+ All Categories
Home > Documents > Boundary conditions for light propagation in diffusive media with nonscattering regions

Boundary conditions for light propagation in diffusive media with nonscattering regions

Date post: 08-Oct-2016
Category:
Upload: hamid
View: 212 times
Download: 0 times
Share this document with a friend
11
Boundary conditions for light propagation in diffusive media with nonscattering regions Jorge Ripoll and Manuel Nieto-Vesperinas Instituto de Ciencia de Materiales de Madrid, Consejo Superior de Investigaciones Cientı ´ficas, Campus de Cantoblanco, 28049 Madrid, Spain Simon R. Arridge and Hamid Dehghani Department of Computer Science, University College London, Gower Street, London, WC1 E6BT, UK Received November 2, 1999; revised manuscript received May 8, 2000; accepted May 11, 2000 The diffusion approximation proves to be valid for light propagation in highly scattering media, but it breaks down in the presence of nonscattering regions. We present a compact expression of the boundary conditions for diffusive media with nonscattering regions, taking into account small-index mismatch. Results from an integral method based on the extinction theorem boundary condition are contrasted with both Monte Carlo and finite-element-method simulations, and a study of its limit of validity is presented. These procedures are il- lustrated by considering the case of the cerebro-spinal fluid in the brain, for which we demonstrate that for practical situations in light diffusion, these boundary conditions yield accurate results. © 2000 Optical Soci- ety of America [S0740-3232(00)00409-9] OCIS codes: 170.5270, 290.1990. 1. INTRODUCTION The study of light propagation through strongly scatter- ing media has received increasing attention during the past few years, partly as a result of to its applications to medical diagnosis. 1,2 In many practical situations it has been shown that visible or near-infrared light transport within turbid media such as human tissues 3 can be ad- equately modeled by the diffusion equation, on the basis of which several imaging methods have been analyzed 420 (see Ref. 20 for a review on the subject of optical tomog- raphy). Specifically, much research is motivated by the ability of optical radiation to diagnose brain tumors. However, a problem arises in studying complex systems such as the brain, in which not all regions diffuse light. These cases are those in which new boundary conditions that model the interaction of the diffuse light with non- diffusive media are needed. 2124 Many studies have been published concerning the way to correctly model the diffuse nondiffuse interface when the scattered light is detected from outside the diffusive medium. 2536 The diffusion approximation does not hold near nondiffuse interfaces, and to date there exists no closed correct expression for such boundary conditions. Even so, it has been shown that in a nonscattering region embedded in a diffuse medium (for instance, an air bubble in milk), the diffusion approximation, in combination with the radiative transfer equation (RTE) in a nonscat- tering medium, can still yield correct results 21,22 : That is, light propagation can still be accurately modeled by the diffusion approximation when light incident on the nonscattering region is already diffuse. An important in- stance of a biological medium with this type of region is the brain, which is mostly a scatterer except for those re- gions filled with cerebro-spinal fluid. The limit of valid- ity of such an approximation will depend both on the ratio between the scattering and the nonscattering volumes and on the refractive index. Therefore a closed expres- sion for the boundary conditions in such media is needed for correct modeling of the interaction at such interfaces. Following previous studies in which advances in this subject were made, 21,22 we present a compact expression of the boundary conditions for diffusive media with non- scattering regions. To obtain the expression for these boundary conditions, we make use of the RTE, 37 on which the diffusion approximation is based. To enhance the clarity and comprehensiveness of the steps followed, in Section 2 we concisely revise the main facts of the RTE, both in free space and in a diffusive medium. Then in Section 3 we derive the expression for the boundary con- ditions on the basis of the theory presented in Section 2. Further, we present the scattering integral equations, which include the interaction of the incident wave front of diffuse light with a nonscattering region by means of these boundary conditions. In Section 4 we present rig- orous numerical results for the corresponding system of integral equations, solved by means of the extinction theorem (ET) integral method, 3840 which, when applied both above and below the interface, yields the nonlocal boundary values for both the average intensity and its normal derivative. The ET method enables one to in- clude any type of mixed boundary conditions and to solve them without any approximation. We compare results derived from the ET with those obtained with the finite- element model 4143 (FEM), which takes into account the nonscattering medium by approximating the boundary conditions. Both methods are then contrasted with Monte Carlo (MC) simulations, and the advantages of us- ing one or another procedure are discussed accordingly. Ripoll et al. Vol. 17, No. 9 / September 2000 / J. Opt. Soc. Am. A 1671 0740-3232/2000/091671-11$15.00 © 2000 Optical Society of America
Transcript

Ripoll et al. Vol. 17, No. 9 /September 2000 /J. Opt. Soc. Am. A 1671

Boundary conditions for light propagation indiffusive media with nonscattering regions

Jorge Ripoll and Manuel Nieto-Vesperinas

Instituto de Ciencia de Materiales de Madrid, Consejo Superior de Investigaciones Cientıficas,Campus de Cantoblanco, 28049 Madrid, Spain

Simon R. Arridge and Hamid Dehghani

Department of Computer Science, University College London, Gower Street, London, WC1 E6BT, UK

Received November 2, 1999; revised manuscript received May 8, 2000; accepted May 11, 2000

The diffusion approximation proves to be valid for light propagation in highly scattering media, but it breaksdown in the presence of nonscattering regions. We present a compact expression of the boundary conditionsfor diffusive media with nonscattering regions, taking into account small-index mismatch. Results from anintegral method based on the extinction theorem boundary condition are contrasted with both Monte Carlo andfinite-element-method simulations, and a study of its limit of validity is presented. These procedures are il-lustrated by considering the case of the cerebro-spinal fluid in the brain, for which we demonstrate that forpractical situations in light diffusion, these boundary conditions yield accurate results. © 2000 Optical Soci-ety of America [S0740-3232(00)00409-9]

OCIS codes: 170.5270, 290.1990.

1. INTRODUCTIONThe study of light propagation through strongly scatter-ing media has received increasing attention during thepast few years, partly as a result of to its applications tomedical diagnosis.1,2 In many practical situations it hasbeen shown that visible or near-infrared light transportwithin turbid media such as human tissues3 can be ad-equately modeled by the diffusion equation, on the basisof which several imaging methods have been analyzed4–20

(see Ref. 20 for a review on the subject of optical tomog-raphy). Specifically, much research is motivated by theability of optical radiation to diagnose brain tumors.However, a problem arises in studying complex systemssuch as the brain, in which not all regions diffuse light.These cases are those in which new boundary conditionsthat model the interaction of the diffuse light with non-diffusive media are needed.21–24

Many studies have been published concerning the wayto correctly model the diffuse–nondiffuse interface whenthe scattered light is detected from outside the diffusivemedium.25–36 The diffusion approximation does not holdnear nondiffuse interfaces, and to date there exists noclosed correct expression for such boundary conditions.Even so, it has been shown that in a nonscattering regionembedded in a diffuse medium (for instance, an air bubblein milk), the diffusion approximation, in combinationwith the radiative transfer equation (RTE) in a nonscat-tering medium, can still yield correct results21,22: Thatis, light propagation can still be accurately modeled bythe diffusion approximation when light incident on thenonscattering region is already diffuse. An important in-stance of a biological medium with this type of region isthe brain, which is mostly a scatterer except for those re-gions filled with cerebro-spinal fluid. The limit of valid-

0740-3232/2000/091671-11$15.00 ©

ity of such an approximation will depend both on the ratiobetween the scattering and the nonscattering volumesand on the refractive index. Therefore a closed expres-sion for the boundary conditions in such media is neededfor correct modeling of the interaction at such interfaces.

Following previous studies in which advances in thissubject were made,21,22 we present a compact expressionof the boundary conditions for diffusive media with non-scattering regions. To obtain the expression for theseboundary conditions, we make use of the RTE,37 on whichthe diffusion approximation is based. To enhance theclarity and comprehensiveness of the steps followed, inSection 2 we concisely revise the main facts of the RTE,both in free space and in a diffusive medium. Then inSection 3 we derive the expression for the boundary con-ditions on the basis of the theory presented in Section 2.Further, we present the scattering integral equations,which include the interaction of the incident wave front ofdiffuse light with a nonscattering region by means ofthese boundary conditions. In Section 4 we present rig-orous numerical results for the corresponding system ofintegral equations, solved by means of the extinctiontheorem (ET) integral method,38–40 which, when appliedboth above and below the interface, yields the nonlocalboundary values for both the average intensity and itsnormal derivative. The ET method enables one to in-clude any type of mixed boundary conditions and to solvethem without any approximation. We compare resultsderived from the ET with those obtained with the finite-element model41–43 (FEM), which takes into account thenonscattering medium by approximating the boundaryconditions. Both methods are then contrasted withMonte Carlo (MC) simulations, and the advantages of us-ing one or another procedure are discussed accordingly.

2000 Optical Society of America

1672 J. Opt. Soc. Am. A/Vol. 17, No. 9 /September 2000 Ripoll et al.

Results are put forward for two distinct situations: thecase of a cylindrical nonscattering region embedded in adiffusive cylinder and the case of a cylindrical nonscatter-ing gap embedded in a diffusive cylinder. We discuss thelimit of validity of the boundary conditions established inthis work by considering the influence of the nonscatter-ing volume versus that of the diffusive volume. Finally,in Section 5 we summarize the main conclusions.

2. GENERAL EXPRESSIONS OF THERADIATIVE TRANSFER EQUATIONFor comprehensiveness and clarity in order to arrive atthe desired expression of the boundary conditions fordiffuse–nondiffuse interfaces, we summarize in this sec-tion the main facts of the theory of the RTE. We followthe notation used in Ref. 37, where a detailed account canbe found.

A way of characterizing the light flow of energy and itsinteraction with the medium is by means of the specificintensity I(r, s), which represents the average power fluxat point r that flows in the direction s and has units@W cm22 sr21#. In terms of the specific intensity, the av-erage intensity U and the total flux density J are definedas

U~r! 5 E4p

I~r, s!dV, (1)

J~r! 5 E4p

I~r, s!sdV, (2)

where both U and J have units of watts per square centi-meter and dV is the solid-angle element dV5 df sin u du. The total flux density that flows througha differential area dS 5 ndS is then

Jn~r! 5 J~r! • n 5 E4p

I~r, s!n • sdV, (3)

where n is the unit outward normal to the area elementdS. Therefore the amount of power dp emitted from dSin the direction of the unit vector s and flowing within asolid angle dV can be written as

d p~r, s! 5 I~r, s!n • sdSdV, (4)

is in units of watts. Then the total power emitted fromdS would be

dP~r! 5 E4p

5 I~r, s!n • sdSdV 5 Jn~r!dS, (5)

which is measured in watts. Let the area S separate twodifferent media with different refractive indices (see Fig.1): an upper medium with refractive index n0 and alower medium with refractive index n1 . Then we canwrite the total downward J2 and upward J1 fluxesthrough dS by means of Eq. (2) as

J1~r! 5 E~2p!1

I1→0t ~r, s!s • ndV, (6)

J2~r! 5 E~2p!2

I0→1t ~r, 2s!~2s • n!dV, (7)

where Ii→jt represents the specific intensity transmitted

on going from medium i into medium j. Their net fluxthrough dS is therefore Jn(r) 5 J1(r) 2 J2(r). Interms of conservation of power, we can then write the to-tal incident power at dS as a sum of the reflected andtransmitted powers; i.e.,

d pi~r, u i! 5 d pr~r, ur! 1 d pt~r, u t!, (8)

where

d pr~r, u t! 5 Rp~u i!d pi~r, u i!, (9)

d pt~r, u t! 5 @1 2 Rp~u i!#d pi~r, u i!. (10)

In Eqs. (9) and (10) Rp is the power reflectivity, whichin terms of the Fresnel reflection coefficients is given byRp 5 uRu2, where R actually denotes R i or R' dependingon whether the polarization is TM or TE. If the wave iscompletely unpolarized, then uRu2 should be equal to1/2(uR iu2 1 uR'u2). By means of the relationship shownin Eqs. (5), (8), and (10), we can rewrite Eqs. (6) and (7) as

J1~r! 5 E~2p!2

@1 2 uR1→0~u i!u2#I1~r, s!s • ndV i , (11)

J2~r! 5 E~2p!1

@1 2 uR0→1~u i!u2#I0~r, 2s!~2s • n!dV i ,

(12)

where I0 and I1 represent the specific intensities incidenton S from medium 0 and 1, respectively. It should be no-ticed that owing to the relationships between dpt and dpi

through the Fresnel coefficients, Eqs. (9) and (10), the in-tegration in Eqs. (11) and (12) is performed over the inci-dent angles, whereas in Eqs. (6) and (7) the integrationwas done over the transmitted angles. That is, in Eqs.(11) and (12), Ri→j(u i) represents the reflection coefficientcorresponding to the angle of incidence u i on going to me-dium j from medium i, and dV i 5 df i sin uidu i . Also,Eqs. (11) and (12) are general for the RTE, since the scat-

Fig. 1. Schematic view of the geometry with the upward (J1)and downward (J2) density flux at an interface.

Ripoll et al. Vol. 17, No. 9 /September 2000 /J. Opt. Soc. Am. A 1673

tering or absorbing specific properties of medium 0 and 1have not yet been introduced.

A. Nonscattering MediumThe expression for the RTE in a nonscattering mediumis37

n

c

]I~r, s!

]t1 s • ¹I~r, s! 1 maI~r, s! 5 0, (13)

where ma is the absorption coefficient of the nonscatteringmedium, is in units of inverse centimeters; n is the refrac-tive index; and c is the speed of light in vacuum. For acontinuous source located at r8, i.e., with no temporal de-pendence, the solution to Eq. (13) is

I~r, ur2r8! 5 I~r8, ur2r8!exp@2uaur 2 r8u#, (14)

where ur2r8 5 (r 2 r8)/ur 2 r8u. With reference to theconfiguration depicted in Fig. 1, consider the case inwhich the surface S separates the nonscattering innermedium (medium 0) with refractive index n0 and absorp-tion coefficient ma0 , from the outer medium (medium 1)with refractive index n1 , which can be either scatteringor nonscattering. The amount of power incident on theelementary area dS P S located at point r and emittedfrom the elementary area dS8 P S located at r8 (see Fig.2) will be

d pi~r, u! 5 I0~r8, ur2r8!exp@2ma0ur 2 r8u#cos udSdV,(15)

where cos u 5 n • ur82r and I0 represents the specific in-tensity inside the nonscattering volume. Note that weconsider S to be the union of all surfaces contributing tothe flux at r, which in general may have a complex topol-ogy; see Ref. 22 for further discussion of this topic. Thetotal power transmitted at r, taking all possible incidenceangles into consideration, is [see Eq. (10)]

dPt~r! 5 E~2p!1

~1 2 uR0→1~u!u2!I0~r8, ur2r8!

3 exp@2ma0ur 2 r8u#n • ur82rdSdV. (16)

By using the fact that dV 5 dS8 • ur2r8 /ur 2 r8u2 (seeFig. 2), and J2(r) 5 dPt(r)/dS [see Eq. (5)], we write thetotal flux density transmitted through every point r of Sin terms of the specific intensity radiated by the completesurface as

J2~r! 5 ESI0~r8, ur2r8!G~r 2 r8!dS8, r P S, (17)

where we have defined

G~r 2 r8! 5 @1 2 uR0→1~u!u2#exp@2ma0ur 2 r8u#

ur 2 r8u2

3 V ~r 2 r8!cos u8 cos u,

cos u 5 n •

~r8 2 r!

ur8 2 ru, cos u8 5 n8 •

~r 2 r8!

ur8 2 ru.

(18)In Eq. (18), V (r 2 r8) is a visibility factor, which is ei-

ther unity if both points r and r8 can be joined by astraight line without intersecting an interface (i.e., when

they are visible to each other) or zero when such astraight line does not exist. From now on, and for thesake of simplicity, this factor will be implicitly assumedand therefore omitted.

An expression similar to that of Eq. (17) for the inwarddensity flux can be found by means of Eq. (1) for the av-erage intensity U0 inside the nonscattering volume:

U0~r! 5 ESI0~r, ur82r!

n8 • ur2r8

ur 2 r8u2 dS8, r P V, (19)

which, with Eq. (14), can be rewritten as

U0~r! 5 ESI0~r8, ur82r!G~r 2 r8!dS8, r P V. (20)

In Eq. (20), G(r 2 r8) is

G~r 2 r8! 5exp@2ma0ur 2 r8u#

ur 2 r8u2 cos u8. (21)

It must be stated that Eqs. (17) and (20) are general forthe RTE, since no approximation has yet been imposed onI0 .

B. Diffusion ApproximationAddressing once again the configuration depicted in Fig.1, consider the case in which medium 1 is a strong scat-terer with n1 as index of refraction and ma1 and ms1 asabsorption and scattering coefficients, respectively.These coefficients are such that ms1 @ ma1 so that the dif-fusion approximation to the RTE is now valid. In thiscase the main assumption is a first-order angular depen-dence of the specific intensity I1(r, s),37

I1~r, s! 5 aU1~r! 1 bJ1~r!s, r P V, (22)

where I1 is the specific intensity, U1 stands for the aver-age intensity [see Eq. (1)], and J1 denotes the total den-sity flux [see Eq. (2)] inside the diffusive medium. Wefind the values of a and b by applying Eqs. (1) and (2),thus obtaining a 5 1/4p and b 5 3/4p. The angular de-pendence of I1 can be seen schematically in Fig. 3. Ex-pression (22) also implies that at any point inside the dif-fusive medium, the total density flux is governed by Fick’slaw (see Ref. 37 for a detailed derivation),

J1~r! 5 21

3~ms18 1 ma1!@¹U1~r! 2 Q1~r!#, r P V,

(23)

Fig. 2. Scheme for the solid-angle relationship between dS5 ndS and dS8 5 n8dS8.

1674 J. Opt. Soc. Am. A/Vol. 17, No. 9 /September 2000 Ripoll et al.

where D1 5 @3(ms18 1 ma1)#21 is the diffusion coefficient(measured in centimeters), ms18 5 (1 2 g)ms1 is the re-duced scattering coefficient, and g is the average cosine ofthe scattering angle.37 In Eq. (23) the term Q1 takes careof the anisotropy and is defined in terms of the angular-dependent photon source e (see Sec. 9-2 of Ref. 37) as

Q1~r! 5 E4p

e~r, s!sdV, (24)

where e is the power radiated by the source per unit vol-ume per unit solid angle in the direction s. However, aswill be shown, the derivation of the boundary conditionsis in terms of Jn1 5 J1 • n and therefore is independentof the expression used for Jn1 .

To find an expression similar to Eq. (17) for the non-scattering medium, we must make use of the value of I1at the interface. To this end, Eq. (23) can be rewritten as

J1 5 Jn1n 1 Jt1t, (25)

where n and t are the surface normal and the tangentialunit vector, respectively. Introducing Eq. (22) into Eq.(11), making use of Eq. (25), we obtain

J1~r! 5 E~2p!2

U1~r!

4p@1 2 uR1→0~u!u2#cos udV

1 E~2p!2

3Jn1~r!

4p@1 2 uR1→0~u!u2#cos2 udV,

(26)

where the contribution from Jt1 is zero, owing to the s• n factor in Eq. (11). Since the total flux density [seeEq. (3)] through the surface must be constant, we also ob-tain the relationship

Jn~r! 5 Jn1~r! 5 Jn0~r!, (27)

where Jn1 inside the diffusive medium is [see Eq. (23)]

Fig. 3. Scheme for the specific intensities inside the diffusivemedium and emerging into the nonscattering medium. Note theangular dependence to first order of I1 in the diffusive medium,whereas in the nonscattering medium I0 is angle independent.The light distribution radiated into the nonscattering medium,i.e., I0(r), gives rise to Lambert’s power law: d p(r)5 I0(r)cos u dS.

Jn1~r! 5 2D1@n • ¹U1~r! 2 n • Q1~r!#

5 2D1

]U1~r!

]n1 D1n • Q1~r!. (28)

Since Jn1 is defined as an integral over all possible angles[see Eq. (3)], neither U1 nor Jn1 has angular dependence.Therefore, with dV 5 2pd(cos u), Eq. (26) can be rewrit-ten as

J1~r! 5 RU

U1~r!

21 RJ

Jn~r!

2, r P S, (29)

where

RU 5 E0

1

@1 2 uR0→1~u!u2#cos ud~cos u!,

RJ 5 3E0

1

@1 2 uR0→1~u!u2#cos2 ud~cos u!. (30)

Relation (27) has been employed in Eq. (29) and will beused from now on. A detailed derivation of these equa-tions can be found in Refs. 33 and 44.

3. BOUNDARY CONDITIONSEquation (17) gives us the expression for the inward fluxin terms of the specific intensity inside the nonscatteringmedium, whereas Eq. (29) gives us the expression for theoutward flux in terms of the average intensity and den-sity flux inside the diffusive medium. What is needed atthis point is a boundary condition for the light wave thatmatches both media. The simplest approximation, whichwill be used throughout this paper, is to consider the lightthat is isotropically radiated from the diffusive medium(see Fig. 3). This involves the assumption that the spe-cific intensity, radiated from the surface S into the non-scattering medium, will not have an angular dependence.Then, since the total flux that emerges from a point in thesurface r is J1, the specific intensity inside the nonscat-tering medium I0 must be expressed as

I0~r, s! 5 I0~r! 5 aJ1~r!, (31)

where a is a constant. Since the total flux radiated intothe nonscattering medium must be J1, by means of Eq.(6) we obtain a 5 1/p. Therefore, the power radiatedinto the nonscattering medium from a certain point r ofthe surface element dS will be [see Eq. (4)]

d p~r, u! 5 I0~r!cos udSdV 5J1~r!

pcos udSdV, (32)

and the total power emitted from dS into the nonscatter-ing medium would be

dP~r! 51

pE

~2p!1J1~r!cos udSdV 5 J1~r!dS. (33)

Equation (32) is Lambert’s cosine law45 and is fre-quently used for describing light emerging from astrongly scattering medium, the surface boundary S act-ing as a secondary source (see Fig. 3). Substituting Eq.(31) into Eq. (17), we obtain the following expression forthe total downward flux:

Ripoll et al. Vol. 17, No. 9 /September 2000 /J. Opt. Soc. Am. A 1675

J2~r! 51

pE

SJ1~r8!G~r 2 r8!dS8, r P S. (34)

The total flux going into the diffusive medium is there-fore a superposition of all outwardgoing fluxes at the sur-face S that are transmitted at point r (see Fig. 2).

Therefore the total normal flux, Jn 5 J1 2 J2, is

Jn~r! 5 RU

U1~r!

21 RJ

Jn~r!

2

21

pE

SFRU

U1~r8!

21 RJ

Jn~r8!

2 G3 G~r 2 r8!dS8, r P S. (35)

Grouping terms, we then obtain that

U1~r! 5 CnJn~r! 11

pE

SFU1~r8! 1

RJ

RUJn~r8!G

3 G~r 2 r8!dS8, r P S, (36)

where Cn 5 (2 2 RJ)/RU , as shown in Refs. 33 and 44.Equation (36) is the nonlocal boundary condition that wewere seeking for light propagation in diffusive media withnonscattering regions. It must be stated that this bound-ary condition cannot take into account multiple reflec-tions inside the nonscattering region when dealing withindex-mismatched media. In the case of a plane inter-face (i.e., an open surface) there is only outgoing flux, andwe recover the expression for the boundary condition inthe diffusion approximation33,37:

U1~r! 5 CnJn~r!, r P S. (37)

We wish to emphasize that more accurate values of Cnbeyond the diffusion approximation can be found in theliterature26,33 for some practical situations in which thesolution reached by using Eq. (37) is not accurate enough.Even so, this is still an active field of research, and newmethods to overcome this problem are needed.

The average intensity inside the nonscattering me-dium, Eq. (20), in this diffusion approximation case can berewritten with the aid of Eq. (31) as

U0~r! 51

pE

SJ1~r8!G~r 2 r8!dS8, r P V. (38)

At this point it must be underlined that Eqs. (36) and(38) have been derived for a three-dimensional configura-tion. The expressions for these equations in two dimen-sions can be found in Appendix A and are those that willbe employed in our numerical simulations. Wheneverdealing with non three-dimensional configurations, greatcare must be taken with the RTE, since conversion fromthree to two dimensions is nontrivial. A first-order ap-proximation to Eq. (36) was used in Refs. 21 and 22 todeal with nonscattering regions in diffusive media for theindex-matched case. This approximation consisted of ap-plying the boundary condition Eq. (37) and including asecondary source at the interface given by

U ~sec!~r! 51

pE

SJn

~0 !~r8!G~r 2 r8!dS8, r P S, (39)

where Jn(0) denotes the zeroth-order approximation to the

total flux, i.e., Jn(0) 5 U/Cn . The main assumption that

this approximation makes is that the total normal flux isequal to the total outward flux, i.e.,

Jn~0 !~r! . J1~r! 5 RU

U1~r!

21 RJ

Jn~0 !~r!

2, r P S,

(40)which gives the boundary condition CnJn

(0) 5 U1 . Thisapproximation is expected to break down when the totalinward flux is not small, but as shown in Refs. 21 and 22,it yields accurate results with biological parameters andwill be employed in Section 4 with the FEM model. Also,whenever dealing with great differences in the refractiveindices, both Eqs. (36) and (40) are expected to breakdown, since they cannot account for multiple reflectionsinside the nonscattering medium. That is, both Eqs. (36)and (40) deal with transmitted light and do not considerthe reflected part, which would give rise to nonlinearequations.

A. Scattering Integral EquationsLet us address the scattering configuration depicted inFig. 4. An infinite homogeneous diffusive medium of vol-ume V with diffusion coefficient D1 , absorption coefficientma1 , and refractive index n1 , contains a nonscattering ob-ject of volume V, with absorption coefficient ma0 and in-dex of refraction n0 , delimited by the surface S. Assum-ing the photon source located at rsource inside V andmodulated by a frequency v, the average intensityU1(r, t) 5 U1(r)exp(2ivt) is governed by the diffusionequation in the frequency domain, which can be ex-pressed in V as2,38

¹2U1~r! 1 k12U1~r! 5 2

S0~r!

D12 ¹ • Q1~r!, r P V.

(41)

Equation (41) is derived from current conservation.37

S0 is the source strength defined as [see Eq. (24)]

S0~r! 5 E4p

e~r, s!dV (42)

and represents the power generated per unit volume(watts per inverse centimeters cubed). D1 is the diffu-sion coefficient previously defined. It must be stated thatwhen we are dealing with an isotropic source, the termQ1 5 0. The complex diffusion wave number k1 is

k12 5 2

ma1

D11 i

vn1

cD1, (43)

c being the speed of light in vacuum. In Eq. (41) we haveneglected the term ¹D1(r) • ¹U1(r)/D1(r) (see Ref. 38),since we consider the case in which both D1 and ma1 areconstant throughout the diffusive medium (i.e., excludingthe nonscattering regions). The Green’s functionG(k1ur 2 r8u) corresponding to the diffusive medium sat-isfies

1676 J. Opt. Soc. Am. A/Vol. 17, No. 9 /September 2000 Ripoll et al.

¹2G~k1ur 2 r8u! 1 k12G~k1ur 2 r8u! 5 24pd ~r 2 r8!,

r, r8 P V, (44)

and its expression is40,46,47

G~k1ur 2 r8u! 5 exp@ik1ur 2 r8u#/ur 2 r8u].

Inside the nonscattering volume V, the solution of Eq.(14) in the case of a modulated source with frequency v is

I0~r, ur2r8! 5 I0~r8, ur2r8!

3 expF S 2ma0 1 ivn0

c DUr 2 r8UG ,

r, r8 P V, (45)

and the average intensity inside V is [see Eq. (38)]

U0~r! 51

pE

SFRU

2U1~r8! 1

RJ

2Jn~r8!GGv~r 2 r8!dS8,

r P V,

Gv~r 2 r8! 5exp$@2ma0 1 i~vn0 /c !#ur 2 r8u%

ur 2 r8u2 cos u8.

(46)

By means of Eq. (46) we are able to find the average in-tensity at any point inside the nonscattering volume V.

Since the total density flux is governed by Fick’s law[Eq. (23)] at the surface S, then, according to Eqs. (23),(27), and (28), Jn(r) 5 2D1]U1(r)/]n 1 D1n • Q1(r).In this case Green’s theorem,46

Ev~U1¹2G 2 G¹2U1!d3r 5 E

s~U1¹G 2 G¹U1! • ds,

for the diffusive medium can be rewritten as

EV

~U1¹2G 2 G¹2U1!d3r

5 ESFU1

]G

]n1 GS Jn

D11 n • Q1D GdS. (47)

Multiplying Eq. (41) by G, Eq. (44) by U, subtractingboth, performing an integral over the volume V, and ap-

Fig. 4. Scattering geometry of a nonscattering region embeddedin a diffusive medium.

plying Green’s theorem [Eq. (47)], we obtain the followingcoupled integral equations for the diffusivemedium40,46,47:

r, r8 P V:

U1~r! 5 U ~ inc!~r! 1 SQ~r! 11

4p

3 ESFU1~r8!

]G~k1ur 2 r8u!

]n8

1 G~k1ur 2 r8u!Jn~r8!

D1GdS8, (48)

r P V, r8 P V:

0 5 U ~ inc!~r! 1 SQ~r!

11

4pE

SFU1~r8!

]G~k1ur 2 r8u!

]n8

1 G~k1ur 2 r8u!Jn~r8!

D1GdS8, (49)

where we have denoted the incident field intensity U (inc)

as

U ~ inc!~r! 5 21

4pE

VFS0~r8!

D11 ¹ • Q1~r8!G

3 G~k1ur 2 r8u!dr8 (50)

and the contribution of the source anisotropy at theboundary, SQ as

SQ~r! 51

4pE

SF n8 • Q1~r8!

D1GG~k1ur 2 r8u!dS8. (51)

The nonlocal boundary condition between the two media,Eq. (36), is

U1~r! 5 CnJn~r! 11

pE

SFU1~r8! 1

RJ

RUJn~r8!G

3 Gv~r 2 r8!dS8, r P S,

Gv~r 2 r8! 5 @1 2 uR0→1~u!u2#Gv~r 2 r8!cos u. (52)

Equations (48)–(52) constitute a closed set of integralequations that can be solved numerically for both U1 andJn . An exact procedure to solve this type of equation inarbitrary diffuse–diffuse interfaces is the extinction theo-rem (ET) method, which can be found in Ref. 38. Its usefor diffusive media with index mismatch is reported inRef. 44. The method has also been used extensively inelectromagnetic scattering, and further references can befound in Refs. 39 and 40. Its use in other areas of re-search can be found in Ref. 48, where it is referred to asthe boundary-element method.

Once the values of U1 and Jn are found by solving Eqs.(48) and (52), U1 and U0 can be univocally determined atany point inside both the diffusive and the nonscatteringmedium by means of Eqs. (48) and (46), respectively. Itshould also be remarked that by similar arguments (seeRef. 38 and references therein), the above equations can

Ripoll et al. Vol. 17, No. 9 /September 2000 /J. Opt. Soc. Am. A 1677

be straightforwardly generalized to cases in which thereare several nonscattering regions embedded in the diffu-sive medium, no matter whether these regions are roughor of any other arbitrary shape. Also, as in the casesolved numerically in Section 4, the volume V can bebounded from the outside by a nonscattering medium:Actually, this would represent the more realistic case of adiffusive object surrounded by air with a nonscatteringregion embedded within (see Fig. 5).

4. NUMERICAL RESULTSTo assess the validity of the boundary condition [Eq. (36)],we study the two-dimensional configurations depicted inFig. 5, where an outer cylinder of fixed radius Rout5 2.5 cm, filled with a diffusive medium of parametersms18 5 20 cm21 and ma1 5 0.1 cm21, is surrounded by air.We shall consider two cases: one in which a nonscatter-ing cylinder of radius R in is included [Fig. 5(a)] and an-other in which there is a nonscattering gap of outer ra-dius Rgap and inner radius R in between two diffusiveregions [Fig. 5(b)]. In both cases the absorption coeffi-cient for the nonscattering region is ma0 5 0.05 cm21.The speed of light is constant throughout the media, withrefractive index n0 5 n1 5 1.4. A continuous isotropicpoint source (v 5 0) is placed at one transport mean freepath ltr 5 3D1 . 0.05 cm from the outer surface withinthe diffusive medium (i.e., rsource 5 (Rout 2 ltr). 2.45 cm). To quantify the limits of validity of theboundary condition, we shall consider several object sizes.For small objects, we must always take into considerationthat for the diffusion approximation to remain valid, lightmust travel at least a few mean free paths. As shown inRef. 13, on comparison with a Monte Carlo result, sizessmaller than or equal to the mean free path start causingdeviations.

To permit reaching numerical convergence with the ETmethod, all surfaces have a discretization dS5 0.025 cm. The numerical calculations took 3 min forFig. 5(a) and 9 min for Fig. 5(b) on a 200-MHz personalcomputer with a 128-Mb RAM. When the FEM wasused, the mesh had between 949 nodes, 1614 elements(R in 5 2 cm) and 2323 nodes, 4456 elements (R in5 0.25 cm) for the case in Fig. 5(a) and .2000 nodes and.4000 linear elements for the case in Fig. 5(b) (exact

Fig. 5. Cases considered: (a) nonscattering cylinder of radiusR in embedded in a diffusive cylinder of radius Rout , (b) nonscat-tering gap of inner radius R in and outer radius Rgap embedded ina diffusive cylinder of radius Rout .

numbers were different for each position of the gap).Computation times were 5–20 s on a 450-MHz PentiumII. The FEM has been reported extensively for manycases,41–43 and with this particular configuration in Ref.22. Therefore we shall not explain it here. The FEMmethod with the approximate boundary conditions has al-ready been contrasted in Ref. 22 with the MC method andwith a discrete ordinate transport code, showing verygood agreement with both. To assess the accuracy of thecalculations performed with the ET and the FEM, wecompared them with results from MC simulations. Thedescription of the MC method for photon diffusion is wellknown (see, for example, Refs. 49 and 50) and will not bedescribed here. The program is the same as that used inRef. 22. 107 photons were launched during performanceof the MC simulations. In all three methods the detectorscanning was performed along the surface of the outercylinder, i.e., at rdetec 5 2.5 cm, with variation of the an-gular distance from the source. Once again, we note thatwhen applying the ET model, we will use the completeboundary condition, Eq. (36), whereas when using theFEM method, we will use the approximate boundary con-dition, Eqs. (39) and (40). Also, whenever we were deal-ing with completely diffusive volumes, i.e., in the absenceof the nonscattering region, the results from ET and FEMagreed to a high degree, thus implying that any differ-ences the two methods may present with dealing withnonscattering volumes are due solely to the difference inthe boundary conditions applied.

Figure 6 shows the results for the MC, FEM, and ET

Fig. 6. Average intensity measured on the outer surface Rout5 2.5 cm versus detector angle separation, for the configurationdepicted in Fig. 5(a). Values of R in 5 0.25, 0.5, 0.75, 1.0, 1.25,1.5, 1.75, and 2.0 cm. Results are presented for simulations per-formed with MC (solid curve), FEM (s), and ET (d). In all casesa dc source (v 5 0) was located at rsource 5 2.45 cm at u 5 0.ma0 5 0.05 cm21, ma1 5 0.1 cm21, ms1 5 20 cm21, g 5 0.

1678 J. Opt. Soc. Am. A/Vol. 17, No. 9 /September 2000 Ripoll et al.

calculations in the case represented in Fig. 5(a). R in hasbeen varied from R in 5 0.25 cm to R in 5 2.0 cm in0.25-cm steps. This implies that the volume ratio of dif-fusive medium to total volume varies from Vr 5 36% upto Vr 5 100%. As shown in Fig. 6, both FEM and the ETmodel account accurately for the photon transport in allcases, with ET being slightly more accurate. The reasonfor this difference is twofold: First, ET is an exactmethod, whereas FEM assumes an approximation on]U/]r. Second, the FEM uses the approximated bound-ary condition, Eqs. (39) and (40), whereas the ET is ableto include the complete boundary condition, Eq. (36). Asexpected, results have a greater deviation from MC forgreater values of R in , i.e., for lower values of Vr . As ageneral rule, we may say that for values of Vr . 75% weobtain quite accurate results with both the FEM and theET. Even so, results obtained for the case with R in5 2.0 cm, i.e., Vr 5 36%, are quite impressive if we takeinto account that then 64% of the total volume is nonscat-tering. Also, if we look at the case with R in 5 0.25 cm,we see that there is a greater deviation than expected.The main reason for this is that such a small nonscatter-ing volume stands in the limit of length scales withinwhich the diffusion approximation works. That is,within the diffusion approximation context, in the samemanner as diffusive regions with sizes of the order of ltrcan barely be considered actually diffusive (see Ref. 13),nonscattering objects with sizes of the order of ltr embed-ded in diffusive media, can barely be considered to be ac-tually nonscattering. Statistically this would mean thatin a particular region of the diffusive object we have alower concentration of scatterers, which, since the diffu-sion approximation always deals with averages, has avery low contribution. Thus, following these consider-ations, the results put forward by the FEM and ET meth-ods for the case R in 5 0.25 cm show lower intensities (i.e.,the nonscattering object is less visible) than those for theMC method.

In Fig. 7 we show the results for the MC, FEM, and ETcalculations in the case represented in Fig. 5(b). Now wehave considered two values of R in , namely, R in 5 1.0 cm[Fig. 7(a)] and R in 5 1.5 cm [Fig. 7(b)]. For each value ofR in we computed five values of Rgap , which conferred thegap widths W 5 Rgap 2 R in 5 0.1, 0.3, 0.5, 0.7, and 0.8cm. In all cases we were always above a volume ratioVr . 75%, namely, between Vr 5 95% (W 5 0.8 cm) andVr 5 99.84% (W 5 0.1 cm), so that a deviation due to lowvalues of the volume ratio is not expected. As seen inFig. 7, both the FEM and the ET show excellent agree-ment with MC. In this configuration, we once againreach more accurate results with the ET for high values ofW, namely, W 5 3, 5, 7, and 8 mm, for the reason statedabove. Also, a greater deviation of the ET occurs forsmaller values of the gap width W, namely, W 5 0.1 cm,in which cases the FEM yields better results. The reasonfor this is that this width is of the order of ltr , and there-fore the arguments that were applied to the case R in5 0.25 cm of Fig. 6 can now be applied here. That is,when dealing with widths on the order of ltr , the diffusionapproximation fails and therefore the boundary condi-tions do not yield correct results. Also, since the FEMuses the approximate boundary condition, it appears that

in cases of low values of W < 3 mm, the approximateboundary condition works better than the completeboundary condition. Also, on comparing the cases W5 0.5, 0.7, and 0.8 cm from Figs. 7(a) and 7(b), we findthat the depth of the gap buried in the diffusive medium,i.e., the value of R in , does not change the accuracy of theresults. We draw the attention to the curve correspond-ing to W 5 0.8 cm in Fig. 7(a), where we find a sharp in-tensity decrease that is also predicted by the ET and theFEM (this effect was also found in Ref. 21, where it is re-ferred to as a ‘‘kink’’). This decrease in intensity appearswhen the inner diffusive volume, r < R in , is in the line ofsight between source and detector, and its magnitude isgreater the closer the outer radius Rout is from the source.

It is important to remark that what has been presentedhere corresponds to an isotropic point source. Wheneverdealing with more realistic sources, such as a laser sourceimpinging the diffusive object from outside, the light en-tering the diffusive region is nonisotropic at least within afew mean free paths, and the correct expression for Q1must be included in Eq. (51). In any case, if the first non-diffuse interface, i.e., Rgap , is of the order of a few meanfree paths ltr near the source (typically ;2ltr), the bound-ary condition Eq. (52) is expected to break down, sincenow light incident at the interface will no longer be welldescribed by the diffusion approximation, Eq. (22) (see

Fig. 7. Average intensity measured on the outer surface Rout5 2.5 cm versus detector angle separation, for the configurationdepicted in Fig. 5(b). Values of (a) R in 5 1.5 cm, Rgap 5 1.6 cm(W 5 0.1 cm), 1.8 cm (W 5 0.3 cm), and 2.0 cm (W 5 0.5 cm).(b) R in 5 1.0 cm, Rgap 5 1.1 cm (W 5 0.1 cm), 1.3 cm (W5 0.3 cm), and 1.5 cm (W 5 0.5 cm). Results are presented forsimulations performed with MC (solid curve), FEM (s), and ET(d). In all cases a dc source (v 5 0) was located at rsource5 2.45 cm at u 5 0. ma0 5 0.05 cm21, ma1 5 0.1 cm21, ms15 20 cm21, g 5 0.

Ripoll et al. Vol. 17, No. 9 /September 2000 /J. Opt. Soc. Am. A 1679

Fig. 3). However, in the cases presented here the valuesof Rgap are such that they are located at least 4ltr from thesource, and as shown in Fig. 7(a) the diffusion approxima-tion still accurately describes light transport.

5. CONCLUSIONSIn this paper we have studied the behavior of light trans-port when it encounters nonscattering regions in a diffu-sive medium. We have presented an expression for theaverage intensity inside both diffusive and nonscatteringmedia and the matching across the interface separatingboth media. This expression gives rise to complete equa-tions for the boundary conditions that model diffuse lightin the presence of nonscattering regions. These expres-sions are rigorous within the diffusion approximation.We have presented an approximation to these boundaryconditions that has already been employed in previousstudies.21,22 To assess the validity of the boundary con-ditions, we have contrasted FEM and ET calculationswith MC simulations. Due to the inherent structure ofthe FEM, the inclusion of the complete boundary condi-tion is quite cumbersome, and hence the approximatedexpression has been used. On the other hand, ET calcu-lations were possible with the complete boundary condi-tion, and the corresponding integral equations were rig-orously solved without any approximation. We haveshown that both the ET with the complete boundary con-dition and the FEM with the approximate expressionyield very good results when compared with MC, al-though slightly better agreement is reached with the ET.These calculations have been carried out for two differenttypes of configurations: a nonscattering cylinder embed-ded in a diffusive cylinder and a nonscattering gap in adiffusive cylinder. We have demonstrated that both situ-ations can be accurately modeled, and the limits of valid-ity of the FEM and the ET have been established.

We stress, nevertheless, that the objective of this workis not to compare the FEM with the ET but rather to as-sess the validity of the boundary conditions put forward.Discussions of the relative advantages and disadvantagesof the two methods can be found elsewhere,38,48 and havenot been included here. Even so, we wish to state thatthe FEM requires less computational time than the ET,and therefore the ET is so far not useful with time-dependent equations. However the ET yields more accu-rate results, and complex boundary conditions can be in-cluded. On the other hand, the FEM, although not soaccurate, is faster, and time dependence can be includedwithout the need for a large amount of computing time.Therefore we suggest the use of ET to assess the accuracyof other faster methods, as has been done in this paper.

The modeling of nonscattering regions embeddedwithin diffusive media is a subject of major interest, sincea great deal of realistic media, such as the brain, includesthem. We have presented here the boundary conditionsthat model such media, no matter the complexity of theirsurface geometry, and have studied their limits of valid-ity. A subject that should be further addressed is the ef-fect of index mismatch between the nonscattering regionsand the surrounding diffusive media. As mentioned inSection 3, the limits of validity of the boundary conditions

presented in this paper are not clear in the case of index-mismatched media. This is an issue still under investi-gation.

APPENDIX A: BOUNDARY CONDITIONS INTWO DIMENSIONSThe expressions for the boundary conditions in two di-mensions are quite useful since the great majority of nu-merical simulations are performed in two dimensions be-cause of their lower computational cost. Let us assumethat in Fig. 2 the configuration is constant in the z direc-tion. We write r 5 (R, z), the unit normal n(r)5 N(R), and the unit area as dS 5 dAdz. Since in thiscase U and J are z independent, we can rewrite Eq. (46)as

U0~R! 51

pE

AJ1~R8!Gv

2D~R 2 R8!dA8, (A1)

Gv2D~R 2 R8! 5 E

2`

1`

Gv~r 2 r8!dz8. (A2)

In Eq. (47), cos u8 can be rewritten as cos u85 cos U8 cos b, where cos U8 5 uR2R8 • N8. With perfor-mance of the change of variable, cos b 5 R/AR2 1 z2, Eq.(A2) can be expressed as

Gv2D~R 2 R8! 5

uR82R • N

uR 2 R8u

3 E2p/2

p/2

expF S 2ma0 1 ivn0

c DR 2 R8

cos bG

3 cos bdb. (A3)

Proceeding in a similar way, we obtain the following ex-pression for the boundary condition Eq. (52):

U1~R! 5 CnJn~R! 11

pE

AFU1~R8! 1

RJ

RUJn~R8!G

3 Gv2D~R 2 R8!dA8, R P A, (A4)

Gv2D~R 2 R8! 5 E

2`

1`

Gv~r 2 r8!dz8. (A5)

Performing the same substitutions as in Eq. (A3), weobtain

Gv2D~R 2 R8! 5

~ uR82R • N!~uR2R8 • N8!

uR 2 R8u

• E2p/2

p/2

@1 2 uR0→1~u~i !!u2#

3 expF S 2ma0 1 ivn0

c D uR 2 R8u

cos bG

3 cos2 bdb, (A6)

where cos u (i) 5 cos b cos U. In the case in which there isno index mismatch, expression (A6) for Gv

2D can be ap-proximated to

1680 J. Opt. Soc. Am. A/Vol. 17, No. 9 /September 2000 Ripoll et al.

Gv2D~R 2 R8! .

p

2

cos U cos U8

uR 2 R8u

3 expF S 2ma0 1 ivn0

c D uR 2 R8uG .

(A7)

Expression (A7) is exact when v 5 0, such as in thecases addressed in Refs. 21 and 22 and in the instancespresented in this paper.

ACKNOWLEDGMENTSWe thank E. Okada and his group for help with the MonteCarlo code. This research has been supported by Comi-sion Interministerial de Ciencia y Tecnologıa of Spain un-der grant PB98-0464 and by the Fundacion RamonAreces. J. Ripoll acknowledges a scholarship from Minis-terio de Educacion y Cultura of Spain. S. R. Arridge andH. Dehghani thank the Welcome Trust and the Engineer-ing and Physical Sciences Research Council (UK).

The authors can be reached at the addresses on thetitle page or by e-mail: [email protected], [email protected], S. [email protected], H. [email protected].

REFERENCES1. E. B. de Haller, ‘‘Time-resolved transillumination and opti-

cal tomography,’’ J. Biomed. Opt. 1, 7–17 (1996).2. A. Yodh and B. Chance, ‘‘Spectroscopy and imaging with

diffusing light,’’ Phys. Today, March 1995, pp. 38–40.3. See related studies in Advances in Optical Imaging and

Photon Migration, J. Fujimoto and M. S. Patterson, eds.,Vol. 21 of Trends in Optics and Photonic Series (Optical So-ciety of America, Washington, D.C., 1998).

4. S. R. Arridge, P. van der Zee, M. Cope, and D. T. Delpy, ‘‘Re-construction methods for near infra-red absorption imag-ing,’’ in Time-Resolved Spectroscopy and Imaging of Tis-sues, B. Chance and A. Katzir, eds., Proc. SPIE 1431, 204–215 (1991).

5. M. A. O’Leary, D. A. Boas, B. Chance, and A. G. Yodh, ‘‘Ex-perimental images of heterogeneous turbid media byfrequency-domain diffusing-photon tomography,’’ Opt. Lett.20, 426–428 (1995).

6. C. P. Gonatas, M. Ishii, J. S. Leigh, and J. C. Schotland,‘‘Optical diffusion imaging using a direct inversionmethod,’’ Phys. Rev. E 52, 4361–4365 (1995).

7. Ch. L. Matson, N. Clark, L. McMackin, and J. S. Fender,‘‘Three-dimensional tumor localization in thick tissue withthe use of diffuse photon-density waves,’’ Appl. Opt. 36,214–220 (1997).

8. X. D. Li, T. Durduran, A. G. Yodh, B. Chance, and D. N.Pattanayak, ‘‘Diffraction tomography for biochemical imag-ing with diffuse-photon density waves,’’ Opt. Lett. 22, 573–575 (1997).

9. S. A. Walker, S. Fantini, and E. Gratton, ‘‘Image recon-struction by backprojection from frequency-domain opticalmeasurements in highly scattering media,’’ Appl. Opt. 36,170–179 (1997).

10. H. Jiang, K. D. Paulsen, U. L. Osterberg, B. W. Pogue, andM. S. Patterson, ‘‘Simultaneous reconstruction of optical ab-sorption and scattering maps in turbid media from near-infrared frequency-domain data,’’ Opt. Lett. 20, 2128–2130(1995).

11. S. B. Colak, D. G. Papaioannou, G. W.’t Hooft, M. B. van derMark, H. Schomberg, J. C. J. Paasschens, J. B. M. Melis-sen, and N. A. A. J. van Asten, ‘‘Tomographic image recon-

struction from optical projections in light-diffusing media,’’Appl. Opt. 36, 181–213 (1997).

12. P. N. den Outer, Th. M. Nieuwenbuizen, and A. Lagendijk,‘‘Location of objects in multiple-scattering media,’’ J. Opt.Soc. Am. A 10, 1209–1218 (1993).

13. S. Feng, F. Zeng, and B. Chance, ‘‘Photon migration in thepresence of a single defect: a perturbation analysis,’’ Appl.Opt. 35, 3826–3836 (1995).

14. S. R. Arridge and J. C. Hebden, ‘‘Optical imaging in medi-cine: II. Modeling and reconstruction,’’ Phys. Med. Biol.42, 841–853 (1997).

15. J. C. Schotland, ‘‘Continuous-wave diffusion imaging,’’ J.Opt. Soc. Am. A 14, 275–279 (1997).

16. S. J. Norton and T. Vo-Dinh, ‘‘Diffraction tomographic im-aging with photon density waves: an explicit solution,’’ J.Opt. Soc. Am. A 15, 2670–2677 (1998).

17. S. A. Walker, D. A. Boas, and E. Gratton, ‘‘Photon densitywaves scattered from cylindrical inhomogeneities: theoryand experiments,’’ Appl. Opt. 37, 1935–1944 (1998).

18. S. Fantini, S. A. Walker, M. A. Franceschini, M. Kaschke,P. M. Schlag, and K. T. Moesta, ‘‘Assessment of the size, po-sition, and optical properties of breast tumor in vivo by non-invasive optical methods,’’ Appl. Opt. 37, 1982–1989 (1998).

19. J. C. Hebden, F. E. W. Schmidt, M. E. Fry, M. Schweiger, E.M. C. Hillman, D. T. Delpy, and S. R. Arridge, ‘‘Simulta-neous reconstruction of absorption and scattering imagesby multichannel measurement of purely temporal data,’’Opt. Lett. 24, 534–536 (1999).

20. S. R. Arridge, ‘‘Optical tomography in medical imaging,’’Topical Rev. Inverse Probl. 15, R41–R93 (1999).

21. M. Firbank, S. R. Arridge, M. Schweiger, and D. T. Delpy,‘‘An investigation of light transport through scattering bod-ies with non-scattering regions,’’ Phys. Med. Biol. 41, 767–783 (1996).

22. S. R. Arridge, H. Dehghani, M. Schweiger, and E. Okada,‘‘The finite element model for the propagation of light inscattering media: a direct method for domains with non-scattering regions,’’ Med. Phys. 27, 252–264 (2000).

23. S. Takahashi, D. Imai, and Y. Yamada, ‘‘Fundamental 3DFEM analysis of light propagation in head model toward 3Doptical tomography,’’ in Optical Tomography and Spectros-copy of Tissue: Theory, Instrumentation, Model and Hu-man Studies II, Proc. SPIE 2979, 130–138 (1997).

24. S. Takahashi and Y. Yamada, ‘‘Simulation of 3D lightpropagation in a layered head model including a clear CSFlayer,’’ in Advances in Optical Imaging and Photon Migra-tion, J. Fujimoto and M. S. Patterson eds., Vol. 21 of Trendsin Optics and Photonic Series (Optical Society of America,Washington, D.C., 1998).

25. M. S. Patterson, B. Chance, and B. C. Wilson, ‘‘Time re-solved reflectance and transmittance for the noninvasivemeasurement of tissue optical properties,’’ Appl. Opt. 28,2331–2336 (1989).

26. N. G. Chen and J. Bai, ‘‘Monte Carlo approach to modelingof boundary conditions for the diffusion equation,’’ Phys.Rev. Lett. 80, 5321–5324 (1998).

27. C. P. Gonatas, M. Miwa, M. Ishii, J. Schotland, B. Chance,and J. S. Leigh, ‘‘Effects due to geometry and boundary con-ditions in multiple light scattering,’’ Phys. Rev. E 48, 2212–2216 (1993).

28. I. Freund, ‘‘Surface reflections and boundary conditions fordiffusive photon transport,’’ Phys. Rev. A 45, 8854–8858(1992).

29. K. Furutsu, ‘‘Boundary conditions of the diffusion equationand applications,’’ Phys. Rev. A 39, 1386–1401 (1989).

30. J. Wu, F. Partovi, M. S. Field, and R. P. Rava, ‘‘Diffuse re-flectance from turbid media: an analytical model of photonmigration,’’ Appl. Opt. 32, 1115–1121 (1993).

31. S. Ito, ‘‘Diffusion of collimated, narrow beam waves in dis-crete random media,’’ Appl. Opt. 34, 7106–7112 (1995).

32. R. C. Haskell, L. O. Vaasand, T. Tsay, T. Feng, M. S. McAd-ams, and B. J. Tromberg, ‘‘Boundary conditions for the dif-fusion equation in radiative transfer,’’ J. Opt. Soc. Am. A11, 2727–2741 (1994).

33. R. Aronson, ‘‘Boundary conditions for diffusion of light,’’ J.Opt. Soc. Am. A 12, 2532–2539 (1995).

Ripoll et al. Vol. 17, No. 9 /September 2000 /J. Opt. Soc. Am. A 1681

34. D. J. Durian and J. Rudnick, ‘‘Spatially resolved back-scattering: implementation of extrapolation boundarycondition and exponential source,’’ J. Opt. Soc. Am. A 16,837–844 (1999).

35. A. D. Kim and A. Ishimaru, ‘‘Optical diffusion ofcontinuous-wave, pulsed, and density waves in scatteringmedia and comparisons with radiative transfer,’’ Appl. Opt.37, 5313–5319 (1998).

36. F. Martinelli, A. Sassaroli, G. Zaccanti, and Y. Yamada,‘‘Properties of the light emerging from a diffusive medium:angular dependence and flux at the external boundary,’’Phys. Med. Biol. 44, 1257–1275 (1999).

37. A. Ishimaru, Wave Propagation and Scattering in RandomMedia (Academic, New York, 1978), Vol. I.

38. J. Ripoll and M. Nieto-Vesperinas, ‘‘Scattering integralequations for diffusive waves. Detection of objects buriedin diffusive media in the presence of rough interfaces,’’ J.Opt. Soc. Am. A 16, 1453–1465 (1999).

39. J. A. Sanchez-Gil and M. Nieto-Vesperinas, ‘‘Light scatter-ing from random rough dielectric surfaces,’’ J. Opt. Soc. Am.A 8, 1270–1286 (1991).

40. A. A. Maradudin, T. Michel, A. R. McGurn, and E. R.Mendez, ‘‘Enhanced backscattering of light from a randomgrating,’’ Ann. Phys. (N.Y.) 203, 255–307 (1990).

41. S. R. Arridge, M. Schweiger, M. Hiraoka, and D. T. Delpy,‘‘A finite element approach for modeling photon transportin tissue,’’ Med. Phys. 20, 299–309 (1993).

42. M. Schweiger, S. R. Arridge, M. Hirakoa, and D. T. Delpy,

‘‘The finite element model for the propagation of light inscattering media: boundary and source conditions,’’ Med.Phys. 22, 1779–1792 (1995).

43. S. R. Arridge, ‘‘Photon-measurement density functions.Part I: Analytical forms,’’ Appl. Opt. 34, 7395–7409(1995).

44. J. Ripoll and M. Nieto-Vesperinas, ‘‘Index mismatch for dif-fuse photon density waves at both flat and rough diffuse–diffuse interfaces,’’ J. Opt. Soc. Am. A 16, 1947–1957(1999).

45. M. Born and E. Wolf, Principles of Optics, 6th ed. (Perga-mon, New York, 1993).

46. G. B. Arfken and H. J. Weber, Mathematical Methods forPhysicists, 4th ed. (Academic, New York, 1995).

47. M. Nieto-Vesperinas, Scattering and Diffraction in PhysicalOptics (Wiley-Interscience, New York, 1991).

48. C. A. Brebbia and J. Dominguez, Boundary Elements, AnIntroductory Course (Computational Mechanics Publica-tions, McGraw-Hill, New York, 1989).

49. L. H. Wang, S. L. Jacques, and L. Q. Zheng, ‘‘MCML-MonteCarlo modeling of photon transport in multilayered tis-sues,’’ Comput. Methods Programs Biomed. 47, 131–146(1995).

50. R. Graaff, M. H. Koelink, F. F. M. de Mul, W. G. Zijlstra,A. C. M. Dassel, and J. G. Aarnoudse, ‘‘Condensed MonteCarlo simulations for the description of light transport,’’Appl. Opt. 32, 426–434 (1993).


Recommended