Open Access
14 July 2014 Analytical Green’s function for the fluorescence simplified spherical harmonics equations in turbid medium
Author Affiliations +
Abstract
It is more complicated to write the analytical expression for the fluorescence simplified spherical harmonics (SPN) equations in a turbid medium, since both the processes of the excitation and emission light and the composite moments of the fluence rate are described by coupled equations. Based on an eigen-decomposition strategy and the well-developed analytical methods of diffusion approximation (DA), we derive the analytical solutions to the fluorescence SPN equations for regular geometries using the Green’s function approach. By means of comparisons with the results of fluorescence DA and Monte Carlo simulations, we have shown the effectiveness of our proposed method and the expected advantages of the SPN equations in the case of small source–detector separation and high absorption.

In recent years, the simplified spherical harmonics (SPN) equations were applied for modeling light propagation in biological tissues.15 They were considered to effectively remedy the applicability of the diffusion approximation (DA) and reduce the complexity of solving the full spherical harmonics (PN) or the discrete ordinates (SN) equations. In a previous paper, we derived the analytical solutions to the SPN equations in turbid medium.3 It will thus be useful to proceed with its extension for studying the response of fluorophores, since fluorescence spectroscopy/imaging has some important applications in the field of biomedical optics.46

In this study, we derive the analytical solutions to the fluorescence simplified spherical harmonics (FSPN) equations based on the Green’s function method in the homogeneous turbid medium. The eigen-decomposition strategy developed in our previous paper and the analytical methods to fluorescence diffusion approximation (FDA) proposed by Yalavarthy and Ayyalasomayajula6 are used. The analytical method validation is presented against the Monte Carlo (MC) simulations and FDA data.

The principle of the SPN equations can be extended to the FSPN equations as a two-stage system of excitation and emission. Here, the FSP3 equations in the steady-state domain are given by the following system of coupled partial differential equations:

Eq. (1)

·13μa1(ν)φ1(ν)(r)+μa0(ν)φ1(ν)(r)23μa0(ν)φ2(ν)(r)=S(ν)(r)·17μa3(ν)φ2(ν)(r)+(49μa0(ν)+59μa2(ν))φ2(ν)(r)(23μa0(ν))φ1(ν)(r)=23S(ν)(r),
where μai(v)=μa0(v)+μs(v)(1gi) is the absorption coefficient of order i, g is the anisotropy factor, μs(v) is the scattering coefficient, and subscript v represents excitation x or emission m. ϕi(v)(r) is the light fluence rate, and S(v)(r) is the light source. For the excitation process, S(x)(r) denotes the excitation light source. For the emission process, S(m)(r)=ϕ(x)(r)Nr(r), where ϕ(x)=ϕ1(x)(2/3)ϕ2(x) is the total excitation light fluence rate, and Nr(r) is the fluorescence yield distribution, which is the product of the quantum efficiency η and the absorption coefficient of the fluorophore μaf. For FSPN equations, the partially reflective boundary conditions can be used.1

In the homogeneous medium with a spatially uniform distribution of fluorophores, Eq. (1) can be rewritten as

Eq. (2)

(2A(ν))Φ(ν)(r)=ε(ν)S(ν)(r),
where Φ(v)(r)=[ϕ1(v)(r)ϕ2(v)(r)]T, ε(v)=[ε1(v)ε2(v)]T, with the coefficient matrix
A(ν)=(3μa0(ν)μa1(ν)2μa0(ν)μa1(ν)143μa0(ν)μa3(ν)289μa0(ν)μa3(ν)+359μa2(ν)μa3(ν))
and the vector ε(v)=[3μa1(v)14/3μa3(v)]T.

By our proposed transformation matrix Φ(v)=B(v)Φ(v) and ε(v)=B(v)1ε(v), where B is the eigenvector and λ1(v)2,λ2(v)2 are the two real eigenvalues of coefficient matrix A(v),3 the excitation and emission equations can be reduced to the form of a system of independent differential equations that are mathematically consistent with the FDA equation

Eq. (3)

(2λi(x)2)φi(x)(r)=εi(x)S(x)(r)(2λi(m)2)φi(m)(r)=εi(m)Nr[φ1(x)(r)(2/3)φ2(x)(r)].

In Ref. 6, the Green’s functions for the time-domain and frequency-domain FDA have been derived in detail. Here, we can further write the universal formula of the steady-state Green’s function for regular geometries

Eq. (4)

ggeoψfl(r,r)=Nrζ2γm2γx2[γx2ggeoψx(r,r)γm2ggeoψm(r,r)],
where ggeoψfl(rr) represents the Green’s function of the emission light fluence rate, with the isotropic source δ(rr) placed at r, and the subscript “geo” represents the relevant geometry under consideration. γx,m2=c/[3(μax,am+μsx,sm)], μax,am and μsx,sm are the absorption and reduced scattering coefficients, respectively, and c is the speed of light in the medium. ζ2=(γm2γx2)/(γm2μaxcγx2μamc), ggeoψx,m(rr) are the Green’s functions evaluated by substituting μax and μam into excitation and emission DA equations, respectively.

Based on Eqs. (3) and (4), we are encouraged to further derive the Green’s functions for the FSP3 equations. For the excitation equation, by applying the transformation relationship of Φ(x)=B(x)Φ(x), the Green’s function of the total excitation light fluence rate is given by

Eq. (5)

g(x)=g1(x)23g2(x)=j=12B1j(x)gj(x)23i=12B2j(x)gj(x).

For the emission equation, using the result of Eq. (4) and the principle of linear superposition theory, we can write the Green’s function of ϕi(m) as

Eq. (6)

gi(ϕfl)=j=12Nrλi(m)2λj(x)2(εi(m)B1j(x)gj(x)εj(x)gi(m))23j=12Nrλi(m)2λj(x)2(εi(m)B2j(x)gj(x)εj(x)gi(m)),
where B1j(x,m) and B2j(x,m) denote the entry of B(x,m) at the first and second row and the j’th column, respectively, gj(x) and gj(m) are the Green’s functions evaluated by substituting λj(x), εj(x), and λi(m), εi(m) into Eq. (3).

The Green function of the total emission light fluence rate for the FSP3 equations can be obtained with

Eq. (7)

g(ϕfl)=g1(ϕfl)23g2(ϕfl)=j=12B1j(m)gj(ϕfl)23j=12B2j(m)gj(ϕfl).

The exiting flux at the boundary for the FSP3 equations can be written as1

Eq. (8)

Γ(ν)=(14+J0)(φ1(ν)23φ2(ν))(0.5+J13μa1(ν))n·φ1(ν)+(516+J2)(13φ2(ν))(J37μa3(ν))n·φ2(ν),
where the constants Ji(i=0,1,2,3) are explicitly given in Ref. 1.

From the above derivation, we note that if gi(x,m) are given, the Green’s functions for the FSP3 equations can be readily obtained. As with our previous paper, the analytical solutions for infinite medium and two-dimensional (2-D) circle are given in the following section.

In the case of an infinite homogeneous medium excited by an isotropic point source, the Green’s functions gi(x,m) can be expressed as3,7

Eq. (9)

gi(x,m)=εi(x,m)eλi(x,m)r/4πr.

The Green’s functions gi(x,m) of 2-D circular domain which represents an infinite cylindrical medium excited by anaxially parallel and infinite isotropic line source at r can be expressed as3,7

Eq. (10)

gi(x,m)(r)=n=[an(x,m)jIn(rλj(x,m))]cos(nθ)+n=[εj(x,m)2πIn(rλj(x,m))Kn(rλj(x,m))]cos(nθ)(r>r),
where In and Kn are the modified Bessel functions of the first and second kinds of n’th respectively, r is the radius of circle, an(x,m)j are unknown coefficients which can be determined by boundary conditions.1

To demonstrate the effectiveness of the proposed method, the analytical solutions to the FSP3 equations are compared with FDA data and MC simulations. The turbid medium under consideration is homogeneous and the entire medium can act as a source for fluorescent light. In the following comparative investigations, we mainly consider the effect of the different absorption coefficients in biological tissue, and keep the others to be constant: the scattering coefficient μs=10mm1, the fluorophore absorption coefficient μaf=0.003, the quantum efficiency of the fluorophore η=0.23, the anisotropy factor g=0.9, and the refractive index n=1.4.

First, the analytical solution to the FSP3 equations in an infinite turbid medium with relatively weak absorption μax=0.03mm1 and μam=0.02mm1 was compared with the FDA calculations and MC simulations. The fluorescence fluence rate versus source–detector distance r (mm) and a plot of the percentage error [(ϕflϕMC)/ϕMC×100%] are shown in Figs. 1(a1) and 1(a2), respectively. As expected, when the source–detector separation is large, both the FSP3 and FDA solutions show an excellent agreement with the MC data. However, when the detector is placed near the isotropic source, the FSP3 shows smaller model error than FDA. Next, a similar comparison was performed for a highly absorbing medium. We increased the absorption coefficient to μax=2mm1 and μam=1mm1. We see that the error of the FDA solution is significantly increased; however, the FSP3 solutions are still in a relatively good agreement with the MC in Figs. 1(b1) and 1(b2).

Fig. 1

Steady-state fluence rates versus the source–detector distance for an infinite geometry with μax=0.03mm1 and μam=0.02mm1 (a1) and μax=2mm1 and μam=1mm1 (b1), calculated from the FSP3, FDA, and MC, respectively, and corresponding model errors (a2), (b2).

JBO_19_7_070503_f001.png

In 2-D circular domain, the radius of circle is r=5mm, the isotropic point light source is placed at (r1/μa1, 0 deg) and 15 detectors are distributed at equal spacing from (r, 22.5 deg) to (r, 337.5 deg). First, we compared the exiting fluorescence flux from the FSP3, FDA, and MC calculations with the small absorption coefficients μax=0.03mm1 and μam=0.02mm1. Figures 2(a1)and 2(a2) describe the exiting fluorescence flux from the boundary versus the detection angle and a plot of the percentage error ((ΓflΓMC)/ΓMC×100%), respectively. Since the absorption coefficient is small, it is natural to observe that the results obtained from FSP3 and FDA are close to MC data, and the maximum model error is within 6.7%. Then we increased the absorption coefficient to μax=0.03mm1 and μam=0.02mm1. As can been seen in Figs. 2(b1) and 2(b2), the error of FDA is more obvious than FSP3, and for the detector at 180 deg the FSP3 and FDA errors reach a maximum of about 29.5% and 16.8%, respectively.

Fig. 2

The exiting fluorescence flux from two-dimensional circle boundary versus the detection angle, with absorption coefficients of μax=0.03mm1, μam=0.02mm1 (a1) and μax=0.3mm1, μax=0.2mm1 (b1), calculated from the FSP3, FDA, and MC, and corresponding model errors (a2), (b2).

JBO_19_7_070503_f002.png

In conclusion, we have derived the analytical Green’s function for the FSP3 equations for the homogeneous turbid medium and shown the effectiveness of our proposed method. As expected, compared with the FDA, the FSP3 equations show the advantages in the case of the small source–detector separation and high absorption. To the authors’ knowledge, this is the first time that the analytical solutions are presented. Although the analytical Green’s functions derived in this paper are limited to the specific geometries (infinite medium and 2-D circle), the proposed derivation process is principly concise and universally extendable to other regular geometries. Moreover, the method is applicable for an arbitrary order of FSPN equations.

Acknowledgments

The authors acknowledge the funding supports from the National Natural Science Foundation of China (81101106, 61108081, 81271618, and 81371602), Research Fund for the Doctoral Program of Higher Education of China (20110032120069 and 20120032110056), and the Natural Science Foundation of Tianjin, China (13JCZDJC28000 and 14JCQNJC14400).

References

1. 

A. D. KloseE. W. Larsen, “Light transport in biological tissue based on the simplified spherical harmonics equations,” J. Comput. Phys., 220 (1), 441 –470 (2006). http://dx.doi.org/10.1016/j.jcp.2006.07.007 JCTPAH 0021-9991 Google Scholar

2. 

J. Bouza DomínguezY. J. Bérubé-Lauzière, “Diffuse optical tomographic imaging of biological media by time-dependent parabolic SPN equations: a two-dimensional study,” J. Biomed. Opt., 17 (8), 086012 (2012). http://dx.doi.org/10.1117/1.JBO.17.8.086012 JBOPFO 1083-3668 Google Scholar

3. 

L. Zhanget al., “Analytical solutions to the simplified spherical harmonics equations using eigen decompositions,” Opt. Lett., 38 (24), 5462 –5465 (2013). http://dx.doi.org/10.1364/OL.38.005462 OPLEDP 0146-9592 Google Scholar

4. 

Y. B. Lauziereet al., “Light propagation from fluorescent probes in biological tissues by coupled time-dependent parabolic simplified spherical harmonics equations,” Biomed. Opt. Exp., 2 (4), 817 –837 (2011). http://dx.doi.org/10.1364/BOE.2.000817 BOEICL 2156-7085 Google Scholar

5. 

Y. Luet al., “A parallel adaptive finite element simplified spherical harmonics approximation solver for frequency domain fluorescence molecular imaging,” Phys. Med. Biol., 55 (162010), 4625 –4645 http://dx.doi.org/10.1088/0031-9155/55/16/002 PHMBA7 0031-9155 Google Scholar

6. 

K. R. AyyalasomayajulaP. K. Yalavarthy, “Analytical solutions for diffuse fluorescence spectroscopy/imaging of biological tissues in regular geometries. Part I: zero and extrapolated boundary conditions,” J. Opt. Soc. Am. A, 30 (3), 537 –552 (2013). http://dx.doi.org/10.1364/JOSAA.30.000537 JOAOD6 0740-3232 Google Scholar

7. 

S. R. Arridgeet al., “The theoretical basis for the determination of optical pathlengths in tissue: temporal and frequency analysis,” Phys. Med. Biol., 37 (71992), 1531 –1560 http://dx.doi.org/10.1088/0031-9155/37/7/005 PHMBA7 0031-9155 Google Scholar
© 2014 Society of Photo-Optical Instrumentation Engineers (SPIE) 0091-3286/2014/$25.00 © 2014 SPIE
Limin Zhang, Xi Yi, Jiao Li, Huijuan Zhao, and Feng Gao "Analytical Green’s function for the fluorescence simplified spherical harmonics equations in turbid medium," Journal of Biomedical Optics 19(7), 070503 (14 July 2014). https://doi.org/10.1117/1.JBO.19.7.070503
Published: 14 July 2014
Lens.org Logo
CITATIONS
Cited by 3 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Luminescence

Absorption

Spherical lenses

Sensors

Biological research

Surface plasmons

Diffusion

Back to Top