Let aC3(Ω¯), kC3(Ω¯×S1) be real valued, with σa=ak0>0, and f0,f~0W3,p(Ω), p>4 be real valued with (f0f~0)/σaC0(Ω¯). Then F:=F~+(f0f~0σa) is a real valued vector field such that the data gf0,F coming from the linear anisotropic source f0+θF, is the same as data gf~0,F~ coming from a different linear anisotropic source f~0+θF~: gf0,F~+(f0f~0σa)=gf~0,F~.

  2. Let a,k0,k1,,kMC3(Ω¯) be real valued with σa=ak0>0. Assume that there are real valued linear anisotropic sources f=f0+θF and f~=f~0+θF~, with isotropic fields f0,f~0W3,p(Ω), p>4, and vector fields F,F~W3,p(Ω;R2), p>4. If the data gf0,F of the linear anisotropic source f equals the data gf~0,F~ of the linear anisotropic source f~. Then F=F~+(f0f~0σa).

Proof.

(i) Assume gf~0,F~ is the data of some real valued anisotropic source f~=f~0+θF~, i.e. it is the trace on Γ×S1 of solution w to the stationary transport boundary value problem: (48) θw+awKw=f~,w|Γ×S1=gf~0,F~,(48) where the operator [Kw](z,θ):=S1k(z,θθ)w(z,θ)dθ, for zΩ and θS1.

Using the subcriticality condition (Equation33): σa=ak0 with σa>0, and isotropic real valued functions ψ and σa, we note: (49) [Kψσa](z,θ)=S1k(z,θθ)[ψσa](z,θ)dθ=ψ(z)σa(z)S1k(z,θθ)dθ=ψ(z)σa(z)k0(z),(49) where second equality use the fact that both ψ and σa are angularly independent functions.

Let u:=w+(f0f~0)/σa and F:=F~+(f0f~0σa). Then θu+auKu=θ(w+f0f~0σa)+a(w+f0f~0σa)K(w+f0f~0σa)=θ∇w+awKw(aσa)f~0+(k0σa)f~0+(aσa)f0(k0σa)f0+θ(f0f~0σa)=(1aσa+k0σa)f~0+(ak0σa)f0+θ(F~+(f0f~0σa))=f0+θF=f,where the second equality uses the linearty of K and (Equation49), the third equality uses (Equation48), and the last equality uses the definition of F. Moreover, since f0f~0/σa vanishes on Γ, we get gf0,F=u|Γ×S1=w|Γ×S1+f0f~0σa|Γ=w|Γ×S1=gf~0,F~.(ii) For isotropic scalar fields f0,f~0W3,p(Ω), p>4, and vector fields F=F1,F2,F~=F1~,F2~W3,p(Ω;R2), p>4, the real valued linear anisotropic sources f=f0+θFW3,p(Ω×S1),p>4, and f~=f~0+θF~W3,p(Ω×S1), p>4.

Let the data gf0,F equals data gf~,F~ i.e. gf~,F~=g=gf0,F.Consider the corresponding boundary value problems (50a) θu+auKu=f(50a) (50b) θw+awKw=f~,(50b) respectively, subject to (50c) u|Γ×S1=g=w|Γ×S1.(50c) Since f,f~W3,p(Ω×S1),p>4, Theorem A.2 (iii), yields solutions u,wW3,p(Ω×S1),p>4. Moreover, by the Sobolev embedding, u,wC2,μ(Ω¯×S1) with μ=12p>12.

The corresponding sequences of non-positive Fourier modes {un}n0 of u satisfy (51) ¯u1¯(z)+u1(z)+[a(z)k0(z)]u0(z)=f0(z),(51) (52) ¯u0(z)+u2(z)+[a(z)k1(z)]u1(z)=(F1(z)+iF2(z))/2,(52) (53) ¯un(z)+un2(z)+[a(z)kn1(z)]un1(z)=0,1nM1,(53) (54) ¯un(z)+un2(z)+a(z)un1(z)=0,nM,(54) whereas the non-positive Fourier modes {wn}n0 of w satisfy (55) ¯w1¯(z)+w1(z)+[a(z)k0(z)]w0(z)=f~0(z),(55) (56) ¯w0(z)+w2(z)+[a(z)k1(z)]w1(z)=(F1~(z)+iF2~(z))/2,(56) (57) ¯wn(z)+wn2(z)+[a(z)kn1(z)]wn1(z)=0,1nM1,(57) (58) ¯wn(z)+wn2(z)+a(z)wn1(z)=0,nM.(58) Since the boundary data g is the same u|Γ×S1=w|Γ×S1, we also have the sequence valued map (59) g:=w|Γ=u|Γ.(59) Moreover, by [Citation26, Proposition 4.1 (ii)], the sequence gYμ(Γ) with μ>12.

Claim 4.1

The above systems subject to boundary condition (Equation59) yields (60) un=wn,foralln1,(60) inside Ω.

Proof of Claim 4.1.

We first show that the systems (Equation54) and (Equation58) subject to (Equation59) yields (61) un(z)=wn(z),zΩ,forallnM.(61) From (Equation54) and (Equation58), the shifted sequence valued maps LMu=uM,uM1, and LMw=wM,wM1,, respectively, solves (62) ¯LMu+L2LMu+aLM+1u=0,and¯LMw+L2LMw+aLM+1w=0.(62) From g in (Equation59), we use the Bukhgeim-Cauchy Integral formula (Equation6) to construct the sequence valued map v and w inside Ω: (63) v:=B(LMeGg),ρ:=B(LMeGg),(63) where eG is the operator in (Equation9).

By [Citation29, Proposition 2.3] and [Citation26, Theorem 3.2 (ii)], v,ρC1,μ(Ω;l1)Cμ(Ω¯;l1)C2(Ω;l) are L2-analytic in Ω: (64) ¯v+L2v=0,and¯ρ+L2ρ=0,(64) and also coincide at the boundary Γ. By uniqueness of L2-analytic functions with a given trace, they coincide inside: (65) v(z)=ρ(z),forzΩ.(65) Using the operator eG in (Equation9), we construct the sequence valued map (66) LMu(z):=eGv(z),andLMw(z):=eGv(z)=eGρ(z)zΩ,(66) and conclude that (Equation61) holds.

Moreover, by Lemma 2.1, the sequences LMu andLMw in (Equation66) satisfies (Equation62).

Next, we show that the systems (Equation53) and (Equation57) subject to boundary condition (Equation59) yield (67) un=wn,forall1nM1,(67) inside Ω.

Define the function (68) ψj(z):=uj(z)wj(z),forzΩ,andj1.(68) Since the boundary data g is the same, Equation (Equation50c) yields (69) ψj|Γ=0,forj1.(69) From (Equation61), we note that (70) ψj0,inΩ,forjM.(70) By subtracting system (Equation57) from (Equation53), and using (Equation68) and (Equation69), yields the boundary value problem (71a) ¯ψM+j=ψM+j2[(akM+j1)ψM+j1],1jM1,(71a) (71b) ψM+j|Γ=0.(71b)

Note that for j = 1 in (Equation71a), the right hand side of (Equation71a) contains modes ψM1 and ψM which are both zero inside Ω by (Equation70). Thus, for j = 1, the mode ψM+1 satisfy the Cauchy problem for the ¯-equation, (72a) ¯Ψ=0,inΩ,(72a) (72b) Ψ=0,onΓ.(72b) The unique solution of the above Cauchy problem is Ψ0. Therefore, resulting ψM+10.

We then solve repeatedly (Equation71a) starting for j=2,,M1, where the right hand side of (Equation71a) in each step is zero, yielding the Cauchy problem (Equation72a) for each subsequent modes, and thus, resulting in the recovering of the modes ψM+1=ψM+2=ψ2=ψ10 in Ω. Hence, establishing (Equation67).

From (Equation61) and (Equation67), we establish (Equation60), and thus Claim 4.1.

By subtracting (Equation55) from (Equation51), and using (Equation60) yields (ak0)(u0w0)=f0f~0. Using σa=ak0 with σa>0, yields (73) u0w0=f0f~0σa.(73) Moreover, by subtracting (Equation56) from (Equation52), and using (Equation60) yields 2¯(u0w0)=(F1F1~)+i(F2F2~).Since both u0 and w0 are real valued we have from (Equation73): FF~=(u0w0)=(f0f~0σa).

Remark 4.1

Note that in Theorem 4.1(i), the assumption on scattering kernels of finite Fourier content in the angular variable is not assumed, and the result holds for a general scattering kernels which depends polynomially on the angle between the directions.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

The work of D. Omogbhe and K. Sadiq were supported by the Austrian Science Fund (FWF), Project P31053-N32 and by the FWF Project F6801–N36 within the Special Research Program SFB F68 ‘Tomography Across the Scales’.

References

  • Cercignani C. The boltzmann equation and its applications. Berlin: Springer-Verlag; 1988.
  • Chandrasekhar S. Radiative transfer. New York: Dover Publ; 1960.
  • Anikonov DS, Kovtanyuk AE, Prokhorov IV. Transport equation and tomography. Brill Academics; 2002. (Inverse and Ill-Posed Problems Series; 30).
  • Choulli M, Stefanov P. An inverse boundary value problem for the stationary transport equation. Osaka J Math. 1999;36:87–104. https://projecteuclid.org/journals/osaka-journal-of-mathematics/volume-36/issue-1/An-inverse-boundary-value-problem-for-the-stationary-transport-equation/ojm/1200788449.pdf
  • Dautray R, Lions J-L. Mathematical analysis and numerical methods for science and technology. Vol. 4. Berlin: Springer-Verlag; 1990.
  • Mokhtar-Kharroubi M. Mathematical topics in neutron transport theory. Singapore: World Scientific; 1997.
  • Stefanov P, Uhlmann G. An inverse source problem in optical molecular imaging. Anal PDE. 2008;1(1):115–126. doi: 10.2140/apde
  • Fujiwara H, Sadiq K, Tamasan A. A two dimensional source reconstruction method in radiative transport using boundary data measured on an arc. Inverse Probl. 2021;37:115005. doi: 10.1088/1361-6420/ac2d75
  • Egger H, Schlottbom M. An Lp theory for stationary radiative transfer. Appl Anal. 2014;93(6):1283–1296. doi: 10.1080/00036811.2013.826798
  • Klibanov MV, Li J, Nguyen LH, et al. Convexification numerical method for a coefficient inverse problem for the radiative transport equation. SIAM J Img Sci Comput. 2023;16:35–63. doi: 10.1137/22M1509837
  • Smirnov AV, Klibanov MV, Nguyen LH. On an inverse source problem for the full radiative transfer equation with incomplete data. SIAM J Sci Comput. 2019;41(5):B929–B952. doi: 10.1137/19M1253605
  • Bal G, Tamasan A. Inverse source problems in transport equations. SIAM J Math Anal. 2007;39:57–76. doi: 10.1137/050647177
  • Sharafutdinov VA. Integral geometry of tensor fields. Utrecht: VSP; 1994.
  • Tamasan A. Tomographic reconstruction of vector fields in variable background media. Inverse Probl. 2007;23:2197–2205. doi: 10.1088/0266-5611/23/5/022
  • Fujiwara H, Sadiq K, Tamasan A. A Fourier approach to the inverse source problem in an absorbing and anisotropic scattering medium. Inverse Probl. 2019;36(1):015005. doi: 10.1088/1361-6420/ab4d98
  • Monard F, Bal G. Inverse source problems in transport via attenuated tensor tomography, arXiv:1908.06508v1.
  • Bukhgeim AL. Inversion formulas in inverse problems, chapter in linear operators and Ill-posed problems by Lavrentiev MM, and Savalev LY. New York: Plenum; 1995.
  • Natterer F. The mathematics of computerized tomography. New York: Wiley; 1986.
  • Natterer F, Wübbeling F. Mathematical methods in image reconstruction. SIAM monographs on mathematical modeling and computation. Philadelphia, PA: SIAM; 2001.
  • Sparr G, Stråhlén K, Lindström K, et al. Doppler tomography for vector fields. Inverse Probl. 1995;11:1051–1061. doi: 10.1088/0266-5611/11/5/009
  • Han W, Eichholz JA, Huang J. RTE-based bioluminescence tomography: a theoretical study. Inverse Probl Sci Eng. 2011;19(4):435–459. doi: 10.1080/17415977.2010.500383
  • Klose AD, Ntziachristos V, Hielscher AH. The inverse source problem based on the radiative transfer equation in optical molecular imaging. J Comput Phys. 2005;202:323–345. doi: 10.1016/j.jcp.2004.07.008
  • Yi HC, Sanchez R, McCormick NJ. Bioluminescence estimation from ocean in situ irradiances. Appl Opt. 1992;31:822–830. doi: 10.1364/AO.31.000822
  • Arbuzov EV, Bukhgeim AL, Kazantsev SG. Two-dimensional tomography problems and the theory of A-analytic functions. Siberian Adv Math. 1998;8:1–20.
  • Finch DV. The attenuated x-ray transform: recent developments, In: Inside out: inverse problems and applications, Math. Sci. Res. Inst. Publ., 47, Cambridge: Cambridge Univ. Press; 2003. p. 47–66.
  • Sadiq K, Tamasan A. On the range of the attenuated radon transform in strictly convex sets. Trans Amer Math Soc. 2015;367(8):5375–5398. doi: 10.1090/tran/2015-367-08
  • Sadiq K, Scherzer O, Tamasan A. On the X-ray transform of planar symmetric 2-tensors. J Math Anal Appl. 2016;442(1):31–49. doi: 10.1016/j.jmaa.2016.04.018
  • Muskhelishvili NI. Singular integral equations. New York: Dover; 2008.
  • Sadiq K, Tamasan A. On the range characterization of the two dimensional attenuated Doppler transform. SIAM J Math Anal. 2015;47(3):2001–2021. doi: 10.1137/140984282
  • Adam R. Sobolev spaces. New York: Academic Press; 1975.

Appendix A.

Some remarks on the regularity of the forward problem

In this section, we revisit the arguments in [Citation7,Citation8], and remark on the well posedness in Lp(Ω×S1) of the boundary value problem (Equation1).

The results in appendix consider both attenuation coefficient and scattering kernel in general setting. Adopting the notation in [Citation7,Citation8], we consider the operators [T11g](x,θ)=0es0a(x+tθ,θ)dtg(x+sθ,θ)ds,and[Kg](x,θ)=S1k(x,θ,θ)g(x,θ)dθ,where the intervening functions are extended by 0 outside Ω.

Using the above operators, the boundary value problem (Equation1) can be rewritten as (A1) (IT11K)u=T11f,u|Γ=0.(A1) If the operator IT11K is invertible, then the problem (EquationA1) is uniquely solvable, and has the form u=(IT11K)1T11f. By using the formal expansion (A2) u=T11f+T11KT11f+T11(KT11K)[IT11K]1T11f.(A2) We recall some results in [Citation8].

Proposition A.1

[Citation8, Proposition 2.1]

Let aC1(Ω¯×S1) and kC2(Ω¯×S1×S1). Then the operator (A3) KT11K:Lp(Ω×S1)W1,p(Ω×S1)is bounded,1<p<.(A3)

Theorem A.1

[Citation8, Theorem 2.1]

Let p>1, aC1(Ω¯×S1), and kC2(Ω¯×S1×S1). At least one of the following statements is true.

  1. IT11K is invertible in Lp(Ω×S1).

  2. there exists ϵ>0 such that IT11(λK) is invertible in Lp(Ω×S1), for any 0<|λ1|<ϵ.

For our main Theorems, we require uW3,p(Ω×S1), p>4 and such a regularity can be achieved for sources fW3,p(Ω×S1), p>4; see Theorem A.2 (iii) below. We refer to [Citation8, Theorem 2.2] for part (i) and (ii) of Theorem A.2, and we include the proof here.

The regularity of the solution u of (Equation1) increases with the regularity of f as follows.

Theorem A.2

Consider the boundary value problem (Equation1) with aC3(Ω¯×S1). For p>1, let kC3(Ω¯×S1×S1) be such that IT11K is invertible in Lp(Ω×S1), and let uLp(Ω×S1) in (EquationA2) be the solution of (Equation1).

  1. If fW1,p(Ω×S1), then uW1,p(Ω×S1).

  2. If fW2,p(Ω×S1), then uW2,p(Ω×S1).

  3. If fW3,p(Ω×S1), then uW3,p(Ω×S1).

Proof.

(i) We consider the regularity of the solution u of (Equation1) term by term as in (EquationA2). It is easy to see that the operator T11 preserve the space W1,p(Ω×S1), and also the operator K preserve the space W1,p(Ω×S1), so that the first two terms, T11f and T11KT11f, both belong to W1,p(Ω×S1). Moreover, (IT11K)1T11fLp(Ω×S1), and now, by using Proposition A.1, the last term is also belong in W1,p(Ω×S1).

(ii) We define the following operators (A4) T01u(x,θ):=0u(x+tθ,θ)dt,Kξju(x,θ):=S1kξj(x,θ,θ)u(x,θ)dθ,T~0,j1u(x,θ):=0u(x+tθ,θ)tjdt,Kηiξju(x,θ):=S12kηiξj(x,θ,θ)u(x,θ)dθ,(A4) where ηi={xi,θi} and ξj={xj,θj} for i, j = 1, 2.

It is easy to see that T01,T~0,j1,Kξj and Kηiξj preserve W1,p(Ω×S1).

By evaluating the radiative transport equation in (Equation1) at x+tθ and integrating in t from to 0, the boundary value problem (Equation1) with zero incoming fluxes is equivalent to the integral equation: (A5) u+T01(au)T01Ku=T01f.(A5) According to part (i), for fW1,p(Ω×S1), the solution uW1,p(Ω×S1), and so uxjLp(Ω×S1). In particular uxj solves the integral equation: (A6) uxj+T01(auxj)T01Kuxj=T01(axju)+T01Kxju+T01fxj.(A6) Moreover, since aC2(Ω¯×S1), kC2(Ω¯×S1×S1), and fW2,p(Ω×S1), the right-hand-side of (EquationA6) lies in W1,p(Ω×S1). By applying part (i) above, we get that the unique solution to (EquationA6) (A7) uxjW1,p(Ω×S1),j=1,2.(A7) For fW1,p(Ω×S1), also according to part (i), uθjLp(Ω×S1). In particular uθj is the unique solution of the integral equation (A8) uθj+T01(auθj)=T~0,11(auxj)T01(aθju)T~0,11(axju)+T01Kθju+T~0,11Kxju+T01fθj,(A8) which is of the type (EquationA5) with K = 0. Moreover, since fW2,p(Ω×S1), and, according to (EquationA7), uxjW1,p(Ω×S1),j=1,2, the right-hand-side of (EquationA8) lies in W1,p(Ω×S1). Again, by applying part (i), we get uθjW1,p(Ω×S1),j=1,2.Thus, uW2,p(Ω×S1).

(iii) For fW2,p(Ω×S1), according to part (ii), uxj,uθjW1,p(Ω×S1), and uxixjLp(Ω×S1). In particular uxixj is the unique solution of the integral equation (A9) uxixj+T01(auxixj)T01(Kuxixj)=T01fxixjT01(axjuxi)T01(axixju)+T01(Kxjuxi)+T01(Kxixju)T01(axiuxj)T01(Kxiuxj).(A9) Moreover, since aC3(Ω¯×S1), kC3(Ω¯×S1×S1), and fW3,p(Ω×S1), the right-hand-side of (EquationA9) lies in W1,p(Ω×S1). By applying part (i) above, we get that the unique solution to (EquationA9) (A10) uxixjW1,p(Ω×S1),i,j=1,2.(A10) For fW2,p(Ω×S1), also according to part (ii), uxj,uθjW1,p(Ω×S1), and uθiθjLp(Ω×S1). In particular uθiθj is the unique solution of the integral equation (A11) uθiθj+T01(auθiθj)=T01(fθiθj)T~0,21(axiuxj)T~0,11(axiuθj)T~0,11(aθiuxj)T01(aθiuθj)T~0,21(axjuxi)T~0,11(axjuθi)T~0,21(axixju)T~0,11(axjθiu)T~0,11(aθjuxi)T01(aθjuθi)T~0,11(aθiθju)T~0,11(Kθjuxi)T01(Kθiθju)T~0,21(Kuxixj)T~0,11(Kθiuxj),(A11) which is of the type (EquationA5) with K = 0.

Moreover, since fW3,p(Ω×S1), and, according to (EquationA10), uxixjW1,p(Ω×S1),j=1,2, the right-hand-side of (EquationA11) lies in W1,p(Ω×S1). Again, by applying part (i), we get (A12) uθiθjW1,p(Ω×S1),i,j=1,2.(A12) For fW2,p(Ω×S1), also according to part (ii), uxiuθjLp(Ω×S1). In particular uxiθj is the unique solution of the integral equation (A13) uxiθj+T01(auxiθj)T01(Kuxjθi)=T01(fxjθi)T~0,11(axiuxj)T01(aθiuxj)T01(axjθiu)T~0,11(axjuxi)T01(uθiaxj)+T~0,11(Kxjuxi)+T01(Kθiuxj)+T01(Kxjθiu),(A13) which is of the type (EquationA5). Moreover, since aC3(Ω¯×S1), kC3(Ω¯×S1×S1), and fW3,p(Ω×S1), the right-hand-side of (EquationA13) lies in W1,p(Ω×S1). Again, by applying part (i), we get (A14) uxiθjW1,p(Ω×S1),i,j=1,2.(A14) From (EquationA10), (EquationA12), and (EquationA14), we get uW3,p(Ω×S1).

We remark that for Theorem A.2 part (i) we only need aC1(Ω¯×S1) and kC2(Ω¯×S1×S1), and we only require aC2(Ω¯×S1) and kC2(Ω¯×S1×S1) for Theorem A.2 part (ii). Moreover, in a similar fashion, one can show that under sufficiently increased regularity of a and k, the solution u of (Equation1) belong to uWm,p(Ω×S1) for Zm1, provided fWm,p(Ω×S1).