This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

{psinputs}

Transport models for wave propagation in scattering media with nonlinear absorption

Joseph Kraisler Department of Applied Physics and Applied Mathematics, Columbia University, New York, NY, 10027; jek2199@columbia.edu    Wei Li Department of Mathematical Sciences, Depaul University, Chicago, IL 60604; wei.li@depaul.edu    Kui Ren Department of Applied Physics and Applied Mathematics, Columbia University, New York, NY 10027;kr2002@columbia.edu    John C. Schotland Department of Mathematics and Department of Physics, Yale University, New Haven, CT 06511; john.schotland@yale.edu    Yimin Zhong Department of Mathematics and Statistics, Auburn University, Auburn, AL 36830; yimin.zhong@auburn.edu
Abstract

This work considers the propagation of high-frequency waves in highly-scattering media where physical absorption of a nonlinear nature occurs. Using the classical tools of the Wigner transform and multiscale analysis, we derive semilinear radiative transport models for the phase-space intensity and the diffusive limits of such transport models. As an application, we consider an inverse problem for the semilinear transport equation, where we reconstruct the absorption coefficients of the equation from a functional of its solution. We obtain a uniqueness result on the inverse problem.

Key words. wave propagation, nonlinear media, nonlinear absorption, semilinear radiative transport equation, semilinear diffusion equation, inverse problem

1 Introduction

The derivation of kinetic models for wave propagation in highly-scattering media is a classical subject [32, 15] that has received significant attention in the past two decades due to its importance in many emerging applications [11, 13, 25, 39, 41, 35]. A significant amount of progress has been made on both the mathematical justification of the derivation (such as those based on multiscale analysis of the Wigner transform) [6, 21, 24, 30, 37, 49] and computational validation of the derived kinetic models [7, 9, 33, 47, 53]; see  [1, 2, 4, 5, 12, 23, 25, 28] and references therein for additional investigations in this field. The obtained models for imaging in complex media have also been utilized in many different settings [8, 9, 13, 16].

In this work, we are interested in developing kinetic models when nonlinear absorption occurs during wave propagation [10, 38, 48, 57]. We are mainly motivated by applications where reconstructing the absorption of the underlying medium from internal or boundary observations is of practical interest. Such applications include, for instance, the case of reconstructing the two-photon absorption coefficient of biological tissues with optical or photoacoustic measurements [20, 36, 42, 44, 45, 51, 54, 55, 56].

Our derivation will be carried out in the frequency domain, where wave propagation is described by the standard Helmholtz equation. The nonlinear absorption mechanism we are interested we study is modeled as a zeroth-order perturbation to the second-order Helmholtz differential operator. This is the essential factor that makes it possible for us to perform the standard calculations in this field for our nonlinear problem. Even though this is a mathematically less attractive nonlinearity to study, the derivation does provide us a formal justification of the semilinear radiative transport model, see, for instance (20), used in many applications. We refer interested readers to [22] for the derivation of transport models for a different type of nonlinearity that makes use a mean-field approximation.

The rest of this paper is organized as follows. In Section 2, we derive the radiative transport model for media with quadratic and higher-order absorption. We then discuss the diffusive limit of the derived transport models in Section 3. As an application, we study in Section 4 an inverse medium problem for our semilinear radiative transport equation. Concluding remarks are offered in Section 5.

2 Derivation of the transport equation

For simplicity, we first consider the case of quadratic nonlinear absorption. This could serve as a model of light propagation in media with two-photon absorption [51, 54]. We will then generalize the result to the case of a general polynomial nonlinearity. Let the wave field p(z,𝐱)p(z,\mathbf{x}) be the solution to the scalar wave equation in the time-harmonic form, that is,

Δ𝐱p+2pz2+k2n2p=0,\Delta_{\mathbf{x}}p+\frac{\partial^{2}p}{\partial z^{2}}+k^{2}n^{2}p=0, (1)

where Δ𝐱\Delta_{\mathbf{x}} is the transverse Laplacian in 𝐱d\mathbf{x}\in\mathbb{R}^{d} (d1d\geq 1), kk is the wave number, and n=n(z,𝐱,p)n=n(z,\mathbf{x},p) is the refractive index. We assume that the refractive index takes the form

n2=12σV(zz,𝐱𝐱)+ik1μK~(z,𝐱)|p|2,n^{2}=1-2\sigma V(\frac{z}{\ell_{z}},\frac{\mathbf{x}}{\ell_{\mathbf{x}}})+ik^{-1}\mu\widetilde{K}(z,\mathbf{x})|p|^{2}\,, (2)

where VV is a real bounded stationary random field with zero mean, with 𝐱\ell_{\mathbf{x}} and z\ell_{z} are the transverse and longitudinal correlation lengths of the random field, respectively. The deterministic function K~\widetilde{K} is non-negative and measures the strength of the second-order absorption. The parameters σ\sigma and μ\mu are the scaling factors quantifying the amplitudes of the fluctuation and the second-order absorption, respectively. Assuming that the field pp possesses a beam-like structure propagating in zz direction and neglecting the back scattering, we may write p(z,𝐱)=eikzψ(z,𝐱)p(z,\mathbf{x})=e^{ikz}\psi(z,\mathbf{x}) with complex amplitude ψ(z,𝐱)\psi(z,\mathbf{x}) satisfing the following equation

2ψz2+2ikψz+Δ𝐱ψ+k2(n21)ψ=0.\frac{\partial^{2}\psi}{\partial z^{2}}+2ik\frac{\partial\psi}{\partial z}+\Delta_{\mathbf{x}}\psi+k^{2}(n^{2}-1)\psi=0\,. (3)

Let L𝐱L_{\mathbf{x}} and LzL_{z} be the characteristic lengths of propagation in 𝐱\mathbf{x} and zz directions, respectively. We rescale the variables 𝐱L𝐱𝐱\mathbf{x}\mapsto L_{\mathbf{x}}\mathbf{x} and zLzzz\mapsto L_{z}z and 𝐱\mathbf{x}, zz are now 𝒪(1)\mathcal{O}(1). Then with the newly-defined variables, we may write (n21)(n^{2}-1) as

n21=2σV(Lzzz,L𝐱𝐱𝐱)+ik1μK~(Lz𝐳,L𝐱𝐱)|ψ|2.n^{2}-1=-2\sigma V(\frac{L_{z}z}{\ell_{z}},\frac{L_{\mathbf{x}}\mathbf{x}}{\ell_{\mathbf{x}}})+ik^{-1}\mu\widetilde{K}(L_{z}\mathbf{z},L_{\mathbf{x}}\mathbf{x})|\psi|^{2}\,. (4)

The equation (3) now becomes

1Lz22ψz2+2ikLzψz+1L𝐱2Δ𝐱ψ2k2σV(Lzzz,L𝐱𝐱𝐱)ψ+ikμK~(Lzz,L𝐱𝐱)|ψ|2ψ=0.\frac{1}{L_{z}^{2}}\frac{\partial^{2}\psi}{\partial z^{2}}+2\frac{ik}{L_{z}}\frac{\partial\psi}{\partial z}+\frac{1}{L_{\mathbf{x}}^{2}}\Delta_{\mathbf{x}}\psi-2k^{2}\sigma V(\frac{L_{z}z}{\ell_{z}},\frac{L_{\mathbf{x}}\mathbf{x}}{\ell_{\mathbf{x}}})\psi+ik\mu\widetilde{K}(L_{z}z,L_{\mathbf{x}}\mathbf{x})|\psi|^{2}\psi=0\,. (5)

With the small aperture assumption that L𝐱LzL_{\mathbf{x}}\ll L_{z}, we can formally approximate the above equation by the paraxial wave equation

iψz+Lz2kL𝐱2Δ𝐱ψkLzσV(Lzzz,L𝐱𝐱𝐱)ψ+iLz2μK~(Lzz,L𝐱𝐱)|ψ|2ψ=0.i\frac{\partial\psi}{\partial z}+\frac{L_{z}}{2kL_{\mathbf{x}}^{2}}\Delta_{\mathbf{x}}\psi-kL_{z}\sigma V(\frac{L_{z}z}{\ell_{z}},\frac{L_{\mathbf{x}}\mathbf{x}}{\ell_{\mathbf{x}}})\psi+i\frac{L_{z}}{2}\mu\widetilde{K}(L_{z}z,L_{\mathbf{x}}\mathbf{x})|\psi|^{2}\psi=0\,. (6)

Our derivation works in the regime where the longitudinal propagation distance LzL_{z} is much larger than the correlation length z\ell_{z} and the correlation length is much larger than the wavelength, that is LzzL_{z}\gg\ell_{z} and zλ:=2πk\ell_{z}\gg\lambda:=\frac{2\pi}{k}. We, therefore, introduce the small parameter ε\varepsilon, and assume the scaling relations in the weak-coupling regime:

𝐱L𝐱=zLz=ε1,k𝐱2=z,σ=1kzε,μ1Lz.\frac{\ell_{\mathbf{x}}}{L_{\mathbf{x}}}=\frac{\ell_{z}}{L_{z}}=\varepsilon\ll 1,\quad k\ell_{\mathbf{x}}^{2}=\ell_{z},\quad\sigma=\frac{1}{k\ell_{z}}\sqrt{\varepsilon},\quad\mu\propto\frac{1}{L_{z}}. (7)

Let us denote the rescaled wave field by ψε(z,𝐱)\psi_{\varepsilon}(z,\mathbf{x}) and take K(z,𝐱)=LzμK~(Lzz,L𝐱𝐱)K(z,\mathbf{x})=L_{z}\mu\widetilde{K}(L_{z}z,L_{\mathbf{x}}\mathbf{x}). Then the paraxial wave equation turns into

iψεz+ε2Δ𝐱ψε1εV(zε,𝐱ε)ψε+i2K(z,𝐱)|ψε|2ψε=0.i\frac{\partial\psi_{\varepsilon}}{\partial z}+\frac{\varepsilon}{2}\Delta_{\mathbf{x}}\psi_{\varepsilon}-\frac{1}{\sqrt{\varepsilon}}V(\frac{z}{\varepsilon},\frac{\mathbf{x}}{\varepsilon})\psi_{\varepsilon}+\frac{i}{2}K(z,\mathbf{x})|\psi_{\varepsilon}|^{2}\psi_{\varepsilon}=0. (8)

We then take the Wigner transform of ψε\psi_{\varepsilon}:

Wε(z,𝐱,𝐤)=dei𝐤𝐲ψε(𝐱ε𝐲2,z)ψε(𝐱+ε𝐲2,z)¯d𝐲(2π)d,W_{\varepsilon}(z,\mathbf{x},\mathbf{k})=\int_{\mathbb{R}^{d}}e^{i\mathbf{k}\cdot\mathbf{y}}\psi_{\varepsilon}(\mathbf{x}-\frac{\varepsilon\mathbf{y}}{2},z)\overline{\psi_{\varepsilon}(\mathbf{x}+\frac{\varepsilon\mathbf{y}}{2},z)}\frac{d\mathbf{y}}{(2\pi)^{d}}, (9)

It is then standard to check Wε(z,𝐱,𝐤)W_{\varepsilon}(z,\mathbf{x},\mathbf{k}) satisfies the Liouville equation:

Wεz+𝐤𝐱Wε+V,εWε+K,εWε=0,\frac{\partial W_{\varepsilon}}{\partial z}+\mathbf{k}\cdot\nabla_{\mathbf{x}}W_{\varepsilon}+\mathcal{L}_{V,\varepsilon}W_{\varepsilon}+\mathcal{L}_{K,\varepsilon}W_{\varepsilon}=0, (10)

where

V,εWε\displaystyle\mathcal{L}_{V,\varepsilon}W_{\varepsilon} =iεdei𝐩𝐱/ε(Wε(z,𝐱,𝐤+𝐩2)Wε(z,𝐱,𝐤𝐩2))V^(zε,𝐩)d𝐩(2π)d,\displaystyle=\frac{i}{\sqrt{\varepsilon}}\int_{\mathbb{R}^{d}}e^{-i\mathbf{p}\cdot\mathbf{x}/\varepsilon}\left(W_{\varepsilon}(z,\mathbf{x},\mathbf{k}+\frac{\mathbf{p}}{2})-W_{\varepsilon}(z,\mathbf{x},\mathbf{k}-\frac{\mathbf{p}}{2})\right)\frac{\widehat{V}(\frac{z}{\varepsilon},\mathbf{p})d\mathbf{p}}{(2\pi)^{d}}, (11)
K,εWε\displaystyle\mathcal{L}_{K,\varepsilon}W_{\varepsilon} =12dei𝐩x(Wε(z,𝐱,𝐤+ε𝐩2)+Wε(z,𝐱,𝐤ε𝐩2))S^ε(z,𝐩)d𝐩(2π)d.\displaystyle=\frac{1}{2}\int_{\mathbb{R}^{d}}e^{-i\mathbf{p}\cdot x}\left(W_{\varepsilon}(z,\mathbf{x},\mathbf{k}+\frac{\varepsilon\mathbf{p}}{2})+W_{\varepsilon}(z,\mathbf{x},\mathbf{k}-\frac{\varepsilon\mathbf{p}}{2})\right)\frac{\widehat{S}_{\varepsilon}(z,\mathbf{p})d\mathbf{p}}{(2\pi)^{d}}\,.

Here Sε(z,𝐱)S_{\varepsilon}(z,\mathbf{x}) is defined as

Sε(z,𝐱):=K(z,𝐱)|ψε(z,𝐱)|2=K(z,𝐱)dWε(z,𝐱,𝐤)𝑑𝐤,\displaystyle S_{\varepsilon}(z,\mathbf{x}):=K(z,\mathbf{x})|\psi_{\varepsilon}(z,\mathbf{x})|^{2}=K(z,\mathbf{x})\int_{\mathbb{R}^{d}}W_{\varepsilon}(z,\mathbf{x},\mathbf{k})d\mathbf{k}\,, (12)

while V^\widehat{V} and S^ε\widehat{S}_{\varepsilon} denote the Fourier transform (𝐱𝐩\mathbf{x}\to\mathbf{p}) of VV and SεS_{\varepsilon} respectively. We use the standard Fourier transform definition

f^(𝐩)=dei𝐩𝐱f(𝐱)d𝐱(2π)d.\widehat{f}(\mathbf{p})=\int_{\mathbb{R}^{d}}e^{i\mathbf{p}\cdot\mathbf{x}}f(\mathbf{x})\frac{d\mathbf{x}}{(2\pi)^{d}}\,.

2.1 Multiscale expansion

In order to find the asymptotic limit as ε0\varepsilon\to 0, we introduce 𝐲=𝐱/ε\mathbf{y}=\mathbf{x}/\varepsilon as the fast variable and denote Wε(z,𝐱,𝐤)=Wε(z,𝐱,𝐲,𝐤)W_{\varepsilon}(z,\mathbf{x},\mathbf{k})=W_{\varepsilon}(z,\mathbf{x},\mathbf{y},\mathbf{k}). Formally we write WεW_{\varepsilon} in its asymptotic expansion in ε\varepsilon:

Wε(z,𝐱,𝐲,𝐤)=W0(z,𝐱,𝐲,𝐤)+εW1(z,𝐱,𝐲,𝐤)+εW2(z,𝐱,𝐲,𝐤)+.W_{\varepsilon}(z,\mathbf{x},\mathbf{y},\mathbf{k})=W_{0}(z,\mathbf{x},\mathbf{y},\mathbf{k})+\sqrt{\varepsilon}W_{1}(z,\mathbf{x},\mathbf{y},\mathbf{k})+\varepsilon W_{2}(z,\mathbf{x},\mathbf{y},\mathbf{k})+\cdots\,.

Using (12), we may also expand Sε(z,𝐱)=Sε(z,𝐱,𝐲)S_{\varepsilon}(z,\mathbf{x})=S_{\varepsilon}(z,\mathbf{x},\mathbf{y}) accordingly as

Sε(z,𝐱,𝐲)=S0(z,𝐱,𝐲)+εS1(z,𝐱,𝐲)+εS2(z,𝐱,𝐲)+.S_{\varepsilon}(z,\mathbf{x},\mathbf{y})=S_{0}(z,\mathbf{x},\mathbf{y})+\sqrt{\varepsilon}S_{1}(z,\mathbf{x},\mathbf{y})+\varepsilon S_{2}(z,\mathbf{x},\mathbf{y})+\cdots\,.

where

S0(z,𝐱,𝐲)=K(z,𝐱)dW0(z,𝐱,𝐲,𝐤)𝑑𝐤.S_{0}(z,\mathbf{x},\mathbf{y})=K(z,\mathbf{x})\int_{\mathbb{R}^{d}}W_{0}(z,\mathbf{x},\mathbf{y},\mathbf{k})d\mathbf{k}\,.

We can now plug in the transform 𝐱𝐱+1ε𝐲\nabla_{\mathbf{x}}\to\nabla_{\mathbf{x}}+\frac{1}{\varepsilon}\nabla_{\mathbf{y}} into (10) to conclude that the leading order equation at 𝒪(ε1)\mathcal{O}(\varepsilon^{-1}) implies 𝐤𝐲W0=0\mathbf{k}\cdot\nabla_{\mathbf{y}}W_{0}=0. This is equivalent to

W0(z,𝐱,𝐲,𝐤)=W0(z,𝐱,𝐤),andS0(z,𝐱,𝐲)=S0(z,𝐱).W_{0}(z,\mathbf{x},\mathbf{y},\mathbf{k})=W_{0}(z,\mathbf{x},\mathbf{k}),\ \ \ \mbox{and}\ \ \ S_{0}(z,\mathbf{x},\mathbf{y})=S_{0}(z,\mathbf{x})\,.

For the order of 𝒪(ε1/2)\mathcal{O}(\varepsilon^{-1/2}), we have

𝐤𝐲W1+αW1+idei𝐩𝐲(W0(z,𝐱,𝐤+𝐩2)W0(z,𝐱,𝐤𝐩2))V^(zε,𝐩)d𝐩(2π)d=0,\mathbf{k}\cdot\nabla_{\mathbf{y}}W_{1}+\alpha W_{1}+i\int_{\mathbb{R}^{d}}e^{-i\mathbf{p}\cdot\mathbf{y}}(W_{0}(z,\mathbf{x},\mathbf{k}+\frac{\mathbf{p}}{2})-W_{0}(z,\mathbf{x},\mathbf{k}-\frac{\mathbf{p}}{2}))\widehat{V}(\frac{z}{\varepsilon},\mathbf{p})\frac{d\mathbf{p}}{(2\pi)^{d}}=0\,, (13)

where α0+\alpha\to 0^{+}. This gives that the Fourier transform of W1W_{1}, W^1\widehat{W}_{1}, is

W^1(z,𝐱,𝐩,𝐤)=(W0(z,𝐱,𝐤+𝐩2)W0(z,𝐱,𝐤𝐩2))V^(zε,𝐩)𝐩𝐤+iα.\widehat{W}_{1}(z,\mathbf{x},\mathbf{p},\mathbf{k})=\frac{(W_{0}(z,\mathbf{x},\mathbf{k}+\frac{\mathbf{p}}{2})-W_{0}(z,\mathbf{x},\mathbf{k}-\frac{\mathbf{p}}{2}))\widehat{V}(\frac{z}{\varepsilon},\mathbf{p})}{\mathbf{p}\cdot\mathbf{k}+i\alpha}\,. (14)

Finally, we derive the equation for 𝒪(1)\mathcal{O}(1) terms. To handle the nonlinearity, we let s>d+2s>d+2 and assume exist positive constants C1C_{1}, C2C_{2} and C3C_{3} such that

W0(z,𝐱,)C1(d)+W0(z,𝐱,)L1(d)<C1,\displaystyle\|W_{0}(z,\mathbf{x},\cdot)\|_{C^{1}(\mathbb{R}^{d})}+\|W_{0}(z,\mathbf{x},\cdot)\|_{L^{1}(\mathbb{R}^{d})}<C_{1}, (z,𝐱)d+1,\displaystyle\forall(z,\mathbf{x})\in\mathbb{R}^{d+1},
dW0(z,,𝐤)𝑑𝐤Hs(d)<C2,\displaystyle\left\|\int_{\mathbb{R}^{d}}W_{0}(z,\cdot,\mathbf{k})d\mathbf{k}\right\|_{H^{s}(\mathbb{R}^{d})}<C_{2}, z,\displaystyle\forall z\in\mathbb{R},
K(z,)Hs(d)<C3,\displaystyle\|K(z,\cdot)\|_{H^{s}(\mathbb{R}^{d})}<C_{3}, z.\displaystyle\forall z\in\mathbb{R}\,.

Then we have that

W0(z,𝐱,𝐤ε2𝐩)+W0(z,𝐱,𝐤+ε2𝐩)=2W0(z,𝐱,𝐤)+𝒪(ε|𝐩|),W_{0}(z,\mathbf{x},\mathbf{k}-\frac{\varepsilon}{2}\mathbf{p})+W_{0}(z,\mathbf{x},\mathbf{k}+\frac{\varepsilon}{2}\mathbf{p})=2W_{0}(z,\mathbf{x},\mathbf{k})+\mathcal{O}(\varepsilon|\mathbf{p}|)\,, (15)

where 𝒪(ε|𝐩|)\mathcal{O}(\varepsilon|\mathbf{p}|) has a uniform constant. Moreover, we have that

|d|𝐩||S^0(z,𝐩)|d𝐩(2π)d|2\displaystyle\left|\int_{\mathbb{R}^{d}}\frac{|\mathbf{p}||\widehat{S}_{0}(z,\mathbf{p})|d\mathbf{p}}{(2\pi)^{d}}\right|^{2} 1(2π)2dd|𝐩|2(1+|𝐩|2)s𝑑𝐩d(1+|𝐩|2)s2|S^0(z,𝐩)|2𝑑𝐩\displaystyle\leq\frac{1}{(2\pi)^{2d}}\int_{\mathbb{R}^{d}}\frac{|\mathbf{p}|^{2}}{(1+|\mathbf{p}|^{2})^{s}}d\mathbf{p}\int_{\mathbb{R}^{d}}(1+|\mathbf{p}|^{2})^{\frac{s}{2}}|\widehat{S}_{0}(z,\mathbf{p})|^{2}d\mathbf{p} (16)
=1(2π)2d(d|𝐩|2(1+|𝐩|2)s𝑑𝐩)S0(z,)Hs(d)2\displaystyle=\frac{1}{(2\pi)^{2d}}\left(\int_{\mathbb{R}^{d}}\frac{|\mathbf{p}|^{2}}{(1+|\mathbf{p}|^{2})^{s}}d\mathbf{p}\right)\|S_{0}(z,\cdot)\|_{H^{s}(\mathbb{R}^{d})}^{2}
CsK(z,)Hs(d)2dW0(z,,𝐤)𝑑𝐤Hs(d)2\displaystyle\leq C_{s}\|K(z,\cdot)\|_{H^{s}(\mathbb{R}^{d})}^{2}\left\|\int_{\mathbb{R}^{d}}W_{0}(z,\cdot,\mathbf{k})d\mathbf{k}\right\|_{H^{s}(\mathbb{R}^{d})}^{2}

is also uniformly bounded, where Cs>0C_{s}>0 is a constant depending only on ss. The last inequality holds since s>d+2s>d+2 implies s>d/2s>d/2. Therefore the 𝒪(1)\mathcal{O}(1) term is

W0z+𝐤𝐱W0+𝐤𝐲W2\displaystyle\frac{\partial W_{0}}{\partial z}+\mathbf{k}\cdot\nabla_{\mathbf{x}}W_{0}+\mathbf{k}\cdot\nabla_{\mathbf{y}}W_{2} (17)
+idei𝐩𝐲(W1(z,𝐱,𝐲,𝐤+𝐩2)W1(z,𝐱,𝐲,𝐤𝐩2))V^(zε,𝐩)d𝐩(2π)d\displaystyle+i\int_{\mathbb{R}^{d}}e^{-i\mathbf{p}\cdot\mathbf{y}}(W_{1}(z,\mathbf{x},\mathbf{y},\mathbf{k}+\frac{\mathbf{p}}{2})-W_{1}(z,\mathbf{x},\mathbf{y},\mathbf{k}-\frac{\mathbf{p}}{2}))\widehat{V}(\frac{z}{\varepsilon},\mathbf{p})\frac{d\mathbf{p}}{(2\pi)^{d}}
+W0(z,𝐱,𝐤)S0(z,𝐱)=0.\displaystyle+W_{0}(z,\mathbf{x},\mathbf{k})S_{0}(z,\mathbf{x})=0\,.

In order to close the equation, we still need to add the orthogonal relation between W0W_{0} and W2W_{2}, that is, 𝔼[𝐤𝐲W2]=0\mathbb{E}[\mathbf{k}\cdot\nabla_{\mathbf{y}}W_{2}]=0. Hence, we have

W0z+𝐤𝐱W0\displaystyle\frac{\partial W_{0}}{\partial z}+\mathbf{k}\cdot\nabla_{\mathbf{x}}W_{0} (18)
+𝔼[idei𝐩𝐲(W1(z,𝐱,𝐲,𝐤+𝐩2)W1(z,𝐱,𝐲,𝐤𝐩2))V^(zε,𝐩)d𝐩(2π)d]\displaystyle+\mathbb{E}\left[i\int_{\mathbb{R}^{d}}e^{-i\mathbf{p}\cdot\mathbf{y}}(W_{1}(z,\mathbf{x},\mathbf{y},\mathbf{k}+\frac{\mathbf{p}}{2})-W_{1}(z,\mathbf{x},\mathbf{y},\mathbf{k}-\frac{\mathbf{p}}{2}))\widehat{V}(\frac{z}{\varepsilon},\mathbf{p})\frac{d\mathbf{p}}{(2\pi)^{d}}\right]
+W0(z,𝐱,𝐤)S0(z,𝐱)=0.\displaystyle+W_{0}(z,\mathbf{x},\mathbf{k})S_{0}(z,\mathbf{x})=0\,.

Let RR be the correlation function of VV, and assume that the power spectrum satisfies

𝔼[V^(z,𝐩)V^(z,𝐪)]=(2π)dR^(𝐩)δ(𝐩+𝐪).\mathbb{E}[\widehat{V}(z,\mathbf{p})\widehat{V}(z,\mathbf{q})]=(2\pi)^{d}\widehat{R}(\mathbf{p})\delta(\mathbf{p}+\mathbf{q})\,. (19)

Then the expectation term in (18) converges weakly to

𝔼[idei𝐩𝐱/ε(W1(z,𝐱,𝐱ε,𝐤+𝐩2)W1(z,𝐱,𝐱ε,𝐤𝐩2))V^(zε,𝐩)d𝐩(2π)d]\displaystyle\mathbb{E}\left[i\int_{\mathbb{R}^{d}}e^{-i\mathbf{p}\cdot\mathbf{x}/\varepsilon}(W_{1}(z,\mathbf{x},\frac{\mathbf{x}}{\varepsilon},\mathbf{k}+\frac{\mathbf{p}}{2})-W_{1}(z,\mathbf{x},\frac{\mathbf{x}}{\varepsilon},\mathbf{k}-\frac{\mathbf{p}}{2}))\widehat{V}(\frac{z}{\varepsilon},\mathbf{p})\frac{d\mathbf{p}}{(2\pi)^{d}}\right]
4πdR^(𝐩𝐤)[W0(z,𝐱,𝐤)W0(z,𝐱,𝐩)]δ(|𝐤|2|𝐩|2)𝑑𝐩.\displaystyle\to 4\pi\int_{\mathbb{R}^{d}}\widehat{R}(\mathbf{p}-\mathbf{k})[W_{0}(z,\mathbf{x},\mathbf{k})-W_{0}(z,\mathbf{x},\mathbf{p})]\delta(|\mathbf{k}|^{2}-|\mathbf{p}|^{2})d\mathbf{p}\,.

Therefore the final radiative transport equation of W0W_{0} is

W0z+𝐤𝐱W0+4πdR^(𝐩𝐤)[W0(z,𝐱,𝐤)W0(z,𝐱,𝐩)]δ(|𝐤|2|𝐩|2)𝑑𝐩\displaystyle\frac{\partial W_{0}}{\partial z}+\mathbf{k}\cdot\nabla_{\mathbf{x}}W_{0}+4\pi\int_{\mathbb{R}^{d}}\widehat{R}(\mathbf{p}-\mathbf{k})[W_{0}(z,\mathbf{x},\mathbf{k})-W_{0}(z,\mathbf{x},\mathbf{p})]\delta(|\mathbf{k}|^{2}-|\mathbf{p}|^{2})d\mathbf{p} (20)
+(K(z,𝐱)dW0(z,𝐱,𝐤)𝑑𝐤)W0(z,𝐱,𝐤)=0.\displaystyle+\left(K(z,\mathbf{x})\int_{\mathbb{R}^{d}}W_{0}(z,\mathbf{x},\mathbf{k}^{\prime})d\mathbf{k}^{\prime}\right)W_{0}(z,\mathbf{x},\mathbf{k})=0\,.

The last term on the left side of  (20) stands for the quadratic absorption, which shows that the absorption coefficient linearly depends on the angular average of W0(𝐱,z,𝐤)W_{0}(\mathbf{x},z,\mathbf{k}).

2.2 Extension to higher order absorption

We now extend the previous result to the case of general polynomial absorption. This could be a physical model for the propagation of light in media with multi-photon absorption [14]. The absorption term in the refractive index is modeled by

n2=12σV(zz,𝐱𝐱)+ik1μl=0mK~l(z,𝐱)|ψ|2l,n^{2}=1-2\sigma V(\frac{z}{\ell_{z}},\frac{\mathbf{x}}{\ell_{\mathbf{x}}})+ik^{-1}\mu\sum_{l=0}^{m}\widetilde{K}_{l}(z,\mathbf{x})|\psi|^{2l}, (21)

where K~l\widetilde{K}_{l} stands for the absorption strength of (l+1)(l+1)-th order. Following the same derivation of the paraxial wave equation, we have a new Liouville equation in the form of  (10) where the only modification is the term

Sε(𝐱,z)=l=0LKl(z,𝐱)|ψε(z,𝐱)|2l,Kl(z,𝐱):=μLzK~l(Lzz,L𝐱𝐱).S_{\varepsilon}(\mathbf{x},z)=\sum_{l=0}^{L}K_{l}(z,\mathbf{x})|\psi_{\varepsilon}(z,\mathbf{x})|^{2l},\quad K_{l}(z,\mathbf{x}):=\mu L_{z}\widetilde{K}_{l}(L_{z}z,L_{\mathbf{x}}\mathbf{x})\,. (22)

Assuming that KlK_{l} and W0W_{0} are sufficiently regular, we can follow the same procedure and obtain the radiative transport equation for W0W_{0} in the setting of the polynomial absorption:

W0z+𝐤𝐱W0+4πdR^(𝐩𝐤)[W0(z,𝐱,𝐤)W0(z,𝐱,𝐩)]δ(|𝐤|2|𝐩|2)𝑑𝐩\displaystyle\frac{\partial W_{0}}{\partial z}+\mathbf{k}\cdot\nabla_{\mathbf{x}}W_{0}+4\pi\int_{\mathbb{R}^{d}}\widehat{R}(\mathbf{p}-\mathbf{k})[W_{0}(z,\mathbf{x},\mathbf{k})-W_{0}(z,\mathbf{x},\mathbf{p})]\delta(|\mathbf{k}|^{2}-|\mathbf{p}|^{2})d\mathbf{p} (23)
+(l=0LKl(z,𝐱)[dW0(z,𝐱,𝐤)𝑑𝐤]l)W0(z,𝐱,𝐤)=0.\displaystyle+\left(\sum_{l=0}^{L}K_{l}(z,\mathbf{x})\left[\int_{\mathbb{R}^{d}}W_{0}(z,\mathbf{x},\mathbf{k}^{\prime})d\mathbf{k}^{\prime}\right]^{l}\right)W_{0}(z,\mathbf{x},\mathbf{k})=0\,.

2.3 The nonlinear RTE in the standard form

For a monochromatic solution W0(z,𝐱,𝐤)W_{0}(z,\mathbf{x},\mathbf{k}) which is supported on 𝐤=1\|\mathbf{k}\|=1, the above radiative transport equation becomes

W0z+𝐤𝐱W0+4π𝕊d1R^(𝐩𝐤)[W0(z,𝐱,𝐤)W0(z,𝐱,𝐩)]𝑑𝐩\displaystyle\frac{\partial W_{0}}{\partial z}+\mathbf{k}\cdot\nabla_{\mathbf{x}}W_{0}+4\pi\int_{\mathbb{S}^{d-1}}\widehat{R}(\mathbf{p}-\mathbf{k})[W_{0}(z,\mathbf{x},\mathbf{k})-W_{0}(z,\mathbf{x},\mathbf{p})]d\mathbf{p} (24)
+(l=0LKl(z,𝐱)[𝕊d1W0(z,𝐱,𝐤)𝑑𝐤]l)W0(z,𝐱,𝐤)=0,\displaystyle+\left(\sum_{l=0}^{L}K_{l}(z,\mathbf{x})\left[\int_{\mathbb{S}^{d-1}}W_{0}(z,\mathbf{x},\mathbf{k}^{\prime})d\mathbf{k}^{\prime}\right]^{l}\right)W_{0}(z,\mathbf{x},\mathbf{k})=0\,,

where 𝕊d1\mathbb{S}^{d-1} is the unit sphere in d\mathbb{R}^{d}. It can further be put in the standard form

W0z+𝐤𝐱W0+Σa(W0)W0+Σs(W0𝒦W0)\displaystyle\frac{\partial W_{0}}{\partial z}+\mathbf{k}\cdot\nabla_{\mathbf{x}}W_{0}+\Sigma_{a}(\langle{W_{0}}\rangle)W_{0}+\Sigma_{s}(W_{0}-\mathcal{K}W_{0}) =0.\displaystyle=0. (25)

Here the total energy of the field at 𝐱\mathbf{x} is denoted by

W0:=𝕊d1W0(z,𝐱,𝐤)𝑑𝐤\langle{W_{0}}\rangle:=\int_{\mathbb{S}^{d-1}}W_{0}(z,\mathbf{x},\mathbf{k})d\mathbf{k}

and

Σa(W0)=l=0LΣa,lW0l\Sigma_{a}(\langle{W_{0}}\rangle)=\sum_{l=0}^{L}\Sigma_{a,l}\langle{W_{0}}\rangle^{l} (26)

is the effective absorption coefficient with Σa,l(z,𝐱)=Kl(z,𝐱)\Sigma_{a,l}(z,\mathbf{x})=K_{l}(z,\mathbf{x}) is the absorption coefficient of (l+1)(l+1)-th order. The scattering coefficient is Σs(𝐤):=4π𝕊d1R^(𝐩𝐤)𝑑𝐩\Sigma_{s}(\mathbf{k}):=4\pi\displaystyle\int_{\mathbb{S}^{d-1}}\widehat{R}(\mathbf{p}-\mathbf{k})d\mathbf{p} and the scattering operator 𝒦\mathcal{K} is defined as

𝒦u(z,𝐱,𝐤):=𝕊d1p(𝐤,𝐤)u(z,𝐱,𝐤)𝑑𝐤\mathcal{K}u(z,\mathbf{x},\mathbf{k}):=\int_{\mathbb{S}^{d-1}}p(\mathbf{k},\mathbf{k}^{\prime})u(z,\mathbf{x},\mathbf{k}^{\prime})d\mathbf{k}^{\prime} (27)

where the scattering phase function

p(𝐤,𝐤)=4πR^(𝐤𝐤)Σs(𝐤).p(\mathbf{k},\mathbf{k}^{\prime})=\frac{4\pi\widehat{R}(\mathbf{k}-\mathbf{k}^{\prime})}{\Sigma_{s}(\mathbf{k})}.

For the problem to be physical, the absorption strengths KlK_{l} (0lL0\leq l\leq L) should be all non-negative.

3 The diffusion limit

We now study the diffusion limit of the transport equation for W0(z,𝐱,𝐤)W_{0}(z,\mathbf{x},\mathbf{k}) with general polynomial nonlinear absorption. We assume that the physical domain Ω\Omega is convex with smooth boundary Ω\partial\Omega. We focus on the following nonlinear radiative transport equation:

W0z+𝐤𝐱W0+Σa(W0)W0+Σs(W0𝒦W0)\displaystyle\frac{\partial W_{0}}{\partial z}+\mathbf{k}\cdot\nabla_{\mathbf{x}}W_{0}+\Sigma_{a}(\langle{W_{0}}\rangle)W_{0}+\Sigma_{s}(W_{0}-\mathcal{K}W_{0}) =0\displaystyle=0\quad in X,\displaystyle X, (28)
W0(z,𝐱,𝐤)\displaystyle W_{0}(z,\mathbf{x},\mathbf{k}) =0\displaystyle=0\quad on Γ,\displaystyle\Gamma_{-},
W0(0,𝐱,𝐤)\displaystyle W_{0}(0,\mathbf{x},\mathbf{k}) =f(𝐱)\displaystyle=f(\mathbf{x})\quad on X0,\displaystyle X_{0},

where X=(0,T)×Ω×𝕊d1X=(0,T)\times\Omega\times\mathbb{S}^{d-1}, Γ:={(z,𝐱,𝐤)(0,T)×Ω×𝕊d1𝐧𝐱𝐤<0}\Gamma_{-}:=\{(z,\mathbf{x},\mathbf{k})\in(0,T)\times\partial\Omega\times\mathbb{S}^{d-1}\mid\mathbf{n}_{\mathbf{x}}\cdot\mathbf{k}<0\}, and 𝐧𝐱\mathbf{n}_{\mathbf{x}} is the unit outward normal vector at 𝐱Ω\mathbf{x}\in\partial\Omega, X0=Ω×𝕊d1X_{0}=\Omega\times\mathbb{S}^{d-1}.

We will focus on power spectra of the form R^(𝐩𝐤)=R^(𝐩𝐤)\widehat{R}(\mathbf{p}-\mathbf{k})=\widehat{R}(\mathbf{p}\cdot\mathbf{k}), in which case the scattering coefficient

Σs(𝐤)Σs\Sigma_{s}(\mathbf{k})\equiv\Sigma_{s}

is a constant, and the scattering phase function is p(𝐤,𝐤)=4πR^(𝐤𝐤)/Σsp(\mathbf{k},\mathbf{k}^{\prime})=4\pi\widehat{R}(\mathbf{k}\cdot\mathbf{k}^{\prime})/\Sigma_{s}. Assume the scattering phase function p(𝐤𝐤)p(\mathbf{k}\cdot\mathbf{k}^{\prime}) is bounded below and above by positive constants θ¯,θ¯0\underline{\theta},\overline{\theta}\geq 0 such that

θ¯p(𝐤,𝐤)θ¯.\underline{\theta}\leq p(\mathbf{k},\mathbf{k}^{\prime})\leq\overline{\theta}\,. (29)

Note that θ¯\underline{\theta} satisfies the condition νd1θ¯1\nu_{d-1}\underline{\theta}\leq 1, where νd1\nu_{d-1} is the measure of the (d1d-1) dimensional unit sphere. For simplicity, we require that the initial condition f(𝐱)L(Ω×𝕊d1)f(\mathbf{x})\in L^{\infty}(\Omega\times\mathbb{S}^{d-1}) is 𝐱\mathbf{x}-dependent and non-negative. The absorption coefficients obey the condition

0<Σa,l¯Σa,l(z,𝐱)Σa,l¯<,0<\underline{\Sigma_{a,l}}\leq\Sigma_{a,l}(z,\mathbf{x})\leq\overline{\Sigma_{a,l}}<\infty, (30)

for some constants Σa,l¯\underline{\Sigma_{a,l}} and Σa,l¯\overline{\Sigma_{a,l}}.

We need the following lemma to have a well-posed problem.

Lemma 3.1 (Kellogg [34]).

Let \mathcal{M} be a bounded convex open subset of a real Banach space, and F:¯¯F:\overline{\mathcal{M}}\to\overline{\mathcal{M}} is a compact continuous map which is continuously Fréchet differentiable on T. If (i) for each mm\in\mathcal{M}, 11 is not an eigenvalue of F(m)F^{\prime}(m), and (ii) for each mm\in\partial\mathcal{M}, mF(m)m\neq F(m), then FF has a unique fixed point in \mathcal{M}.

We define the space, using dS(𝐱)dS(\mathbf{x}) to denote the surface measure of Ω\partial\Omega,

𝒲p={uLp((0,T)×Ω×𝕊d1);uz+𝐤uLp((0,T)×Ω×𝕊d1);\displaystyle\mathcal{W}_{p}=\{u\in L^{p}((0,T)\times\Omega\times\mathbb{S}^{d-1});\frac{\partial u}{\partial z}+\mathbf{k}\cdot\nabla u\in L^{p}((0,T)\times\Omega\times\mathbb{S}^{d-1});
u(0,,)Lp(Ω×𝕊d1);u|ΓLp((0,T)×Γ,|𝐧𝐱𝐤|dzd𝐤dS(𝐱))}\displaystyle u(0,\cdot,\cdot)\in L^{p}(\Omega\times\mathbb{S}^{d-1});u|_{\Gamma_{-}}\in L^{p}((0,T)\times\Gamma_{-},|\mathbf{n}_{\mathbf{x}}\cdot\mathbf{k}|dz\,d\mathbf{k}dS(\mathbf{x}))\}

and the equipped norm

u𝒲pp=uLp((0,T)×Ω×d1)p+zu+𝐤uLp((0,T)×Ω×d1)p.\|u\|_{\mathcal{W}_{p}}^{p}=\|u\|_{L^{p}((0,T)\times\Omega\times\mathbb{R}^{d-1})}^{p}+\|\partial_{z}u+\mathbf{k}\cdot\nabla u\|_{L^{p}((0,T)\times\Omega\times\mathbb{R}^{d-1})}^{p}.
Theorem 3.2.

Let the initial condition f(𝐱)f(\mathbf{x}) satisfy:

(i)f1νd1inf(z,𝐱)(0,T)×ΩΣa,k1kΣa,k,1kL(i)\qquad\|f\|_{\infty}\leq\frac{1}{\nu_{d-1}}\inf_{(z,\mathbf{x})\in(0,T)\times\Omega}\frac{\Sigma_{a,k-1}}{k\Sigma_{a,k}},\quad\forall 1\leq k\leq L

or

(ii)νd1Σa(f)f2Σsθ¯,(z,𝐱)(0,T)×Ω,(ii)\qquad\nu_{d-1}\Sigma_{a}^{\prime}(\|f\|_{\infty})\|f\|_{\infty}\leq 2\Sigma_{s}\underline{\theta},\quad\forall(z,\mathbf{x})\in(0,T)\times\Omega,

where Σa\Sigma_{a}^{\prime} is the Fréchet derivative of Σa\Sigma_{a}, that is,

Σa(m)=s=1LsΣa,l(z,𝐱)ms1.\Sigma_{a}^{\prime}(m)=\sum_{s=1}^{L}s\Sigma_{a,l}(z,\mathbf{x})m^{s-1}\,.

Then equation (28) admits a unique non-negative solution in 𝒲((0,T)×Ω×𝕊d1)\mathcal{W}_{\infty}((0,T)\times\Omega\times\mathbb{S}^{d-1}).

Proof.

Let F:mϕF:m\mapsto\phi be the map defined through the relation ϕ=Fm\langle{\phi}\rangle=Fm, where ϕ\phi solves the transport equation

ϕz+𝐤ϕ+Σa(m)ϕ+Σs(ϕ𝒦ϕ)\displaystyle\frac{\partial\phi}{\partial z}+\mathbf{k}\cdot\nabla\phi+\Sigma_{a}(m)\phi+\Sigma_{s}\left(\phi-\mathcal{K}\phi\right) =0\displaystyle=0\quad in X,\displaystyle X, (31)
ϕ(z,𝐱,𝐤)\displaystyle\phi(z,\mathbf{x},\mathbf{k}) =0\displaystyle=0\quad on Γ,\displaystyle\Gamma_{-},
ϕ(0,𝐱,𝐤)\displaystyle\phi(0,\mathbf{x},\mathbf{k}) =f(𝐱)\displaystyle=f(\mathbf{x})\quad on X0.\displaystyle X_{0}\,.

We denote by \mathcal{M} the set

={mL((0,T)×Ω)L2((0,T)×Ω)0mf+δ}\mathcal{M}=\{m\in L^{\infty}((0,T)\times\Omega)\cap L^{2}((0,T)\times\Omega)\mid 0\leq m\leq\|f\|_{\infty}+\delta\}

with δ>0\delta>0 being arbitrary. It is straightforward to check that \mathcal{M} is convex, bounded, and closed under the usual L2L^{2} topology. For any mm\in\mathcal{M}, we have that Σa(m)\Sigma_{a}(m) is non-negative. Therefore, by the maximum principle for the linear transport equation (31) (see for instance [19]), ϕ\langle{\phi}\rangle\in\mathcal{M}. This shows that F:F:\mathcal{M}\to\mathcal{M}. To show that FF is a continuous operator, we denote by ϕ1\phi_{1} and ϕ2\phi_{2} the solutions to (31) corresponding to m1m_{1}\in\mathcal{M} and m2m_{2}\in\mathcal{M} respectively. We then introduce w=ϕ1ϕ2w=\phi_{1}-\phi_{2}. It is then clear that ww solves the linear transport equation:

wz+𝐤w+Σa(m1)w+Σs(w𝒦w)\displaystyle\frac{\partial w}{\partial z}+\mathbf{k}\cdot\nabla w+\Sigma_{a}(m_{1})w+\Sigma_{s}\left(w-\mathcal{K}w\right) =(Σa(m2)Σa(m1))ϕ2\displaystyle=(\Sigma_{a}(m_{2})-\Sigma_{a}(m_{1}))\phi_{2}\quad in X,\displaystyle X, (32)
w(z,𝐱,𝐤)\displaystyle w(z,\mathbf{x},\mathbf{k}) =0\displaystyle=0\quad on Γ,\displaystyle\Gamma_{-},
w(0,𝐱,𝐤)\displaystyle w(0,\mathbf{x},\mathbf{k}) =0\displaystyle=0\quad on X0.\displaystyle X_{0}\,.

By standard linear theory [19], this equation admit a unique w𝒲2w\in\mathcal{W}_{2} such that

w𝒲2((0,T)×Ω×𝕊d1)\displaystyle\|w\|_{\mathcal{W}_{2}((0,T)\times\Omega\times\mathbb{S}^{d-1})} C(Σa(m2)Σa(m1))ϕ2L2((0,T)×Ω×𝕊d1)\displaystyle\leq C\|(\Sigma_{a}(m_{2})-\Sigma_{a}(m_{1}))\phi_{2}\|_{L^{2}((0,T)\times\Omega\times\mathbb{S}^{d-1})} (33)
CΣa(f+δ)fm2m1L2((0,T)×Ω)\displaystyle\leq C\Sigma_{a}^{\prime}(\|f\|_{\infty}+\delta)\|f\|_{\infty}\|m_{2}-m_{1}\|_{L^{2}((0,T)\times\Omega)}

for some constant C>0C>0. Using the averaging lemma [27], we obtain that there exists a constant C>0C^{\prime}>0 that

Fm1Fm2H1/2((0,T)×Ω)\displaystyle\|Fm_{1}-Fm_{2}\|_{H^{1/2}((0,T)\times\Omega)} =wH1/2((0,T)×Ω)Cw𝒲2((0,T)×Ω×𝕊d1)\displaystyle=\|\langle{w}\rangle\|_{H^{1/2}((0,T)\times\Omega)}\leq C^{\prime}\|w\|_{\mathcal{W}_{2}((0,T)\times\Omega\times\mathbb{S}^{d-1})} (34)
CCΣa(f+δ)fm2m1L2((0,T)×Ω).\displaystyle\leq CC^{\prime}\Sigma_{a}^{\prime}(\|f\|_{\infty}+\delta)\|f\|_{\infty}\|m_{2}-m_{1}\|_{L^{2}((0,T)\times\Omega)}.

Combining this with the Kondrachov embedding theorem, we have shown that F:F:\mathcal{M}\to\mathcal{M} is a continuous compact operator. By the Schauder fixed point theorem, there exists a fixed point for \mathcal{M}, and hence (28) has a non-negative solution.

To prove the uniqueness of the solution, we use Lemma 3.1. We first observe that for any fixed point mm^{\ast} of FF, it must satisfy the conditions: mfm^{\ast}\leq\|f\|_{\infty} and m>0m^{\ast}>0. This is due to the fact that ff is strictly positive. Hence there are no fixed points on the boundary \partial\mathcal{M}. Next, we show that F(m)F^{\prime}(m) cannot have 11 as its eigenvalue. Let ϕ\phi be the solution to (31) with mm\in\mathcal{M} and δm\delta m be a perturbation that m+δmm+\delta m\in\mathcal{M}, we denote by ww the solution to the following equation:

wz+𝐤w+Σa(m)w+Σs(w𝒦w)\displaystyle\frac{\partial w}{\partial z}+\mathbf{k}\cdot\nabla w+\Sigma_{a}(m)w+\Sigma_{s}\left(w-\mathcal{K}w\right) =Σa(m)δmϕ\displaystyle=-\Sigma_{a}^{\prime}(m)\delta m\phi\quad in X,\displaystyle X, (35)
w(z,𝐱,𝐤)\displaystyle w(z,\mathbf{x},\mathbf{k}) =0\displaystyle=0\quad on Γ,\displaystyle\Gamma_{-},
w(0,𝐱,𝐤)\displaystyle w(0,\mathbf{x},\mathbf{k}) =0\displaystyle=0\quad on X0.\displaystyle X_{0}\,.

Then the Fréchet derivative F(m)[δm]=wF^{\prime}(m)[\delta m]=\langle{w}\rangle. Suppose F(m)F^{\prime}(m) indeed has 11 as its eigenvalue and w\langle{w}\rangle as the corresponding nonzero eigenfunction. Then F(w)=wF^{\prime}(\langle{w}\rangle)=\langle{w}\rangle and

wz+𝐤w+Σa(m)w+Σs(w𝒦w)\displaystyle\frac{\partial w}{\partial z}+\mathbf{k}\cdot\nabla w+\Sigma_{a}(m)w+\Sigma_{s}\left(w-\mathcal{K}w\right) =Σa(m)wϕ\displaystyle=-\Sigma_{a}^{\prime}(m)\langle{w}\rangle\phi\quad in X,\displaystyle X, (36)
w(z,𝐱,𝐤)\displaystyle w(z,\mathbf{x},\mathbf{k}) =0\displaystyle=0\quad on Γ,\displaystyle\Gamma_{-},
w(0,𝐱,𝐤)\displaystyle w(0,\mathbf{x},\mathbf{k}) =0\displaystyle=0\quad on X0.\displaystyle X_{0}\,.

Using the notations Σt=Σa(m)+Σs\Sigma_{t}=\Sigma_{a}(m)+\Sigma_{s} and R=Σs𝒦wΣa(m)wϕR=\Sigma_{s}\mathcal{K}w-\Sigma_{a}^{\prime}(m)\langle{w}\rangle\phi, we can write the solution to  (36) as

w(z,𝐱,𝐤)\displaystyle w(z,\mathbf{x},\mathbf{k}) =0zτ(𝐱,𝐤)exp0sΣt(zl,𝐱l𝐤)𝑑lR(zs,𝐱s𝐤,𝐤)𝑑s\displaystyle=\int_{0}^{z\wedge\tau_{-}(\mathbf{x},\mathbf{k})}\exp^{-\int_{0}^{s}\Sigma_{t}(z-l,\mathbf{x}-l\mathbf{k})dl}R(z-s,\mathbf{x}-s\mathbf{k},\mathbf{k})ds
=0zτ(𝐱,𝐤)Σt(zs,𝐱s𝐤)e0sΣt(zl,𝐱l𝐤)𝑑lR(zs,𝐱s𝐤,𝐤)Σt(zs,𝐱s𝐤)𝑑s,\displaystyle=\int_{0}^{z\wedge\tau_{-}(\mathbf{x},\mathbf{k})}\Sigma_{t}(z-s,\mathbf{x}-s\mathbf{k})e^{-\int_{0}^{s}\Sigma_{t}(z-l,\mathbf{x}-l\mathbf{k})dl}\frac{R(z-s,\mathbf{x}-s\mathbf{k},\mathbf{k})}{\Sigma_{t}(z-s,\mathbf{x}-s\mathbf{k})}ds\,,

which then gives the bound

|w(z,𝐱,𝐤)|\displaystyle|w(z,\mathbf{x},\mathbf{k})| 0zτ(𝐱,𝐤)Σt(zs,𝐱s𝐤)e0sΣt(zl,𝐱l𝐤)𝑑l|R(zs,𝐱s𝐤,𝐤)Σt(zs,𝐱s𝐤)|𝑑s\displaystyle\leq\int_{0}^{z\wedge\tau_{-}(\mathbf{x},\mathbf{k})}\Sigma_{t}(z-s,\mathbf{x}-s\mathbf{k})e^{-\int_{0}^{s}\Sigma_{t}(z-l,\mathbf{x}-l\mathbf{k})dl}\left|\frac{R(z-s,\mathbf{x}-s\mathbf{k},\mathbf{k})}{\Sigma_{t}(z-s,\mathbf{x}-s\mathbf{k})}\right|ds (37)
(1e0zτ(𝐱,𝐤)Σt(zl,𝐱l𝐤)𝑑l)sup0szτ(𝐱,𝐤)|R(zs,𝐱s𝐤,𝐤)Σt(zs,𝐱s𝐤)|\displaystyle\leq\left(1-e^{-\int_{0}^{z\wedge\tau_{-}(\mathbf{x},\mathbf{k})}\Sigma_{t}(z-l,\mathbf{x}-l\mathbf{k})dl}\right)\sup_{0\leq s\leq z\wedge\tau_{-}(\mathbf{x},\mathbf{k})}\left|\frac{R(z-s,\mathbf{x}-s\mathbf{k},\mathbf{k})}{\Sigma_{t}(z-s,\mathbf{x}-s\mathbf{k})}\right|
γsup(0,T)×Ω×𝕊d1|R(z,𝐱,𝐤)Σt(z,𝐱)|\displaystyle\leq\gamma\sup_{(0,T)\times\Omega\times\mathbb{S}^{d-1}}\left|\frac{R(z,\mathbf{x},\mathbf{k})}{\Sigma_{t}(z,\mathbf{x})}\right|

for some 0<γ<10<\gamma<1. Here ab=min(a,b)a\wedge b=\min(a,b) and τ(𝐱,𝐤)\tau_{-}(\mathbf{x},\mathbf{k}) is the distance from 𝐱\mathbf{x} to Γ\Gamma_{-} in the direction of 𝐤-\mathbf{k}. The next step is to show that sup(0,T)×Ω×𝕊d1|R(z,𝐱,𝐤)Σt(z,𝐱)|w\sup_{(0,T)\times\Omega\times\mathbb{S}^{d-1}}\left|\frac{R(z,\mathbf{x},\mathbf{k})}{\Sigma_{t}(z,\mathbf{x})}\right|\leq\|w\|_{\infty}, which, when combined with the bound in (37), leads to the bound |w(z,𝐱,𝐤)|γw|w(z,\mathbf{x},\mathbf{k})|\leq\gamma\|w\|_{\infty} (and hence w=0w=0). This contradicts the assumption that w\langle{w}\rangle is the eigenfunction corresponding to the eigenvalue 11 of FF^{\prime}. We derive the bound under the two assumptions in the theorem.

  1. (a)

    When condition (i) is satisfied, we deduce from it that

    νd1Σa(m)fΣa(m).\nu_{d-1}\Sigma_{a}^{\prime}(m)\|f\|_{\infty}\leq\Sigma_{a}(m)\,.

    Meanwhile, we also have that

    |R(z,𝐱,𝐤)|Σsw+νd1Σa(m)fw,|R(z,\mathbf{x},\mathbf{k})|\leq\Sigma_{s}\|w\|_{\infty}+\nu_{d-1}\Sigma_{a}^{\prime}(m)\|f\|_{\infty}\|w\|_{\infty},

    Combining the above two bounds gives us

    |R(z,𝐱,𝐤)|(Σs+Σa(m))w=Σtw.|R(z,\mathbf{x},\mathbf{k})|\leq(\Sigma_{s}+\Sigma_{a}(m))\|w\|_{\infty}=\Sigma_{t}\|w\|_{\infty}.

    Therefore, we have sup(0,T)×Ω×𝕊d1|R(z,𝐱,𝐤)Σt(z,𝐱)|w\sup_{(0,T)\times\Omega\times\mathbb{S}^{d-1}}\left|\frac{R(z,\mathbf{x},\mathbf{k})}{\Sigma_{t}(z,\mathbf{x})}\right|\leq\|w\|_{\infty}.

  2. (b)

    For the case when condition (ii) is satisfied, we first observe that

    R(z,𝐱,𝐤)=Σs(𝒦wθ¯w)+(Σsθ¯Σa(m)ϕ)wR(z,\mathbf{x},\mathbf{k})=\Sigma_{s}\left(\mathcal{K}w-\underline{\theta}\langle{w}\rangle\right)+\left(\Sigma_{s}\underline{\theta}-\Sigma_{a}^{\prime}(m)\phi\right)\langle{w}\rangle

    implies

    |R(z,𝐱,𝐤)Σt(z,𝐱)|Σs(1νd1θ¯)+νd1|Σsθ¯Σa(m)ϕ|Σt(z,𝐱)w.\left|\frac{R(z,\mathbf{x},\mathbf{k})}{\Sigma_{t}(z,\mathbf{x})}\right|\leq\frac{\Sigma_{s}(1-\nu_{d-1}\underline{\theta})+\nu_{d-1}|\Sigma_{s}\underline{\theta}-\Sigma_{a}^{\prime}(m)\phi|}{\Sigma_{t}(z,\mathbf{x})}\|w\|_{\infty}. (38)

    When f\|f\|_{\infty} satisfies

    f1νd12Σsθ¯Σa(f+δ),(z,𝐱)(0,T)×Ω,\|f\|_{\infty}\leq\frac{1}{\nu_{d-1}}\frac{2\Sigma_{s}\underline{\theta}}{\Sigma_{a}^{\prime}(\|f\|_{\infty}+\delta)},\quad\forall(z,\mathbf{x})\in(0,T)\times\Omega, (39)

    we obtain that Σs(1νd1θ¯)+νd1|Σsθ¯Σa(m)ϕ|Σs\Sigma_{s}(1-\nu_{d-1}\underline{\theta})+\nu_{d-1}|\Sigma_{s}\underline{\theta}-\Sigma_{a}^{\prime}(m)\phi|\leq\Sigma_{s} which, when combined with (38), implies that

    |R(z,𝐱,𝐤)Σt(z,𝐱)|w.\left|\frac{R(z,\mathbf{x},\mathbf{k})}{\Sigma_{t}(z,\mathbf{x})}\right|\leq\|w\|_{\infty}.

    Since δ>0\delta>0 is arbitrary, taking δ0\delta\to 0, the condition (39) becomes

    f1νd12Σsθ¯Σa(f),(z,𝐱)(0,T)×Ω,\|f\|_{\infty}\leq\frac{1}{\nu_{d-1}}\frac{2\Sigma_{s}\underline{\theta}}{\Sigma_{a}^{\prime}(\|f\|_{\infty})},\quad\forall(z,\mathbf{x})\in(0,T)\times\Omega\,,

    which is simply (ii).

The proof is now complete. ∎

To study the diffusion limit, we introduce a small parameter ϵ\epsilon and denote by WϵW_{\epsilon} the solution to the scaled radiative transport equation:

Wϵz+1ϵ𝐤Wϵ+Σa(Wϵ)Wϵ+1ϵ2Σs(Wϵ𝒦Wϵ)\displaystyle\frac{\partial W_{\epsilon}}{\partial z}+\frac{1}{\epsilon}\mathbf{k}\cdot\nabla W_{\epsilon}+\Sigma_{a}(\langle{W_{\epsilon}}\rangle)W_{\epsilon}+\frac{1}{\epsilon^{2}}\Sigma_{s}\left(W_{\epsilon}-\mathcal{K}W_{\epsilon}\right) =0\displaystyle=0\quad in X,\displaystyle X, (40)
Wϵ(z,𝐱,𝐤)\displaystyle W_{\epsilon}(z,\mathbf{x},\mathbf{k}) =0\displaystyle=0\quad on Γ,\displaystyle\Gamma_{-},
Wϵ(0,𝐱,𝐤)\displaystyle W_{\epsilon}(0,\mathbf{x},\mathbf{k}) =f(𝐱)\displaystyle=f(\mathbf{x})\quad on X0.\displaystyle X_{0}\,.

We have the following corollary using condition (ii) in Theorem 3.2.

Corollary 3.3.

If ϵ\epsilon is sufficiently small such that

Σa(f)f1νd12θ¯Σsϵ2,\Sigma_{a}^{\prime}(\|f\|_{\infty})\|f\|_{\infty}\leq\frac{1}{\nu_{d-1}}\frac{2\underline{\theta}\Sigma_{s}}{\epsilon^{2}},

then  (40) admits a unique non-negative solution in 𝒲((0,T)×Ω×𝕊d1)\mathcal{W}_{\infty}((0,T)\times\Omega\times\mathbb{S}^{d-1}). Moreover, the solution satisfies Wϵf\|W_{\epsilon}\|_{\infty}\leq\|f\|_{\infty}.

3.1 Asymptotic expansion

Let ϵ\epsilon be sufficiently small such that (40) admits a unique solution. We formally expand the solution in powers of ϵ\epsilon:

Wϵ(z,𝐱,𝐯)=W0(z,𝐱,𝐯)+ϵW1(z,𝐱,𝐯)+ϵ2W2(z,𝐱,𝐯)+ϕϵ(z,𝐱,𝐯).W_{\epsilon}(z,\mathbf{x},\mathbf{v})=W_{0}(z,\mathbf{x},\mathbf{v})+\epsilon W_{1}(z,\mathbf{x},\mathbf{v})+\epsilon^{2}W_{2}(z,\mathbf{x},\mathbf{v})+\phi_{\epsilon}(z,\mathbf{x},\mathbf{v})\,. (41)

Substituting this into (40) and matching the equations at orders ϵ2\epsilon^{-2}, ϵ1\epsilon^{-1} and ϵ0\epsilon^{0} gives the system:

(𝒦)W0\displaystyle(\mathcal{I}-\mathcal{K})W_{0} =0,\displaystyle=0, (42)
Σs(𝒦)W1+𝐤W0\displaystyle\Sigma_{s}(\mathcal{I}-\mathcal{K})W_{1}+\mathbf{k}\cdot\nabla W_{0} =0,\displaystyle=0,
W0z+Σs(𝒦)W2+𝐤W1+Σa(W0)W0\displaystyle\frac{\partial W_{0}}{\partial z}+\Sigma_{s}(\mathcal{I}-\mathcal{K})W_{2}+\mathbf{k}\cdot\nabla W_{1}+\Sigma_{a}({W_{0}})W_{0} =0.\displaystyle=0\,.

Following standard procedures [15, 19], we obtain from the first two equations that

W0(z,𝐱,𝐤)=W0(z,𝐱),W1(z,𝐱,𝐤)=i=1dDi(𝐤)ΣsxiW0(z,𝐱),W_{0}(z,\mathbf{x},\mathbf{k})=W_{0}(z,\mathbf{x}),\quad W_{1}(z,\mathbf{x},\mathbf{k})=-\sum_{i=1}^{d}\frac{D_{i}(\mathbf{k})}{\Sigma_{s}}\partial_{x_{i}}W_{0}(z,\mathbf{x})\,, (43)

where Di(v)L(𝕊d1)D_{i}(v)\in L^{\infty}(\mathbb{S}^{d-1}) are the unique solutions to

(𝒦)Di(𝐤)=𝐤𝐞i,𝕊d1Di(𝐤)𝑑𝐤=0,i=1,2,,d.(\mathcal{I}-\mathcal{K})D_{i}(\mathbf{k})=\mathbf{k}\cdot\mathbf{e}_{i},\quad\int_{\mathbb{S}^{d-1}}D_{i}(\mathbf{k})d\mathbf{k}=0\,,\quad i=1,2,\dots,d. (44)

Next, we integrate the third equation in (42) over 𝕊d1\mathbb{S}^{d-1}, and utilize the fact that

(𝒦)W2=0\langle{(\mathcal{I}-\mathcal{K})W_{2}}\rangle=0

to get the equation for W0W_{0}. This leads to

𝕊d1(W0z+(𝐤W1+Σa(W0)W0))𝑑𝐤=0.\int_{\mathbb{S}^{d-1}}\left(\frac{\partial W_{0}}{\partial z}+\left(\mathbf{k}\cdot\nabla W_{1}+\Sigma_{a}(W_{0})W_{0}\right)\right)d\mathbf{k}=0\,.

Since W0W_{0} is independent of 𝐤\mathbf{k}, we have that W0W_{0} solves the diffusion equation:

W0z(AΣsW0)+Σa(W0)W0\displaystyle\frac{\partial W_{0}}{\partial z}-\nabla\cdot\left(\frac{A}{\Sigma_{s}}\nabla W_{0}\right)+\Sigma_{a}(W_{0})W_{0} =0,\displaystyle=0,\quad in (0,T)×Ω,\displaystyle(0,T)\times\Omega, (45)
W0(z,𝐱)\displaystyle W_{0}(z,\mathbf{x}) =0,\displaystyle=0,\quad on (0,T)×Ω,\displaystyle(0,T)\times\partial\Omega,
W0(0,𝐱)\displaystyle W_{0}(0,\mathbf{x}) =f(𝐱)\displaystyle=f(\mathbf{x})\quad on Ω,\displaystyle\Omega,

where the matrix AA is positive definite and is defined as

Aij=𝕊d1𝐤𝐞iDj(𝐤)𝑑𝐤.A_{ij}=\int_{\mathbb{S}^{d-1}}\mathbf{k}\cdot\mathbf{e}_{i}D_{j}(\mathbf{k})d\mathbf{k}\,.

Under our assumption that the scattering phase function p(𝐤𝐤)p(\mathbf{k}\cdot\mathbf{k}^{\prime}) is rotation invariant, we have that Aij=1d(1g)δijA_{ij}=\frac{1}{d(1-g)}\delta_{ij}, where g(1,1)g\in(-1,1) is the anisotropy parameter.

Let ϕϵ\phi_{\epsilon} be the remainder in the expansion. Then we can check that ϕϵ\phi_{\epsilon} satisfies the equation:

zϕϵ+1ϵ𝐤ϕϵ+Σsϵ2(𝒦)ϕϵ+Σa(Wϵ)ϕϵ\displaystyle\partial_{z}\phi_{\epsilon}+\frac{1}{\epsilon}\mathbf{k}\cdot\nabla\phi_{\epsilon}+\frac{\Sigma_{s}}{\epsilon^{2}}\left(\mathcal{I}-\mathcal{K}\right)\phi_{\epsilon}+\Sigma_{a}(\langle{W_{\epsilon}}\rangle)\phi_{\epsilon} =ϵh1,\displaystyle=\epsilon h_{1},\; in X\displaystyle X (46)
ϕϵ(z,𝐱,𝐤)\displaystyle\phi_{\epsilon}(z,\mathbf{x},\mathbf{k}) =ϵh2,\displaystyle=\epsilon h_{2},\quad on Γ,\displaystyle\Gamma_{-},
ϕϵ(0,𝐱,𝐤)\displaystyle\phi_{\epsilon}(0,\mathbf{x},\mathbf{k}) =ϵh3,\displaystyle=\epsilon h_{3},\quad on X0,\displaystyle X_{0}\,,

where the functions h1h_{1}, h2h_{2}, and h3h_{3} are respectively of the forms:

h1(z,𝐱,𝐤)\displaystyle h_{1}(z,\mathbf{x},\mathbf{k}) =zW1𝐤W2Σa(Wϵ)(W1+ϵW2)\displaystyle=-\partial_{z}W_{1}-\mathbf{k}\cdot\nabla W_{2}-\Sigma_{a}(\langle{W_{\epsilon}}\rangle)(W_{1}+\epsilon W_{2})
ϵzW2+1ϵ[Σa(W0)Σa(Wϵ)]W0,\displaystyle\quad-\epsilon\partial_{z}W_{2}+\frac{1}{\epsilon}\left[\Sigma_{a}(W_{0})-\Sigma_{a}(\langle{W_{\epsilon}}\rangle)\right]W_{0},
h2(z,𝐱,𝐤)\displaystyle h_{2}(z,\mathbf{x},\mathbf{k}) =i=1dDi(𝐤)ΣsW0xiϵW2,\displaystyle=-\sum_{i=1}^{d}\frac{D_{i}(\mathbf{k})}{\Sigma_{s}}\frac{\partial W_{0}}{\partial x_{i}}-\epsilon W_{2},
h3(0,𝐱,𝐤)\displaystyle h_{3}(0,\mathbf{x},\mathbf{k}) =W1ϵW2.\displaystyle=-W_{1}-\epsilon W_{2}\,.

Let δ:=12inf[0,T]×ΩΣa,0(z,𝐱)\delta:=\frac{1}{2}\inf_{[0,T]\times\Omega}\Sigma_{a,0}(z,\mathbf{x}). Then by the assumtion in (30), δ>0\delta>0. We then have that for any z[0,T)z\in[0,T),

ϕϵ(z,,)\displaystyle\|\phi_{\epsilon}(z,\cdot,\cdot)\|_{\infty} ϵh3eδz+ϵ0zeδ(zs)(h1+δh2)𝑑s\displaystyle\leq\epsilon\|h_{3}\|_{\infty}e^{-\delta z}+\epsilon\int_{0}^{z}e^{-\delta(z-s)}(\|h_{1}\|_{\infty}+\delta\|h_{2}\|_{\infty})ds (47)
ϵT(h1+δh2)+ϵh3.\displaystyle\leq\epsilon T(\|h_{1}\|_{\infty}+\delta\|h_{2}\|_{\infty})+\epsilon\|h_{3}\|_{\infty}.

It remains to show that h1h_{1}, h2h_{2}, and h3h_{3} are bounded. We first observe from the equations in (42) that

W1C1W0C([0,T),C1(Ω)),W2C2W0C([0,T),C2(Ω))\|W_{1}\|_{\infty}\leq C_{1}\|W_{0}\|_{C([0,T),C^{1}(\Omega))},\quad\|W_{2}\|_{\infty}\leq C_{2}\|W_{0}\|_{C([0,T),C^{2}(\Omega))}

We then take the derivative with respect to zz of the equations in (42) to deduce that

zW1C3W0C([0,T),C3(Ω)),zW2C4W0C([0,T),C4(Ω))\|\partial_{z}W_{1}\|_{\infty}\leq C_{3}\|W_{0}\|_{C([0,T),C^{3}(\Omega))},\quad\|\partial_{z}W_{2}\|_{\infty}\leq C_{4}\|W_{0}\|_{C([0,T),C^{4}(\Omega))}

together with

𝐤W2C5W0C([0,T),C3(Ω)).\|\mathbf{k}\cdot\nabla W_{2}\|_{\infty}\leq C_{5}\|W_{0}\|_{C([0,T),C^{3}(\Omega))}\,.

Therefore, given that W0C([0,T),C4(Ω))W_{0}\in C([0,T),C^{4}(\Omega)), we have the following estimate for the diffusion approximation:

WϵW0ϕϵ+ϵW1+ϵ2W2=C(T,W0C([0,T),C4(Ω)))ϵ.\|W_{\epsilon}-W_{0}\|_{\infty}\leq\|\phi_{\epsilon}\|_{\infty}+\epsilon\|W_{1}\|_{\infty}+\epsilon^{2}\|W_{2}\|_{\infty}=C(T,\|W_{0}\|_{C([0,T),C^{4}(\Omega))})\epsilon.

To ensure the regularity of the solution W0W_{0}, at least for a short time, we simply need the initial condition ff to be smooth enough since W0W_{0} would be smoother than the initial condition due to the diffusive nature.

We can prove the following result.

Theorem 3.4.

Assume that fC04(Ω)f\in C^{4}_{0}(\Omega) is non-negative, and the absorption and scattering coefficients satisfy the condition

Σa,lC2(Ω),Σa,l0,Σs>0.\Sigma_{a,l}\in C^{2}(\Omega),\quad\Sigma_{a,l}\geq 0,\quad\Sigma_{s}>0.

Then the diffusion equation (45) admits a unique strong solution W0C([0,T),C4(Ω))W_{0}\in C([0,T),C^{4}(\Omega)) when 1d31\leq d\leq 3.

Proof.

Let L=(AΣs)L=-\nabla\cdot\left(\frac{A}{\Sigma_{s}}\nabla\right). Then L-L is the infinitesimal generator of an analytic semigroup G(t)G(t) on L2(Ω)L^{2}(\Omega) and G(t)1\|G(t)\|\leq 1 for all t0t\geq 0. We denote

𝒟(L)=H2(Ω)H01(Ω).\mathcal{D}(L)=H^{2}(\Omega)\cap H^{1}_{0}(\Omega).

By [43, Theorem 8.4.4] and [43, Theorem 6.3.1], we have that when f𝒟(L)f\in\mathcal{D}(L), there exists a unique local strong solution W0W_{0} to the equation (45) on (0,T)×Ω(0,T)\times\Omega, that is,

W0C([0,T),L2(Ω))C((0,T),H2(Ω)H01(Ω))C1((0,T),L2(Ω)).W_{0}\in C([0,T),L^{2}(\Omega))\cap C((0,T),H^{2}(\Omega)\cap H^{1}_{0}(\Omega))\cap C^{1}((0,T),L^{2}(\Omega))\,. (48)

Moreover, 0W0f0\leq W_{0}\leq\|f\|_{\infty} by the comparison principle. Then the result of  [43, Corollary 6.3.2] ensures that zW0\partial_{z}W_{0} is locally Hölder continuous for z(0,T)z\in(0,T). Hence W0(z,𝐱)W_{0}(z,\mathbf{x}) and zW0(z,𝐱)\partial_{z}W_{0}(z,\mathbf{x}) are both continuous on (0,T)×Ω¯(0,T)\times\overline{\Omega}. This means that Σa(W0)\Sigma_{a}(W_{0}) is also continuous. Therefore we must have W0(z,)C2(Ω)W_{0}(z,\cdot)\in C^{2}(\Omega), which means W0W_{0} is a classical solution.

Let g(𝐱):=LfΣa(f)g(\mathbf{x}):=-Lf-\Sigma_{a}(f), fC2(Ω)𝒟(L)f\in C^{2}(\Omega)\in\mathcal{D}(L). By differentiating (45), we find that ψ:=zW0\psi:=\partial_{z}W_{0} satisfies the following equation:

zψ+Lψ+(Σa(W0)W0+Σa(W0))ψ\displaystyle\partial_{z}\psi+L\psi+\left(\Sigma_{a}^{\prime}(W_{0})W_{0}+\Sigma_{a}(W_{0})\right)\psi =0\displaystyle=0\quad in (0,T)×Ω,\displaystyle(0,T)\times\Omega, (49)
ψ(z,𝐱)\displaystyle\psi(z,\mathbf{x}) =0\displaystyle=0\quad on (0,T)×Ω,\displaystyle(0,T)\times\partial\Omega,
ψ(0,𝐱)\displaystyle\psi(0,\mathbf{x}) =g(𝐱)\displaystyle=g(\mathbf{x})\quad on {0}×Ω.\displaystyle\{0\}\times\Omega.

Following a similar process, we can deduce that ψ(z,)C2(Ω)\psi(z,\cdot)\in C^{2}(\Omega). Since W0(z,)C2(Ω)W_{0}(z,\cdot)\in C^{2}(\Omega) and zW0(z,)C2(Ω)\partial_{z}W_{0}(z,\cdot)\in C^{2}(\Omega) for z(0,T)z\in(0,T), we have zW0(z,)+Σa(W0(z,))W0(z,)C2(Ω)\partial_{z}W_{0}(z,\cdot)+\Sigma_{a}(W_{0}(z,\cdot))W_{0}(z,\cdot)\in C^{2}(\Omega). By classical regularity theory for elliptic equations [26], W0(z,)C4(Ω)W_{0}(z,\cdot)\in C^{4}(\Omega). ∎

Remark 3.5.

We have assumed so far that the initial condition ff is independent of the variable 𝐤\mathbf{k}. In fact, the case of ff depending on 𝐤\mathbf{k}, that is, f=f(𝐱,𝐤)f=f(\mathbf{x},\mathbf{k}), can be treated in a similar manner by introducing another fast variable θ=zϵ2\theta=\frac{z}{\epsilon^{2}}, as in [19, Section XXI.5.3]. We will not repeat the calculations here.

3.2 The case of degenerate coefficients

Let us now briefly consider the case when the problem is degenerate, that is, when the absorption coefficient can vanish in part of the domain of interest. More precisely, we relax the requirement that all Σa,l>0\Sigma_{a,l}>0 to the following:

Σa,l0,Σa(f)>0,(z,𝐱)[0,T]×Ω.\Sigma_{a,l}\geq 0,\quad\Sigma_{a}(\|f\|_{\infty})>0,\quad\forall(z,\mathbf{x})\in[0,T]\times\Omega. (50)

In this case, Σa(f)>0\Sigma_{a}^{\prime}(\|f\|_{\infty})>0. When ϵ\epsilon is sufficiently small, the scaled transport equation (40) admits a unique solution in L(X)L^{\infty}(X). Let wϵw_{\epsilon} be the solution to the following linear transport equation:

zwϵ+1ϵ𝐤wϵ+Σa(f)wϵ+Σsϵ2(𝒦)wϵ\displaystyle\partial_{z}w_{\epsilon}+\frac{1}{\epsilon}\mathbf{k}\cdot\nabla w_{\epsilon}+\Sigma_{a}(\|f\|_{\infty})w_{\epsilon}+\frac{\Sigma_{s}}{\epsilon^{2}}(\mathcal{I}-\mathcal{K})w_{\epsilon} =0\displaystyle=0\quad in X,\displaystyle X, (51)
wϵ(z,𝐱,𝐤)\displaystyle w_{\epsilon}(z,\mathbf{x},\mathbf{k}) =0\displaystyle=0\quad on Γ,\displaystyle\Gamma_{-},
wϵ(0,𝐱,𝐤)\displaystyle w_{\epsilon}(0,\mathbf{x},\mathbf{k}) =f(𝐱)\displaystyle=f(\mathbf{x})\quad on X0.\displaystyle X_{0}.

Since the absorption coefficient Σa(f)Σa(Wϵ)\Sigma_{a}(\|f\|_{\infty})\geq\Sigma_{a}(\langle{W_{\epsilon}}\rangle), we conclude that wϵWϵw_{\epsilon}\leq W_{\epsilon} for any ϵ>0\epsilon>0. On the other hand, as ϵ0\epsilon\to 0, we have wϵw0w_{\epsilon}\to w_{0}, where w0w_{0} is the solution to

zw0(AΣsw0)+Σa(f)w0\displaystyle\partial_{z}w_{0}-\nabla\cdot\left(\frac{A}{\Sigma_{s}}\nabla w_{0}\right)+\Sigma_{a}(\|f\|_{\infty})w_{0} =0\displaystyle=0\quad in (0,T)×Ω,\displaystyle(0,T)\times\Omega, (52)
w0(z,𝐱)\displaystyle w_{0}(z,\mathbf{x}) =0\displaystyle=0\quad on (0,T)×Ω,\displaystyle(0,T)\times\partial\Omega,
w0(0,𝐱)\displaystyle w_{0}(0,\mathbf{x}) =f(𝐱)\displaystyle=f(\mathbf{x})\quad on Ω.\displaystyle\Omega.

Hence Wϵwϵinf[0,T]×Ωw0Cϵ\|W_{\epsilon}\|\geq\|w_{\epsilon}\|\geq\inf_{[0,T]\times\Omega}w_{0}-C\epsilon for some C>0C>0. Because inf(0,T)×Ωw0>0\inf_{(0,T)\times\Omega}w_{0}>0 strictly for ϵ\epsilon sufficiently small, we conclude that WϵW_{\epsilon} is bounded from below by a positive number, which implies Σa(Wϵ)\Sigma_{a}(\langle{W_{\epsilon}}\rangle) is also strictly positive. Then repeat the process in (47) by setting δ=12inf(0,T)×ΩΣa(Wϵ)\delta=\frac{1}{2}\inf_{(0,T)\times\Omega}\Sigma_{a}(\langle{W_{\epsilon}}\rangle) instead, we obtain the same conclusion that WϵW0Cϵ\|W_{\epsilon}-W_{0}\|_{\infty}\leq C^{\prime}\epsilon for a constant CC^{\prime} independent of ϵ\epsilon.

4 An application in inverse problems

We now consider the following inverse medium problem as a direct application of the transport model:

zu+𝐤u+Σa(u)u+Σs(𝒦)u\displaystyle\partial_{z}u+\mathbf{k}\cdot\nabla u+\Sigma_{a}(\langle{u}\rangle)u+\Sigma_{s}(\mathcal{I}-\mathcal{K})u =0\displaystyle=0\quad in X,\displaystyle X, (53)
u(z,𝐱,𝐤)\displaystyle u(z,\mathbf{x},\mathbf{k}) =0\displaystyle=0\quad on Γ,\displaystyle\Gamma_{-},
u(0,𝐱,𝐤)\displaystyle u(0,\mathbf{x},\mathbf{k}) =f(𝐱)\displaystyle=f(\mathbf{x})\quad on X0,\displaystyle X_{0},

where f(𝐱)L(X0)f(\mathbf{x})\in L^{\infty}(X_{0}) is a strictly positive source function. We assume that (53) has a unique positive solution u𝒲u\in\mathcal{W}_{\infty}.

We assume that the absorption coefficient Σa\Sigma_{a} is not known, but we have additional data that is the density of the solution, that is,

g(z,𝐱)=u:=𝕊d1u(z,𝐱,𝐤)𝑑𝐤.g(z,\mathbf{x})=\langle{u}\rangle:=\int_{\mathbb{S}^{d-1}}u(z,\mathbf{x},\mathbf{k})d\mathbf{k}\,. (54)

The inverse problem amounts to finding the unknown absorption coefficients Σa,l\Sigma_{a,l} from the observed data gg from a given ff.

We can prove the following result.

Theorem 4.1.

Let gg and g~\widetilde{g} be data of the form (54) generated from (53) with coefficients Σa\Sigma_{a} and Σ~a\widetilde{\Sigma}_{a} respectively. Then g=g~g=\widetilde{g} implies Σa(u)=Σ~a(u~)\Sigma_{a}(\langle{u}\rangle)=\widetilde{\Sigma}_{a}(\langle{\widetilde{u}}\rangle).

Proof.

Let δu=uu~\delta u=u-\widetilde{u}. We verify that for any δu\delta u, we have the identity

Ω×𝕊d1[𝐤δu]δuu~𝑑𝐱𝑑𝐤=Ω×𝕊d1𝐤|δu|22u~d𝐱d𝐤Ω×𝕊d1[𝐤1u~]|δu|22𝑑𝐱𝑑𝐤,\int_{\Omega\times\mathbb{S}^{d-1}}\left[\mathbf{k}\cdot\nabla\delta u\right]\frac{\delta u}{\tilde{u}}d\mathbf{x}d\mathbf{k}=\int_{\Omega\times\mathbb{S}^{d-1}}\mathbf{k}\cdot\nabla\frac{|\delta u|^{2}}{2\tilde{u}}d\mathbf{x}d\mathbf{k}-\int_{\Omega\times\mathbb{S}^{d-1}}\left[\mathbf{k}\cdot\nabla\frac{1}{\tilde{u}}\right]\frac{|\delta u|^{2}}{2}d\mathbf{x}d\mathbf{k}\,, (55)

and the identity

𝐤1u~=1u~2zu~+Σa~(u~)u~+Σs(𝒦)u~|u~|2.\mathbf{k}\cdot\nabla\frac{1}{\tilde{u}}=\frac{1}{\tilde{u}^{2}}\partial_{z}\tilde{u}+\frac{\widetilde{\Sigma_{a}}(\langle{\tilde{u}}\rangle)}{\tilde{u}}+\frac{\Sigma_{s}(\mathcal{I}-\mathcal{K})\tilde{u}}{|\tilde{u}|^{2}}\,. (56)

Using the fact that g=g~g=\widetilde{g}, we can also conclude that

δu=0.\langle{\delta u}\rangle=0\,. (57)

It is also straightforward to check that δu\delta u solves the transport equation:

zδu+𝐤δu+Σa(u)δu+Σs(𝒦)δu=(Σ~a(u~)Σa(u))u~,\partial_{z}\delta u+\mathbf{k}\cdot\nabla\delta u+{\Sigma_{a}}(\langle{{u}}\rangle)\delta u+\Sigma_{s}(\mathcal{I}-\mathcal{K})\delta u=(\widetilde{\Sigma}_{a}(\langle{\tilde{u}}\rangle)-\Sigma_{a}(\langle{{u}}\rangle))\tilde{u}\,, (58)

with zero initial and incoming boundary conditions. We multiply this equation by δuu~\frac{\delta u}{\tilde{u}} and integrate over Ω×𝕊d1\Omega\times\mathbb{S}^{d-1} to obtain

Ω×𝕊d1(zδu)δuu~𝑑𝐱𝑑𝐤+Ω×𝕊d1𝐤|δu|22u~d𝐱d𝐤Ω×𝕊d1|δu|22|u~|2zu~d𝐱d𝐤\displaystyle\int_{\Omega\times\mathbb{S}^{d-1}}(\partial_{z}\delta u)\frac{\delta u}{\tilde{u}}d\mathbf{x}d\mathbf{k}+\int_{\Omega\times\mathbb{S}^{d-1}}\mathbf{k}\cdot\nabla\frac{|\delta u|^{2}}{2\tilde{u}}d\mathbf{x}d\mathbf{k}-\int_{\Omega\times\mathbb{S}^{d-1}}\frac{|\delta u|^{2}}{2|\tilde{u}|^{2}}\partial_{z}\tilde{u}d\mathbf{x}d\mathbf{k} (59)
Ω×𝕊d1Σa~(u~)+Σs2u~|δu|2𝑑𝐱𝑑𝐤+Ω×𝕊d1(Σs𝒦u~)|δu|2|u~|2𝑑𝐱𝑑𝐤\displaystyle-\int_{\Omega\times\mathbb{S}^{d-1}}\frac{\widetilde{\Sigma_{a}}(\langle{\tilde{u}}\rangle)+\Sigma_{s}}{2\tilde{u}}|\delta u|^{2}d\mathbf{x}d\mathbf{k}+\int_{\Omega\times\mathbb{S}^{d-1}}\frac{(\Sigma_{s}\mathcal{K}\tilde{u})|\delta u|^{2}}{|\tilde{u}|^{2}}d\mathbf{x}d\mathbf{k}
+Ω×𝕊d1Σa(u)u~|δu|2𝑑𝐱𝑑𝐤+ΣsΩ×𝕊d1[(𝒦)δu]δuu~𝑑𝐱𝑑𝐤\displaystyle+\int_{\Omega\times\mathbb{S}^{d-1}}\frac{\Sigma_{a}(\langle{u}\rangle)}{\tilde{u}}|\delta u|^{2}d\mathbf{x}d\mathbf{k}+\Sigma_{s}\int_{\Omega\times\mathbb{S}^{d-1}}\left[(\mathcal{I}-\mathcal{K})\delta u\right]\frac{\delta u}{\tilde{u}}d\mathbf{x}d\mathbf{k}
=Ω×𝕊d1(Σa~(u~)Σa(u))δu𝑑𝐱𝑑𝐤.\displaystyle=\int_{\Omega\times\mathbb{S}^{d-1}}(\widetilde{\Sigma_{a}}(\langle{\tilde{u}}\rangle)-\Sigma_{a}(\langle{{u}}\rangle))\delta ud\mathbf{x}d\mathbf{k}.

where we have used the identities (55) and  (56).

We first observe that since Σa(u)\Sigma_{a}(\langle{{u}}\rangle) and Σa~(u~)\widetilde{\Sigma_{a}}(\langle{\tilde{u}}\rangle) do not depend on 𝐤\mathbf{k}, the right hand side of (59) vanishes due to (57).

To handle the left hand side of (59), we observe that

Ω×𝕊d1𝐤|δu|22u~d𝐱d𝐤=Γ+𝐤𝐧|δu|22u~0,\int_{\Omega\times\mathbb{S}^{d-1}}\mathbf{k}\cdot\nabla\frac{|\delta u|^{2}}{2\tilde{u}}d\mathbf{x}d\mathbf{k}=\int_{\Gamma_{+}}\mathbf{k}\cdot\mathbf{n}\frac{|\delta u|^{2}}{2\tilde{u}}\geq 0\,, (60)
(zδu)δuu~=12z[|δu|2u~]+12|δu|2|u~|2zu~,(\partial_{z}\delta u)\frac{\delta u}{\tilde{u}}=\frac{1}{2}\partial_{z}\left[\frac{|\delta u|^{2}}{\tilde{u}}\right]+\frac{1}{2}\frac{|\delta u|^{2}}{|\tilde{u}|^{2}}\partial_{z}\tilde{u}\,, (61)

and

ΣsΩ×𝕊d1[(𝒦)δu]δuu~𝑑𝐱𝑑𝐤=ΣsΩ×𝕊d1|δu|2u~𝑑𝐱𝑑𝐤ΣsΩ×𝕊d1(𝒦δu)δuu~𝑑𝐱𝑑𝐤.\Sigma_{s}\int_{\Omega\times\mathbb{S}^{d-1}}\left[(\mathcal{I}-\mathcal{K})\delta u\right]\frac{\delta u}{\tilde{u}}d\mathbf{x}d\mathbf{k}=\Sigma_{s}\int_{\Omega\times\mathbb{S}^{d-1}}\frac{|\delta u|^{2}}{\tilde{u}}d\mathbf{x}d\mathbf{k}-\Sigma_{s}\int_{\Omega\times\mathbb{S}^{d-1}}(\mathcal{K}\delta u)\frac{\delta u}{\tilde{u}}d\mathbf{x}d\mathbf{k}\,. (62)

We can also prove the following inequality (see Appendix A),

Ω×𝕊d1(𝒦δu)δuu~𝑑𝐱𝑑𝐤Ω×𝕊d1(𝒦u~)|δu|2|u~|2𝑑𝐱𝑑𝐤+νd1κ24Ω×𝕊d1|δu|2u~𝑑𝐱𝑑𝐤,\int_{\Omega\times\mathbb{S}^{d-1}}(\mathcal{K}\delta u)\frac{\delta u}{\tilde{u}}d\mathbf{x}d\mathbf{k}\leq\int_{\Omega\times\mathbb{S}^{d-1}}(\mathcal{K}\tilde{u})\frac{|\delta u|^{2}}{|\tilde{u}|^{2}}d\mathbf{x}d\mathbf{k}+\nu_{d-1}\frac{\kappa^{2}}{4}\int_{\Omega\times\mathbb{S}^{d-1}}\frac{|\delta u|^{2}}{\tilde{u}}d\mathbf{x}d\mathbf{k}, (63)

where κ=(θ¯θ¯)2θ¯\kappa=\frac{(\overline{\theta}-\underline{\theta})}{2\sqrt{\underline{\theta}}}, the constants θ¯\overline{\theta} and θ¯\underline{\theta} are defined in (29).

Let M(z,𝐱):=Σa(u)Σa~(u~)2+Σs(2νd1κ24)M(z,\mathbf{x}):=\Sigma_{a}(\langle{u}\rangle)-\frac{\widetilde{\Sigma_{a}}(\langle{\tilde{u}}\rangle)}{2}+\Sigma_{s}\left(\frac{2-\nu_{d-1}\kappa^{2}}{4}\right). We can then deduce from (59), using (60), (61), (62), and (63), that

12zΩ×𝕊d1|δu|2u~𝑑𝐱𝑑𝐤+Ω×𝕊d1M(z,𝐱)|δu|2u~𝑑𝐱𝑑𝐤0.\displaystyle\frac{1}{2}\partial_{z}\int_{\Omega\times\mathbb{S}^{d-1}}\frac{|\delta u|^{2}}{\tilde{u}}d\mathbf{x}d\mathbf{k}+\int_{\Omega\times\mathbb{S}^{d-1}}M(z,\mathbf{x})\frac{|\delta u|^{2}}{\tilde{u}}d\mathbf{x}d\mathbf{k}\leq 0. (64)

Since the coefficients Σa,l\Sigma_{a,l} and Σs\Sigma_{s} are finite and both u,u~\langle{u}\rangle,\langle{\tilde{u}}\rangle are bounded by fL(X0)\|f\|_{L^{\infty}(X_{0})}, there exists a constant M¯=inf(0,T)×ΩM(z,𝐱)\underline{M}=\inf_{(0,T)\times\Omega}M(z,\mathbf{x}) that

12zΩ×𝕊d1|δu|2u~𝑑𝐱𝑑𝐤+M¯Ω×𝕊d1|δu|2u~𝑑𝐱𝑑𝐤0.\frac{1}{2}\partial_{z}\int_{\Omega\times\mathbb{S}^{d-1}}\frac{|\delta u|^{2}}{\tilde{u}}d\mathbf{x}d\mathbf{k}+\underline{M}\int_{\Omega\times\mathbb{S}^{d-1}}\frac{|\delta u|^{2}}{\tilde{u}}d\mathbf{x}d\mathbf{k}\leq 0.

Then by the Grönwall inequality and the initial condition δu=0\delta u=0 at z=0z=0, we must have that δu0\delta u\equiv 0. Therefore u=u~u=\tilde{u}, which implies that

Σa(u)=zu+𝐤u+Σs(𝒦)uu=zu~+𝐤u~+Σs(𝒦)u~u~=Σa~(u~).\Sigma_{a}(\langle{u}\rangle)=-\frac{\partial_{z}u+\mathbf{k}\cdot\nabla u+\Sigma_{s}(\mathcal{I}-\mathcal{K})u}{u}=-\frac{\partial_{z}\tilde{u}+\mathbf{k}\cdot\nabla\tilde{u}+\Sigma_{s}(\mathcal{I}-\mathcal{K})\tilde{u}}{\tilde{u}}=\widetilde{\Sigma_{a}}(\langle{\tilde{u}}\rangle)\,. (65)

The proof is complete. ∎

The following corollary is a direct result of the comparison principle and Theorem 4.1.

Corollary 4.2.

Under the assumption of Theorem 4.1, the coefficients Σa,l\Sigma_{a,l} can be uniquely determined with finitely many data sets uj\langle{u_{j}}\rangle, j=1,2,,L+1j=1,2,\dots,L+1 if the initial conditions satisfy 0<f1<f2<<fL+10<f_{1}<f_{2}<\dots<f_{L+1} on Ω\Omega.

5 Concluding remarks

This work describes the derivation of semilinear radiative transport models for wave propagation in highly-scattering media with nonlinear absorption. While the technical aspects of the derivation are relatively standard, we believe that our work provides a theoretical justification for the semilinear radiative transport models, as well as their diffusion approximations, used in applications such as multi-photon imaging [10, 36, 44, 45, 51, 55, 56, 57].

As we have remarked before, one concrete example of the quadratic absorption we considered here is two-photon absorption in nonlinear optics [38, 48]. The radiative transport equation we derive for this case, equation (20), is different from the two-photon radiative transport equation of [40] where the phase space intensity corresponds to a two-photon entangled state of light, not two-photon absorption.

The calculation we have presented here does not generalize to media with nonlinearities such as those that arise in the Kerr effect and second harmonic generation [3, 29, 46, 52]. The derivation of transport equations for wave propagation in such media is a topic of great interest, but is much more challenging due to the richness of the behavior of the corresponding wave equations [17, 18, 50]. We point to the derivation in [22] in the context of the nonlinear Schrödinger equation within the mean-field approximation, and leave further investigations in this direction to future work. We note that the acousto-optic effect, in which light undergoes a frequency shift due to interaction with an acoustic wave, has also been studied in random media. Although this effect is not nonlinear in the sense considered in this paper, it is possible to develop a suitable kinetic model and associated radiative transport equations [31].

Acknowledgment

The work of JK is supported in part by the Simons Foundation Math + X Investigator Award #376319 (to Michael I. Weinstein). KR is partially supported by the National Science Foundation through grants DMS-1913309 and DMS-1937254. JCS acknowledges support from the NSF Grant DMS-1912821 and the AFOSR Grant FA9550-19-1-0320.

Appendix A Proof of inequality (63)

Proof.

Using the AM-GM inequality, we deduce that

Ω×𝕊d1(𝒦δu)δuu~𝑑𝐱𝑑𝐤Ω×𝕊d1(𝒦u~)|δu|2|u~|2𝑑𝐱𝑑𝐤+14Ω×𝕊d1|𝒦δu𝒦u~|2𝑑𝐱𝑑𝐤.\int_{\Omega\times\mathbb{S}^{d-1}}(\mathcal{K}\delta u)\frac{\delta u}{\tilde{u}}d\mathbf{x}d\mathbf{k}\leq\int_{\Omega\times\mathbb{S}^{d-1}}(\mathcal{K}\tilde{u})\frac{|\delta u|^{2}}{|\tilde{u}|^{2}}d\mathbf{x}d\mathbf{k}+\frac{1}{4}\int_{\Omega\times\mathbb{S}^{d-1}}\left|\frac{\mathcal{K}\delta u}{\sqrt{\mathcal{K}\tilde{u}}}\right|^{2}d\mathbf{x}d\mathbf{k}.

Since θ¯p(𝐤,𝐤)θ¯\underline{\theta}\leq p(\mathbf{k},\mathbf{k}^{\prime})\leq\overline{\theta} and δu=0\langle{\delta u}\rangle=0, we have the preliminary estimate

|𝒦δu𝒦u~|=|𝒦δuθ¯+θ¯2δu𝒦u~|(θ¯θ¯)|δu|2θ¯u~.\left|\frac{\mathcal{K}\delta u}{\sqrt{\mathcal{K}\tilde{u}}}\right|=\left|\frac{\mathcal{K}\delta u-\frac{\overline{\theta}+\underline{\theta}}{2}\langle{\delta u}\rangle}{\sqrt{\mathcal{K}\tilde{u}}}\right|\leq\frac{(\overline{\theta}-\underline{\theta})\langle{|\delta u|}\rangle}{2\sqrt{\underline{\theta}\langle{\tilde{u}}\rangle}}.

Denote by κ\kappa the constant κ:=(θ¯θ¯)2θ¯\kappa:=\frac{(\overline{\theta}-\underline{\theta})}{2\sqrt{\underline{\theta}}}. We then have

Ω×𝕊d1|𝒦δu𝒦u~|2𝑑𝐱𝑑𝐤κ2Ω×𝕊d1|δu|2u~𝑑𝐱𝑑𝐤=νd1κ2Ω|δu|2u~𝑑𝐱.\int_{\Omega\times\mathbb{S}^{d-1}}\left|\frac{\mathcal{K}\delta u}{\sqrt{\mathcal{K}\tilde{u}}}\right|^{2}d\mathbf{x}d\mathbf{k}\leq\kappa^{2}\int_{\Omega\times\mathbb{S}^{d-1}}\frac{\langle{|\delta u|}\rangle^{2}}{{\langle{\tilde{u}}\rangle}}d\mathbf{x}d\mathbf{k}=\nu_{d-1}\kappa^{2}\int_{\Omega}\frac{\langle{|\delta u|}\rangle^{2}}{{\langle{\tilde{u}}\rangle}}d\mathbf{x}.

Using the Cauchy-Schwartz inequality, we arrive at

𝕊d1|δu(𝐱,𝐤)|2u~(𝐱,𝐤)𝑑𝐤𝕊d1u~(𝐱,𝐤)𝑑𝐤(𝕊d1|δu(𝐱,𝐤)|𝑑𝐤)2=|δu|2.\int_{\mathbb{S}^{d-1}}\frac{|\delta u(\mathbf{x},\mathbf{k})|^{2}}{\tilde{u}(\mathbf{x},\mathbf{k})}d\mathbf{k}\int_{\mathbb{S}^{d-1}}\tilde{u}(\mathbf{x},\mathbf{k})d\mathbf{k}\geq\left(\int_{\mathbb{S}^{d-1}}|\delta u(\mathbf{x},\mathbf{k})|d\mathbf{k}\right)^{2}=\langle{|\delta u|}\rangle^{2}\,.

Therefore, we have

|δu|2u~𝕊d1|δu(𝐱,𝐤)|2u~(𝐱,𝐤)𝑑𝐤.\frac{\langle{|\delta u|}\rangle^{2}}{{\langle{\tilde{u}}\rangle}}\leq\int_{\mathbb{S}^{d-1}}\frac{|\delta u(\mathbf{x},\mathbf{k})|^{2}}{\tilde{u}(\mathbf{x},\mathbf{k})}d\mathbf{k}.

This implies that

Ω×𝕊d1(𝒦δu)δuu~𝑑𝐱𝑑𝐤Ω×𝕊d1(𝒦u~)|δu|2|u~|2𝑑𝐱𝑑𝐤+νd1κ24Ω×𝕊d1|δu|2u~𝑑𝐱𝑑𝐤.\int_{\Omega\times\mathbb{S}^{d-1}}(\mathcal{K}\delta u)\frac{\delta u}{\tilde{u}}d\mathbf{x}d\mathbf{k}\leq\int_{\Omega\times\mathbb{S}^{d-1}}(\mathcal{K}\tilde{u})\frac{|\delta u|^{2}}{|\tilde{u}|^{2}}d\mathbf{x}d\mathbf{k}+\nu_{d-1}\frac{\kappa^{2}}{4}\int_{\Omega\times\mathbb{S}^{d-1}}\frac{|\delta u|^{2}}{\tilde{u}}d\mathbf{x}d\mathbf{k}.

This completes the proof. ∎

References

  • [1] J.-L. Akian and E. Savin, Kinetic modeling of multiple scattering of acoustic waves in randomly heterogeneous flows, Multiscale Modeling & Simulation, 19 (2021), pp. 1394–1424.
  • [2] H. Ammari, E. Bossy, J. Garnier, W. Jing, and L. Seppecher, Radiative transfer and diffusion limits for wave field correlations in locally shifted random media, J. Math. Phys., 54 (2013). 021501.
  • [3] Y. M. Assylbekov and T. Zhou, Direct and inverse problems for the nonlinear time-harmonic Maxwell equations in Kerr-type media, J. Spectr. Theor., (2020).
  • [4] G. Bal, Kinetics of scalar wave fields in random media, Wave Motion, 43 (2005), pp. 132–157.
  • [5] G. Bal, T. Komorowski, and L. Ryzhik, Kinetic limits for waves in a random medium, Kinet. Relat. Models, 3 (2010), pp. 529–644.
  • [6] G. Bal, G. Papanicolaou, and L. Ryzhik, Radiative transport limit for the random Schrödinger equation, Nonlinearity, 15 (2002), pp. 513–529.
  • [7] G. Bal and O. Pinaud, Accuracy of transport models for waves in random media, Wave Motion, 43 (2006).
  • [8]  , Kinetic models for imaging in random media, SIAM Multiscale Model. Simul., 6 (2007), pp. 792–819.
  • [9] G. Bal and K. Ren, Transport-based imaging in random media, SIAM J. Appl. Math., 68 (2008), pp. 1738–1762.
  • [10] P. Bardsley, K. Ren, and R. Zhang, Quantitative photoacoustic imaging of two-photon absorption, J. Biomed. Opt., 23 (2018). 016002.
  • [11] I. Baydoun, É. Savin, R. Cottereau, D. Clouteau, and J. Guilleminot, Kinetic modeling of multiple scattering of elastic waves in heterogeneous anisotropic media, Wave Motion, 51 (2014), pp. 1325–1348.
  • [12] L. Borcea and J. Garnier, Derivation of a one-way radiative transfer equation in random media, Phys. Rev. E, 93 (2016). 022115.
  • [13] L. Borcea, J. Garnier, and K. Solna, Wave propagation and imaging in moving random media, Multiscale Modeling & Simulation, 17 (2019), pp. 31–67.
  • [14] R. W. Boyd, Nonlinear optics, Academic press, 2020.
  • [15] R. Carminati and J. C. Schotland, Principles of Scattering and Transport of Light, Cambridge University Press, 2021.
  • [16] S. Chen and Q. Li, Semiclassical limit of an inverse problem for the Schrödinger equation, Research in the Mathematical Sciences, 8 (2021), pp. 1–18.
  • [17] C. Conti, L. Angelani, and G. Ruocco, Light diffusion and localization in three-dimensional nonlinear disordered media, Phys. Rev. A, 75 (2007). 033812.
  • [18] C. Conti and L. Leuzzi, Complexity of waves in nonlinear disordered media, Phys. Rev. B, 83 (2011). 134204.
  • [19] R. Dautray and J.-L. Lions, Mathematical Analysis and Numerical Methods for Science and Technology, Vol VI, Springer-Verlag, Berlin, 1993.
  • [20] W. Denk, J. Strickler, and W. Webb, Two photon laser scanning fluorescence microscopy, Science, 248 (1990), pp. 73–76.
  • [21] L. Erdös and H. T. Yau, Linear Boltzmann equation as the weak coupling limit of a random Schrödinger Equation, Comm. Pure Appl. Math., 53(6) (2000), pp. 667–735.
  • [22] A. Fannjiang, G. Papanicolaou, and S. Jin, High frequency behavior of the focusing nonlinear Schrödinger equation with random inhomogeneities, SIAM J. Appl. Math., 63 (2003), pp. 1328–1358.
  • [23] A. Fannjiang and L. V. Ryzhik, Radiative transfer of sound waves in a random flow: Turbulent scattering, straining, and mode-coupling, SIAM J. Appl. Math., 61 (2001), pp. 1545–1577.
  • [24] A. C. Fannjiang, Radiative transfer limit of two-frequency Wigner distribution for random parabolic waves: an exact solution, Comptes Rendus Physique, 8 (2007), pp. 267–271.
  • [25] J. Garnier and K. Solna, Effective transport equations and enhanced backscattering in random waveguides, SIAM J. Appl. Math., 68 (2008), pp. 1574–1599.
  • [26] D. Gilbarg and N. S. Trudinger, Elliptic Partial Differential Equations of Second Order, Springer-Verlag, Berlin, 2000.
  • [27] F. Golse, P.-L. Lions, B. Perthame, and R. Sentis, Regularity of the moments of the solution of a transport equation, J. Func. Anal., 76 (1988), pp. 110–125.
  • [28] C. Gomez, Radiative transport limit for the random schrödinger equation with long-range correlations, J. Math. Pures Appl., 98 (2012), pp. 295–327.
  • [29] A. Goy and D. Psaltis, Digital reverse propagation in focusing Kerr media, Phys. Rev. A, 83 (2011). 031802.
  • [30] M. Guo and X.-P. Wang, Transport equations for a general class of evolution equations with random perturbations, J. Math. Phys., 40 (1999), pp. 4828–4848.
  • [31] J. Hoskins and J. Schotland, Acousto-optic effect in random media, Phys. Rev. E, 95 (2017). 033002.
  • [32] A. Ishimaru, Wave Propagation and Scattering in Random Media, IEEE Press, 1997.
  • [33] Y. Jing and N. Xiang, One-dimensional transport equation models for sound energy propagation in long spaces: Simulations and experiments, J. Acoust. Soc. Am., 127 (2010), pp. 2323–2331.
  • [34] R. Kellogg, Uniqueness in the schauder fixed point theorem, Proceedings of the American Mathematical Society, 60 (1976), pp. 207–210.
  • [35] J. Kraisler and J. C. Schotland, Collective spontaneous emission and kinetic equations for one-photon light in random media, J. Math. Phys., 63 (2022). 031901.
  • [36] R.-Y. Lai, K. Ren, and T. Zhou, Inverse transport and diffusion problems in photoacoustic imaging with nonlinear absorption, SIAM J. Appl. Math., 82 (2022), pp. 602–624. arXiv:2107.08118.
  • [37] J. Lukkarinen and H. Spohn, Kinetic limit for wave propagation in a random medium, Arch. Ration. Mech. Anal., 183 (2007), pp. 93–162.
  • [38] H. Mahr, Two-photon absorption spectroscopy, in Quantum Electronics: A Treatise, H. Rabin and C. L. Tang, eds., Academic Press, 2012.
  • [39] L. Margerin, Introduction to radiative transfer of seismic waves, in Seismic Earth: Array Analysis of Broadband Seismograms, A. Levander and G. Nolet, eds., American Geophysical Union, Washington, DC, 2005.
  • [40] V. A. Markel and J. C. Schotland, Radiative transport for two-photon light, Phys. Rev. A, 90 (2014). 033815.
  • [41] M. I. Mishchenko, L. D. Travis, and A. A. Lacis, Multiple scattering of light by particles: radiative transfer and coherent back scattering, Cambridge University Press, 2006.
  • [42] P. P. Mondal, G. Vicidomini, and A. Diaspro, Image reconstruction for multiphoton fluorescence microscopy, Appl. Phys. Lett., 92 (2008). 103902.
  • [43] A. Pazy, Semigroups of Linear Operators and Applications to Partial Differential Equations, Springer Science & Business Media, 2012.
  • [44] K. Ren and R. Zhang, Nonlinear quantitative photoacoustic tomography with two-photon absorption, SIAM J. Appl. Math., 78 (2018), pp. 479–503.
  • [45] K. Ren and Y. Zhong, Unique determination of absorption coeffients in a semilinear transport equation, SIAM J. Math. Anal., 53 (2021), pp. 5158–5184. arXiv:2007.09516.
  • [46] M.-L. Ren and Z.-Y. Li, Exact iterative solution of second harmonic generation in quasi-phase-matched structures, Optics Express, 18 (2010), pp. 7288–7299.
  • [47] L. Roux, P. Mareschal, N. Vukadinovic, J.-B. Thibaud, and J.-J. Greffet, Scattering by a slab containing randomly located cylinders: comparison between radiative transfer and electromagnetic simulation, J. Opt. Soc. Am. A, 18 (2001), pp. 374–384.
  • [48] M. Rumi and J. W. Perry, Two-photon absorption: an overview of measurements and principles, Adv. Opt. Photon., 2 (2010), pp. 451–518.
  • [49] L. Ryzhik, G. Papanicolaou, and J. B. Keller, Transport equations for elastic and other waves in random media, Wave Motion, 24 (1996), pp. 327–370.
  • [50] S. E. Skipetrov, Langevin description of speckle dynamics in nonlinear disordered media, Phys. Rev. E, 67 (2003). 016601.
  • [51] P. Stefanov and Y. Zhong, Inverse boundary problem for the two photon absorption transport equation, SIAM J. Math. Anal., 54 (2022), pp. 2753–2767.
  • [52] T. Szarvas and Z. Kis, Numerical simulation of nonlinear second harmonic wave generation by the finite difference frequency domain method, J. Opt. Soc. Am. B, 35 (2018), pp. 731–740.
  • [53] F. Voit, J. Schäfer, and A. Kienle, Light scattering by multiple spheres: comparison between Maxwell theory and radiative-transfer-theory calculations, Opt. Lett., 34 (2009), pp. 2593–2595.
  • [54] D. Xu, X. Wang, Z. Xu, W. Zhou, and J. Yin, A super-resolution reconstruction algorithm for two-photon fluorescence polarization microscopy, Optics Communications, 499 (2021). 127116.
  • [55] Y. Yamaoka, M. Nambu, and T. Takamatsu, Fine depth resolution of two-photon absorption-induced photoacoustic microscopy using low-frequency bandpass filtering, Optics Express, 19 (2011), pp. 13365–13377.
  • [56] C. S. Yelleswarapu and S. R. Kothapalli, Nonlinear photoacoustics for measuring the nonlinear optical absorption coefficient, Optics Express, 18 (2010), pp. 9020–9025.
  • [57] J. Ying, F. Liu, and R. R. Alfano, Spatial distribution of two-photon-excited fluorescence in scattering media, Appl. Opt., 38 (1999), pp. 224–229.