17.08.2013 Views

Transport Theory for Light Propagation in Tissues*

Transport Theory for Light Propagation in Tissues*

Transport Theory for Light Propagation in Tissues*

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Transport</strong> <strong>Theory</strong> <strong>for</strong> <strong>Light</strong> <strong>Propagation</strong> <strong>in</strong> <strong>Tissues*</strong><br />

Arnold D. Kim<br />

Department of Mathematics, Stan<strong>for</strong>d University, USA<br />

ABSTRACT<br />

We review the basic theory <strong>for</strong> the radiative transport equation govern<strong>in</strong>g light propagation <strong>in</strong> biological tissues.<br />

The Green’s function is the fundamental solution to the transport equation from which all other solutions can<br />

be computed. We compute the Green’s function as an analytical expansion <strong>in</strong> plane wave modes. We calculate<br />

these plane wave modes numerically us<strong>in</strong>g the discrete ord<strong>in</strong>ate method. We use the Green’s function to compute<br />

the po<strong>in</strong>t spread function <strong>in</strong> a half space composed of a uni<strong>for</strong>m scatter<strong>in</strong>g and absorb<strong>in</strong>g medium.<br />

1. INTRODUCTION<br />

The radiative transport equation governs light propagation <strong>in</strong> biological tissue. 1 Hence, solutions of it provide<br />

fundamental <strong>in</strong>sight <strong>in</strong>to the <strong>in</strong>teraction between light and tissue. The Green’s function is the fundamental<br />

solution to the transport equation from which all other solutions can be computed. This theory <strong>for</strong> the radiative<br />

transport equation is well known. 2, 3 However, the Green’s function is not known except <strong>for</strong> relatively simple<br />

situations.<br />

We shall describe a method to compute the Green’s function <strong>for</strong> the transport equation. It is given analytically<br />

as an expansion <strong>in</strong> plane wave modes. Because plane wave solutions are not known analytically, we calculate<br />

them numerically us<strong>in</strong>g a discrete ord<strong>in</strong>ate method. Kim and Keller 4 have given several results regard<strong>in</strong>g plane<br />

wave solutions <strong>for</strong> the transport equation. In particular, they have identified an important symmetry property.<br />

Kim 5 has extended those results by identify<strong>in</strong>g the orthogonality of plane wave modes. We shall use both the<br />

symmetry and orthogonal properties of plane wave modes to compute the Green’s function.<br />

We shall demonstrate the use of the Green’s function <strong>for</strong> the transport equation by comput<strong>in</strong>g the po<strong>in</strong>t<br />

spread function <strong>in</strong> a half space. The half space is composed of a uni<strong>for</strong>m scatter<strong>in</strong>g and absorb<strong>in</strong>g medium.<br />

For an optically thick medium such as biological tissue, the diffusion approximation is often used. 1 It assumes<br />

that light undergoes sufficient multiple scatter<strong>in</strong>g that the radiance becomes isotropic. We shall show that the<br />

po<strong>in</strong>t spread function posseses a non-trivial direction dependence even <strong>for</strong> sources located deep <strong>in</strong> the half space.<br />

Hence, the key assumption <strong>in</strong> the diffusion approximation is not valid <strong>for</strong> the po<strong>in</strong>t spread function.<br />

We review the radiative transport equation <strong>for</strong> light propagation <strong>in</strong> biological tissue <strong>in</strong> Section 2. In Section<br />

3 we discuss plane wave modes and the numerical method used to compute them. In Section 4 we compute<br />

the Green’s function as an expansion <strong>in</strong> plane wave modes. We derive the po<strong>in</strong>t spread function <strong>in</strong> a half space<br />

composed of a uni<strong>for</strong>m scatter<strong>in</strong>g and absorb<strong>in</strong>g medium and present results of it <strong>in</strong> Section 5. We present our<br />

conclusions <strong>in</strong> Section 6.<br />

2. THE RADIATIVE TRANSPORT EQUATION<br />

The time-<strong>in</strong>dependent radiance Ψ is the radiant power per unit solid angle per unit area perpendicular to the<br />

direction of propagation. It depends on a position vector r and a unit direction vector ω. The radiative transport<br />

equation<br />

ω · ∇Ψ + σaΨ + σsLΨ = Q (1)<br />

* This paper is an abridgement of a paper by the authors with the same title [to appear <strong>in</strong> J. Opt. Soc. Am. A.].<br />

For author <strong>in</strong><strong>for</strong>mation: E-mail: adkim@math.stan<strong>for</strong>d.edu, Telephone: 1-650-723-2975, Fax: 1-650-725-4066, Address:<br />

Department of Mathematics, Stan<strong>for</strong>d University, Stan<strong>for</strong>d, CA 94305-2125, USA


governs Ψ <strong>in</strong> a scatter<strong>in</strong>g and absorb<strong>in</strong>g medium. The absorption and scatter<strong>in</strong>g coefficients are denoted by σa<br />

and σs, respectively. Q denotes a source. The scatter<strong>in</strong>g operator L is def<strong>in</strong>ed as<br />

<br />

LΨ = Ψ(ω, r) −<br />

Ω<br />

f(ω · ω ′ )Ψ(ω ′ , r)dω ′ . (2)<br />

Integration <strong>in</strong> (2) takes place over the unit sphere Ω. The scatter<strong>in</strong>g phase function f gives the fraction of light<br />

scattered <strong>in</strong> direction ω due to light <strong>in</strong>cident <strong>in</strong> direction ω ′ . We assume that it depends only on the cos<strong>in</strong>e of<br />

the scatter<strong>in</strong>g angle ω · ω ′ .<br />

The solution of (1) <strong>in</strong> a spatial doma<strong>in</strong> D with boundary surface S is determ<strong>in</strong>ed uniquely by <strong>in</strong>ternal sources<br />

and the radiance <strong>in</strong>cident on the surface. 2 Hence, (1) is a well-posed problem provided that it is supplemented<br />

with boundary conditions of the <strong>for</strong>m<br />

Ψ(ω, rs) = h(ω, rs), ω ∈ Ω<strong>in</strong>(rs), rs ∈ S. (3)<br />

Ω<strong>in</strong>(rs) is the set of unit direction vectors po<strong>in</strong>t<strong>in</strong>g <strong>in</strong>to D at rs. It is def<strong>in</strong>ed as Ω<strong>in</strong>(rs) = {ω : ω · ˆn(rs) > 0}<br />

with ˆn(rs) denot<strong>in</strong>g the unit <strong>in</strong>ward normal at rs.<br />

The solution to (2) with (3) <strong>in</strong> a doma<strong>in</strong> D with boundary S is given by the general representation <strong>for</strong>mula2, 3 :<br />

<br />

Ψ(ω, r) = G(ω, r; ω ′ , r ′ )Q(ω ′ , r ′ )dω ′ dr ′ <br />

+ G(ω, r; ω ′ , r ′ s) [ω ′ · ˆn(r ′ s)] Ψ(ω ′ , r ′ s)dω ′ dr ′ s. (4)<br />

D<br />

Ω<br />

The Green’s function G(ω, r; ω ′ , r ′ ) is the solution of (1) <strong>in</strong> the whole space with Q(ω, r) = δ(ω − ω ′ )δ(r − r ′ )<br />

that is bounded <strong>for</strong> all r = r ′ . The surface <strong>in</strong>tegral term <strong>in</strong> (4) conta<strong>in</strong>s the Green’s function <strong>for</strong> a source located<br />

at the boundary surface. It is <strong>for</strong>mally def<strong>in</strong>ed as the limit<strong>in</strong>g value of the Green’s function as the source location<br />

2, 3<br />

approaches the boundary from with<strong>in</strong> the doma<strong>in</strong>.<br />

The second term <strong>in</strong> (4) requires that the radiance at the boundary surface be known <strong>for</strong> all directions.<br />

However, we know it only <strong>for</strong> directions po<strong>in</strong>t<strong>in</strong>g <strong>in</strong>to the doma<strong>in</strong> from (3). To obta<strong>in</strong> the radiance over the<br />

rema<strong>in</strong><strong>in</strong>g directions, we must solve the surface <strong>in</strong>tegral equation3 <br />

Ψ(ω, rs) = G(ω, rs; ω ′ , r ′ )Q(ω ′ , r ′ )dω ′ dr ′ <br />

+<br />

D<br />

Ω<br />

S<br />

S<br />

Ω<br />

Ω<br />

G(ω, rs; ω ′ , r ′ s) [ω ′ · ˆn(r ′ s)] Ψ(ω ′ , r ′ s)dω ′ dr ′ s. (5)<br />

This <strong>in</strong>tegral equation is <strong>for</strong> the radiance over all directions. Hence it is a coupled system of <strong>in</strong>tegral equations<br />

<strong>for</strong><br />

<br />

Ψ(ω, rs) =<br />

Ψ<strong>in</strong>(ω, rs)<br />

Ψout(ω, rs)<br />

ω ∈ Ω<strong>in</strong>(rs),<br />

ω ∈ Ωout(rs),<br />

(6)<br />

with Ωout(rs) = {ω : ω · ˆn(rs) < 0}. By solv<strong>in</strong>g (5), we determ<strong>in</strong>e Ψ on S <strong>for</strong> all directions. The solution<br />

anywhere <strong>in</strong> D follows from evaluat<strong>in</strong>g (4).<br />

3. PLANE WAVE MODES<br />

We study plane wave mode solutions to the homogeneous transport equation which is (1) with Q = 0. Plane<br />

wave solutions take the <strong>for</strong>m<br />

Ψ(ν, µ, ρ, z) = 1<br />

(2π) 2<br />

<br />

V (ν, µ; κ)e λ(κ)z+iκ·ρ dκ. (7)<br />

R 2<br />

In (7) we express the position vector as r = (ρ, z) and the direction vector as ω = (ν, µ). Substitut<strong>in</strong>g (7) <strong>in</strong>to<br />

(1) and Fourier trans<strong>for</strong>m<strong>in</strong>g that result with respect to ρ yields the eigenvalue problem<br />

λµV + iν · κV + σaV + σsLV = 0. (8)<br />

Both λ and V depend on κ parametrically. A portion of the eigenvalue spectrum is cont<strong>in</strong>uous. 2 In the<br />

discussion that follows, discrete sums are meant to <strong>in</strong>clude <strong>in</strong>tegration over the cont<strong>in</strong>uous spectrum.


3.1. Symmetry<br />

For each pair [λ(κ), V (ν, µ; κ)] solv<strong>in</strong>g (8), the pair [−λ(κ), V (ν, −µ; κ)] also solves it. 4 In light of this symmetry,<br />

we order the eigenvalues by their real part. Furthermore, we <strong>in</strong>dex the eigenvalues so that Re[λj] > 0 <strong>for</strong> j > 0<br />

and Re[λj] < 0 <strong>for</strong> j < 0. Hence, the symmetry property of plane wave modes is given as<br />

3.2. Orthogonality<br />

λ−j = −λj, V−j(ν, µ; κ) = Vj(ν, −µ; κ), j = 1, 2, . . . (9)<br />

For two different pairs [λj, Vj(ν, µ)] and [λk, Vk(ν, µ)] satisfy<strong>in</strong>g<br />

λjµVj + iν · κVj + σaVj + σsLVj = 0, (10a)<br />

λkµVk + iν · κVk + σaVk + σsLVk = 0. (10b)<br />

We multiply (10a) by Vk and (10b) by Vj, <strong>in</strong>tegrate both results over the unit sphere and take the difference.<br />

This manipulation yields<br />

<br />

(λj − λk) Vj(ν, µ)Vk(ν, µ)µdω = 0. (11)<br />

Ω<br />

From (11) we determ<strong>in</strong>e that the plane wave modes are orthogonal with respect to an <strong>in</strong>ner product hav<strong>in</strong>g<br />

weight µ.<br />

Suppose C is the value of the <strong>in</strong>tegral<br />

<br />

Ω<br />

Vj(ν, µ)Vj(ν, µ)µdω = C. (12)<br />

Chang<strong>in</strong>g µ to −µ <strong>in</strong> (12) and us<strong>in</strong>g (9), we determ<strong>in</strong>e that<br />

<br />

V−j(ν, µ)V−j(ν, µ)µdω = −C. (13)<br />

In view of (13) we normalize plane wave modes as<br />

3.3. Completeness<br />

<br />

Ω<br />

Ω<br />

Vj(ν, µ)Vj(ν, µ)µdω =<br />

<br />

+1 <strong>for</strong> j < 0,<br />

−1 <strong>for</strong> j > 0.<br />

It can be shown that plane wave modes are complete over the full range. 2 We now show that half of them are<br />

complete over the half range. Consider the one dimensional problem, homogeneous half space problem<br />

<strong>in</strong> the half space z > 0 with the boundary condition<br />

(14)<br />

µ∂zΨ(ω, z) + σaΨ(ω, z) + σsLΨ(ω, z) = 0, (15)<br />

Ψ(ω, 0) = h(ω), ω · ˆz > 0. (16)<br />

We impose also that the solution is bounded <strong>for</strong> all z > 0. The solution to this problem exists and is unique.<br />

Because of planar symmetry, we need only to consider plane wave modes <strong>for</strong> κ = (0, 0). Us<strong>in</strong>g completeness<br />

of plane wave modes, we write the general solution to (15) as<br />

Ψ(ω, z) = <br />

aje λjz Vj(ω). (17)<br />

j


To impose boundedness <strong>for</strong> all z > 0, we set aj = 0 <strong>for</strong> j > 0 to remove exponentially grow<strong>in</strong>g terms <strong>in</strong> (17).<br />

Impos<strong>in</strong>g the boundary condition (16), we obta<strong>in</strong> the l<strong>in</strong>ear system<br />

<br />

ajVj(ω) = h(ω), ω · ˆz > 0. (18)<br />

j 0}. There<strong>for</strong>e, half of the plane wave modes (those <strong>for</strong> with <strong>in</strong>dices j < 0) are<br />

complete over the half range. A similar analysis can be done to show that the plane wave modes <strong>for</strong> which j > 0<br />

are complete over the hemisphere Ω− = {ω : ω · ˆz < 0}.<br />

3.4. Numerical Method to Calculate Plane Wave Modes<br />

We solve (10) with L def<strong>in</strong>ed by (2) us<strong>in</strong>g the discrete ord<strong>in</strong>ate method. 1 The direction vector ω is given <strong>in</strong><br />

terms of the cos<strong>in</strong>e of the polar angle µ = cos θ and the azimuthal angle ϕ as<br />

<br />

ω = 1 − µ 2 cos ϕ, 1 − µ 2 s<strong>in</strong> ϕ, µ . (19)<br />

We use a product Gaussian quadrature rule 6 to approximate the <strong>in</strong>tegral over the sphere. This method uses an<br />

M-po<strong>in</strong>t Gauss-Legendre quadrature rule <strong>for</strong> µ with abscissas µm and weights wm, and an 2M-po<strong>in</strong>t extended<br />

trapezoid rule with abscissas ϕn = (n − 1)∆ϕ <strong>for</strong> n = 1, . . . , 2M and constant weights ∆ϕ = π/M. Us<strong>in</strong>g these<br />

quadrature rules, we obta<strong>in</strong> the approximation <strong>for</strong> the scatter<strong>in</strong>g operator:<br />

LV (µm, ϕm) ≈ V (µm, ϕn) −<br />

2M<br />

M<br />

n ′ =1 m ′ =1<br />

f(µm, ϕn; µm ′, ϕn ′)V (µm ′, ϕn ′)wm ′∆ϕ. (20)<br />

We use (20) <strong>in</strong> (8) and seek the values of V at the discrete po<strong>in</strong>ts (µm, ϕn). This yields the 2M 2 × 2M 2<br />

matrix eigenvalue problem<br />

λµmV (µm, ϕn) + i 1 − µ 2 m (κx cos ϕn + κy s<strong>in</strong> ϕn) V (µm, ϕn)<br />

Here κ = (κx, κy). Solv<strong>in</strong>g (21) yields 2M 2 eigenvalues.<br />

+ σaV (µm, ϕn) + σsLV (µm, ϕn) = 0, m = 1, . . . , M, n = 1, . . . , 2M. (21)<br />

To normalize the plane wave modes as <strong>in</strong> (14), we use the quadrature rules to compute the normalization<br />

factor<br />

2M M<br />

γj = Vj(µm, ϕn)Vj(µm, ϕn)µmwm∆ϕ. (22)<br />

n=1 m=1<br />

Then we scale the calculated plane wave modes by (−γj) 1/2 <strong>for</strong> j > 0 and (+γj) 1/2 <strong>for</strong> j < 0.<br />

Because of the symmetric quadrature rules used <strong>in</strong> (20), the symmetry given <strong>in</strong> (9) is reta<strong>in</strong>ed so that<br />

λ−j = −λj and V−j(µm, ϕn) = Vj(µM−m+1, ϕn) <strong>for</strong> j = 1, . . . , M 2 . We order the eigenvalues by their real parts,<br />

and <strong>in</strong>dex them as follows:<br />

Re[λ −M 2] < · · · < Re[λ−1] < Re[λ+1] < · · · < Re[λ +M 2]. (23)<br />

We solve (23) <strong>for</strong> each (κx, κy) on an equally spaced grid. This grid is chosen <strong>in</strong> relation to an equally<br />

spaced grid <strong>in</strong> (x, y) <strong>for</strong> use <strong>in</strong> a two dimensional discrete Fourier trans<strong>for</strong>m. Those results are stored <strong>in</strong> physical<br />

memory. In fact, only half of the eigenvalues and eigenvectors (e.g. those <strong>for</strong> which j > 0) are kept. The other<br />

ones can be computed readily us<strong>in</strong>g the symmetric property of plane wave modes.


4. THE GREEN’S FUNCTION<br />

Suppose we have calculated all of the plane wave modes. We now use them to compute the Green’s function.<br />

The Green’s function satisfies<br />

It is translationally <strong>in</strong>variant. Hence, we seek G <strong>in</strong> the <strong>for</strong>m<br />

<br />

where G solves<br />

ω · ∇G + σaG + σsLG = δ(ω − ω ′ )δ(r − r ′ ). (24)<br />

G(ω, r; ω ′ , r ′ ) = 1<br />

(2π) 2<br />

R 2<br />

G(ω, z; ω ′ , z ′ , κ)e iκ·(ρ−ρ′ ) dκ (25)<br />

µ∂z G + iν · κ G + σa G + σsL G = δ(ω − ω ′ )δ(z − z ′ ). (26)<br />

We impose that G is bounded <strong>for</strong> all z = z ′ . By <strong>in</strong>tegrat<strong>in</strong>g (26) about a small <strong>in</strong>terval about z ′ , we obta<strong>in</strong> the<br />

jump condition<br />

µ G(ω, z ′ + 0; ω ′ , z ′ , κ) − µ G(ω, z ′ − 0; ω ′ , z ′ , κ) = δ(ω − ω ′ ). (27)<br />

We seek G as a plane wave mode expansion<br />

Substitut<strong>in</strong>g (28) <strong>in</strong>to (26) and us<strong>in</strong>g (10b), we obta<strong>in</strong><br />

<br />

j<br />

G(ω, z; ω ′ , z ′ , κ) = <br />

gj(z; ω ′ , z ′ , κ)Vj(ω). (28)<br />

j<br />

{[∂zgj(z; ω ′ , z ′ , κ) − λj(κ)gj(z; ω ′ , z ′ , κ)] µVj(ω; κ)} = δ(ω − ω ′ )δ(z − z ′ ). (29)<br />

By orthogonal properties of plane wave modes given <strong>in</strong> (11), we obta<strong>in</strong> n<br />

∂zgj(z; ω ′ , z ′ , κ) − λj(κ)gj(z; ω ′ , z ′ , κ) = −sgn(j)Vj(ω ′ ; κ)δ(z − z ′ ) (30)<br />

with sgn(j) = j/|j|. From (30) we deduce that gj = −sgn(j)Cj(z; z ′ , κ)Vj(ω ′ ) where Cj satisfies<br />

∂zCj(z; z ′ , κ) − λj(κ)Cj(z; z ′ , κ) = δ(z − z ′ ). (31)<br />

We seek the solution to (31) that is bounded <strong>for</strong> all z = z ′ and satisfies<br />

It is readily found to be<br />

There<strong>for</strong>e, G is given by<br />

Cj(z ′ + 0; z ′ , κ) − Cj(z ′ − 0; z ′ , κ) = 1. (32)<br />

⎧<br />

−e λj(κ)(z−z′ ) , z < z ′ , j > 0,<br />

Cj(z; z ′ ⎪⎨<br />

0, z < z<br />

, κ) =<br />

⎪⎩<br />

′ , j < 0,<br />

0, z > z ′ , j > 0,<br />

+e λj(κ)(z−z′ ) , z > z ′ , j < 0.<br />

G(ω, z; ω ′ , z ′ j>0 , κ) =<br />

eλj(κ)(z−z′ ) Vj(ω; κ)Vj(ω ′ ; κ), z < z ′ ,<br />

<br />

j z ′ .<br />

The Green’s function G(ω, r; ω ′ , r ′ ) is recovered by substitut<strong>in</strong>g (34) <strong>in</strong>to (25).<br />

Eq. (34) is approximate because we calculate the plane wave modes numerically. If we knew the plane<br />

wave modes exactly, then (34) would be exact. For the case of isotropic scatter<strong>in</strong>g with planar symmetry, Case<br />

and Zweifel 2 have computed the plane wave modes analytically. Their analysis extends to weakly anisotropic<br />

scatter<strong>in</strong>g phase function. However, this analysis is too difficult to do <strong>for</strong> general scatter<strong>in</strong>g problems. Another<br />

way to <strong>in</strong>terpret (34) with the numerically calculated plane wave modes is that it is the exact Green’s function<br />

<strong>for</strong> the system of differential equation result<strong>in</strong>g from the discrete ord<strong>in</strong>ate method.<br />

(33)<br />

(34)


5. THE POINT SPREAD FUNCTION<br />

We use the Green’s function to compute the po<strong>in</strong>t spread function <strong>in</strong> a half space composed of a uni<strong>for</strong>m scatter<strong>in</strong>g<br />

and absorb<strong>in</strong>g medium. The po<strong>in</strong>t spread function is the radiance exit<strong>in</strong>g the half space z > 0 at the boundary<br />

z = 0 due to a unit source <strong>in</strong> position and direction located <strong>in</strong>side the half space.<br />

We wish to solve<br />

<strong>in</strong> the half space z > 0 subject to the boundary condition<br />

The po<strong>in</strong>t spread function is Ψ(ω, ρ, 0) <strong>for</strong> ω · ˆz < 0.<br />

ω · ∇Ψ + σaΨ + σsLΨ = δ(ρ)δ(z − z ′ )δ(ω − ω ′ ) (35)<br />

Ψ(ω, ρ, z = 0) = 0, ω · ˆz > 0, ρ ∈ R 2 . (36)<br />

The Green’s function solves (35), but does not satisfy (36). Hence, we seek Ψ <strong>in</strong> the <strong>for</strong>m<br />

Ψ = G − Y (37)<br />

where Y is a bounded and regular solution of the homogeneous problem <strong>in</strong> the half space. We impose (36) us<strong>in</strong>g<br />

(37) and determ<strong>in</strong>e that<br />

Y (ω, ρ, 0; ω ′ , r ′ ) = G(ω, ρ, 0; ω ′ , r ′ ), ω · ˆz > 0. (38)<br />

Hence, Y satisfies the homogeneous half space problem with boundary condition (38). From the discussion on<br />

the half-range completeness of plane wave modes, we know that Y takes the <strong>for</strong>m<br />

<br />

with<br />

Y (ω, r; ω ′ , r ′ ) = 1<br />

(2π) 2<br />

R 2<br />

Y (ω, z; ω ′ , z ′ , κ)e iκ·ρ dκ. (39)<br />

Y (ω, z; ω ′ , z ′ , κ) = <br />

yj(ω ′ , z ′ ; κ)e λj(κ)z Vj(ω; κ). (40)<br />

j 0. (41)<br />

G(ω, 0; ω ′ , z ′ , κ) = <br />

j>0<br />

e −λj(κ)z′<br />

Vj(ω; κ)Vj(ω ′ ; κ). (42)<br />

Eq. (42) must hold <strong>for</strong> all ω ′ , but not <strong>for</strong> all ω. Hence, we seek yj as the plane wave mode expansion<br />

yj(ω ′ , z ′ ; κ) = <br />

k<br />

′<br />

−λk(κ)z<br />

djk(κ)e Vk(ω ′ ; κ) (43)<br />

Substitut<strong>in</strong>g (43) and (42) <strong>in</strong>to (41), multiply<strong>in</strong>g by µ ′ and <strong>in</strong>tegrat<strong>in</strong>g over Ω yields the l<strong>in</strong>ear system<br />

<br />

djk(κ)Vj(ω; κ) = Vk(ω; κ), ω · ˆz > 0, k > 0. (44)<br />

j


5.1. Numerical Result<br />

We have calculated the po<strong>in</strong>t spread function us<strong>in</strong>g the radiative transport equation. For these calculations we<br />

have used the Henyey-Greenste<strong>in</strong> phase function<br />

f(ω · ω ′ ) = 1 1 − g<br />

4π<br />

2<br />

(1 + g2 − 2gω · ω ′ ) 3/2<br />

with g denot<strong>in</strong>g the asymmetry parameter. The half space is composed of a medium with σa/σs = 0.01 and<br />

g = 0.85. We have computed the plane wave modes us<strong>in</strong>g M = 16. Hence, the matrix eigenvalue problem (23)<br />

has size 512 × 512. We have solved it <strong>for</strong> a 32 × 32 grid <strong>in</strong> (x, y) chosen to sample 50 square mean free paths<br />

ls = σ −1<br />

s .<br />

Fig. 1 shows contour plots of the <strong>in</strong>tegrated backscattered flux<br />

0<br />

−1<br />

(46)<br />

2π0<br />

F (x, y) = Ψ(µ, ϕ, x, y)µdµdϕ (47)<br />

<strong>for</strong> various source directions (µ ′ , ϕ ′ ). The source is located 10 ls deep <strong>in</strong> the half space. The x and y axes are<br />

normalized with respect to ls.<br />

Because the source is not directed normal to the boundary surface <strong>in</strong> all cases, there is no axi-symmetry.<br />

Fig. 1 shows that the peaks of the <strong>in</strong>tegrated fluxes are skewed slightly towards the −ˆx direction. As µ → −1<br />

we observe that the peak of the flux becomes larger and narrower.<br />

In Fig. 2 we show the angular dependence of Ψ evaluated at the orig<strong>in</strong> (x, y) = (8 ls, 8 ls). We show it only<br />

<strong>for</strong> µ < 0 s<strong>in</strong>ce impos<strong>in</strong>g the boundary condition yields Ψ = 0 <strong>for</strong> µ > 0. Fig. 2 shows a non-trivial angular<br />

dependence. The diffusion approximation 1 is used often <strong>in</strong>stead of the transport equation to model light <strong>in</strong><br />

tissues. The fundamental assumption <strong>in</strong> this approximation is that the radiance has become nearly isotropic due<br />

to multiple scatter<strong>in</strong>g. S<strong>in</strong>ce Fig. 2 shows a significant angular dependence, it <strong>in</strong>dicates that this fundamental<br />

assumption is not valid.<br />

5.2. Asymptotic Result<br />

Suppose that the source is situated deep <strong>in</strong>side the half space. In that case the contribution from most of the<br />

plane wave modes <strong>in</strong> (42) and (45) to the po<strong>in</strong>t spread function is small. By neglect<strong>in</strong>g all plane wave modes<br />

except <strong>for</strong> the slowest decay<strong>in</strong>g one, we compute the asymptotic result of the the po<strong>in</strong>t spread function as z ′ → ∞.<br />

The Fourier trans<strong>for</strong>m of the po<strong>in</strong>t spread function def<strong>in</strong>ed as<br />

<br />

Ψ(ω, 0; κ) = Ψ(ω, ρ, 0)e −iκ·ρ dρ. (48)<br />

R 2<br />

Fourier trans<strong>for</strong>m<strong>in</strong>g (37) yields Ψ = G − Y . Because the eigenvalues λj are ordered by their real parts, the<br />

terms <strong>in</strong> a plane wave mode expansion <strong>in</strong>volv<strong>in</strong>g exp[−λ1(κ)z ′ ] <strong>in</strong> (42) and (45) are the slowest decay<strong>in</strong>g ones<br />

as z ′ → ∞. Neglect<strong>in</strong>g terms <strong>in</strong>volv<strong>in</strong>g faster decay<strong>in</strong>g modes yields the asymptotic result<br />

Ψ(ω, 0; κ) ∼ e −λ1(κ)z′<br />

V1(ω ′ ⎡<br />

; κ) ⎣V1(ω; κ) − <br />

⎤<br />

dj,1(κ)Vj(ω; κ) ⎦ , z ′ → ∞. (49)<br />

The coefficients dj,1 are determ<strong>in</strong>ed from (47) which do not depend on z ′ . The summation term <strong>in</strong> (49) shows<br />

that the po<strong>in</strong>t spread function still ma<strong>in</strong>ta<strong>in</strong>s a significant amount of directional <strong>in</strong><strong>for</strong>mation. It is because the<br />

radiance must satisfy the boundary condition (36). This directional variation is what is seen <strong>in</strong> Fig. 2.<br />

j


y<br />

y<br />

20<br />

10<br />

0<br />

−10<br />

−20<br />

20<br />

10<br />

0<br />

−10<br />

−20<br />

−20 −10 0 10 20<br />

x<br />

(a) source direction (µ ′ , ϕ ′ ) = (0.9894, π)<br />

−20 −10 0 10 20<br />

x<br />

(c) source direction (µ ′ , ϕ ′ ) = (−0.7554, π)<br />

−22<br />

−24<br />

−26<br />

−28<br />

−30<br />

−14<br />

−16<br />

−18<br />

−20<br />

−22<br />

−24<br />

−26<br />

y<br />

y<br />

20<br />

10<br />

0<br />

−10<br />

−20<br />

20<br />

10<br />

0<br />

−10<br />

−20<br />

−20 −10 0 10 20<br />

x<br />

(b) source direction (µ ′ , ϕ ′ ) = (0.7554, π)<br />

−20 −10 0 10 20<br />

x<br />

(d) source direction (µ ′ , ϕ ′ ) = (−0.9894, π)<br />

Figure 1. Integrated backscattered flux (47) <strong>for</strong> various source directions. The optical properties of the medium are given<br />

by σa/σs = 0.01 and g = 0.80. It is <strong>for</strong> a source at depth z ′ = 10 ls. The x and y axes are normalized by the scatter<strong>in</strong>g<br />

mean free path ls. The gray scale of the contours is given <strong>in</strong> dB.<br />

6. CONCLUSIONS<br />

We have discussed a method to compute the Green’s function <strong>for</strong> the radiative transport equation. It is given as<br />

an expansion <strong>in</strong> plane wave modes. These plane wave modes have been calculated numerically us<strong>in</strong>g the discrete<br />

ord<strong>in</strong>ate method.<br />

We have used the Green’s function <strong>for</strong> the radiative transport equation to compute the po<strong>in</strong>t spread function<br />

<strong>in</strong> a half space composed of a uni<strong>for</strong>m scatter<strong>in</strong>g and absorb<strong>in</strong>g medium. We have derived an asymptotic result<br />

<strong>for</strong> the po<strong>in</strong>t spread function when the source is deep <strong>in</strong>side the half space. It shows that the po<strong>in</strong>t spread function<br />

has a non-trivial direction dependence. Hence, the fundamental assumption <strong>in</strong> the diffusion approximation is<br />

not satisifed. There<strong>for</strong>e, the diffusion approximation should not be used to compute the po<strong>in</strong>t spread function.<br />

ACKNOWLEDGMENTS<br />

The work of A. D. Kim was partially supported by grants AFOSR: F49620-01-1-0465, ONR: N00014-02-1-0088<br />

−20<br />

−22<br />

−24<br />

−26<br />

−28<br />

−12<br />

−14<br />

−16<br />

−18<br />

−20<br />

−22<br />

−24<br />

−26


φ<br />

φ<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

−0.8 −0.6 −0.4 −0.2<br />

μ<br />

(a) source direction (µ ′ , ϕ ′ ) = (0.9894, π)<br />

−0.8 −0.6 −0.4 −0.2<br />

μ<br />

(c) source direction (µ ′ , ϕ ′ ) = (−0.7554, π)<br />

−26<br />

−27<br />

−28<br />

−29<br />

−30<br />

−31<br />

−32<br />

−23<br />

−24<br />

−25<br />

−26<br />

−27<br />

−28<br />

−29<br />

−30<br />

φ<br />

φ<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

−0.8 −0.6 −0.4 −0.2<br />

μ<br />

(b) source direction (µ ′ , ϕ ′ ) = (0.7554, π)<br />

−0.8 −0.6 −0.4 −0.2<br />

μ<br />

(d) source direction (µ ′ , ϕ ′ ) = (−0.9894, π)<br />

Figure 2. The angular dependence of the radiance at the orig<strong>in</strong>, Ψ(µ, ϕ, 8 ls, 8 ls) at various source angles. All other<br />

parameters are the same as <strong>in</strong> Fig. 1.<br />

and NSF: DMS-0071578.<br />

REFERENCES<br />

1. A. Ishimaru, Wave <strong>Propagation</strong> and Scatter<strong>in</strong>g <strong>in</strong> Random Media (New York: IEEE Press, 1996).<br />

2. K. M. Case and P. F. Zweifel, L<strong>in</strong>ear transport theory (Read<strong>in</strong>g, MA: Addison-Wesley, 1967).<br />

3. K. M. Case, “On boundary value problems of l<strong>in</strong>ear transport theory,” <strong>in</strong> Proc. Symp. <strong>in</strong> Applied Mathematics,<br />

Vol. 1, R. Bellman, G. Birkhoff and I. Abu-Shumays eds., American Mathematical Society 1969,<br />

pp. 17-36.<br />

4. A. D. Kim and J. B. Keller, “<strong>Light</strong> propagation <strong>in</strong> biological tissue,” J. Opt. Soc. Am. A 20 92-98 (2003).<br />

5. A. D. Kim, “<strong>Transport</strong> theory <strong>for</strong> light propagation <strong>in</strong> biological tissue,” to appear <strong>in</strong> J. Opt. Soc. Am. A.<br />

6. K. Atk<strong>in</strong>son, “Numerical <strong>in</strong>tegration on the sphere,” J. Austral. Math. Soc. Ser. B 23 332-347 (1982).<br />

−26<br />

−27<br />

−28<br />

−29<br />

−30<br />

−31<br />

−32<br />

−22<br />

−24<br />

−26<br />

−28<br />

−30

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!