Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Light transport in homogeneous tissue with m-dependent anisotropic scattering I: Case’s singular eigenfunctions solution and Chandrasekhar polynomials

Open Access Open Access

Abstract

This paper is the first of two deriving the analytical solutions for light transport in infinite homogeneous tissue with an azimuth-dependent (m-dependent) anisotropic scattering kernel by two approaches, Case’s singular eigenfuncions expansion and Fourier transform, as well as proving the consistence of the two solutions. In this paper, Case’s method was applied and extended to the general m-dependent anisotropic scattering case. The explicit Green’s function of radiance distributions, which was regarded as the comparative standard for the equivalent solution via Fourier transform and inversion in our second accompanying paper, was expanded into a complete set of the discrete and continuous eigenfunctions. Considering that the two kinds of m-dependent Chandrasekhar orthogonal polynomials that play vital roles in these analytical solutions are very sensitive to the typical optical parameters of biological tissue as well as the degrees or orders, four numerical evaluation methods were benchmarked to find the stable, reliable and feasible numerical evaluation methods in high degrees and high orders.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Though for decades many approximation methods, such as the spherical harmonics (PN) [1, 2], the simplified spherical harmonics (SPN) [3–6], the diffusion equation (DE) [7–10], FN method [11, 12], etc., are applied to deal with the light transport equations and make great success, the traditional but effective Case’s singular eigenfunctions (CSEs) expansion solutions [13] are still regarded as the exact analytical solutions and as the criteria for testing the approximate methods. Since K. M. Case reported the singular eigenfuntions expansion method of building the elementary solutions for the radiative transport equation (RTE) and proved the orthogonality and completeness of the singular eigenfunctions in isotropic scattering [13], the Case’s method has become the hot hits and soon extended to linear anisotropic scattering [14, 15], general anisotropic scattering with [16, 17] and without [18] azimuthal symmetry. Due to its complexity, the numerical approximation was the main efforts in quite a long time. In recent years, the Case’s method was revisited by some researchers [19–21] to seek the more “exact” solutions in biological optics, neutron transport theory, heat transport theory, and so on.

As for the analytical solutions, the Case’s method and Fourier transform method are both indispensable. For the regular infinite space, the solutions obtained from the two methods should be consistent with each other. In fact, K. M. Case [22] applied the Fourier transform method to linear transport theory early. B. D. Ganapol [23–26] explicitly proved the consistence between the CSEs solution and the Fourier transform solution in the casem=0. The purpose of our two papers is to extend the consistent theory suitable for the neutron transport to the azimuth-dependent or m-dependent light transport in biological tissue. So in our first paper, we firstly derived the Green’s function based on the CSEs expansion for the light transport in infinite homogeneous tissue with m-dependent anisotropic scattering kernel and made a comparative standard for the Fourier transform solution which is to be discussed in our second accompanying paper.

In Case’s method, the Chandrasekhar orthogonal polynomials occupy the most important and fundamental position undisputedly. Without the orthogonal polynomials, it is impossible to build the eigenfunctions and elementary solutions, or to find the discrete eigenvalues. Inönü [27] and K. M. Case [28, 29] discussed the polynomials thoroughly and laid a substantial mathematical foundation for successive researchers. C. E. Siewert, etc [30–32], proved many useful identities about Chandrasekhar polynomials and gave some feasible methods to compute the polynomials with high orders and high degrees for the sake of finding the discrete eigenvalues. In fact, in our reachable scope, there are not many literatures on the Chandrasekhar polynomials, especially in the evaluations of the polynomials. Recently, Ganapol [33] focused on the Chandrasekhar polynomials and benchmarked the various computational methods for zero order,m=0. However, our numerical evaluation experiments showed that in biological optics, unlike the neutron transport domain or the extremely critical testing parameters, the Chandrasekhar orthogonal polynomials are very sensitive to the optical parameters of biological tissue, such as single particle albedo ϖand anisotropy factorg. So in this paper, we extended Ganapol’s methods in the case m=0to benchmark the m-dependent Chandrasekhar polynomials with the typical optical parameters of biological tissue to find the stable, reliable, feasible numerical evaluation methods in high degrees and high orders.

2. Case’s singular eigenfunctions solution

2.1 Eigenvalues and singular eigenfunctions

The light transport in the infinite homogeneous tissue with an m-dependent anisotropic scattering kernel is governed by integro-differential equation [11, 12, 34, 35]:

μψ(x,μ,φ;μ0,φ0)x+μtψ(x,μ,φ;μ0,φ0)=μs02π1+1p(cosγ)ψ(x,μ,φ;μ0,φ0)dμdφ+δ(x)δ(μμ0)δ(φφ0)
where ψ(x,μ,φ;μ0,φ0) is the radiance of the scattering light which only depends on the position coordinatex, the direction cosineμmeasured with the angle between the propagating radiation and the positive xaxis, and the azimuthal angle φ. μ0, φ0are the direction cosine and azimuthal angle of the source positioned atx=0, respectively. μsis the scattering coefficient, μa is the absorption coefficient, and μt=μa+μs is the total attenuation coefficient [35]. p(cosγ) is the scattering phase function where γis the scattering angle. We assume that the scatters are spherical symmetric and express the phase function as spherical harmonics expansion [21, 34]:
p(cosγ)=14πl=0L(2l+1)flm=ll(lm)!(l+m)!Plm(μ)Plm(μ)eim(φφ)
where Lis the truncated position of the scattering kernel and L>1.f0=1,0<fl<1 for1lL, and fl=0 forl>L. In biological tissue optics, the Henyey-Greenstein anisotropic scattering kernel is often applied and then fl=gl [35]. The anisotropy factor gis the average cosine of angle of scattering [35]. Plm(μ) is the associated Legendre functions of the first kind [36].

To develop the singular eigenfunctions solution of Eq. (1), we expand the ψ(x,μ,φ;μ0,φ0) as Fourier series [18]

ψ(x,μ,φ;μ0,φ0)=12πm=ψm(x,μ;μ0)eimφ

Substituting Eq. (3) and δ(φφ0)=12πm=eim(φφ0) into Eq. (1), we get

μψm(x,μ;μ0)x+μtψm(x,μ;μ0)=μs2l=|m|LβlmPlm(μ)1+1Plm(μ)ψm(x,μ;μ0)dμ+δ(x)δ(μμ0)eimφ0
where βlm=(lm)!(l+m)!(2l+1)fl.Thus, according to the Case’s method [13] and separation of variables in homogeneous equation of Eq. (4), the solution in full-space has the form:
ψm(x,μ)exp(μtx/v)ϕm(v,μ)
where vis the eigenvalues and ϕm(v,μ) is the corresponding eigenfunctions. Substituting Eq. (5) into the homogeneous equation of Eq. (4), we get the equation forϕm(v,μ):
(1μv)ϕm(v,μ)=ϖ2l=|m|LβlmPlm(μ)1+1Plm(μ)ϕm(v,μ)dμ
where ϖ=μs/μtis the single particle albedo, and it is obvious that 0<ϖ<1in absorbing tissue, and ϖ=1 when μa=0 in conservative tissue.

Considering the Hobson definition [36] of the associated Legendre functions of the first kind which is used by the mathematical computing softwares, such as Matlab, Mathematica, etc., we define

glm(v)(1)m1+1Plm(μ)ϕm(v,μ)dμ
glm(v) is the Chandrasekhar polynomials that we will discuss in Section 3.

Then, whenv[1,1], the discrete eigenfunctions are

ϕm(v,μ)=ϖv2gm(v,μ)vμ
and satisfy the normalization condition
11ϕm(v,μ)(1μ2)|m|/2dμ=1
where

gm(v,μ)=l=|m|Lβlmglm(v)Plm(μ)

So we define the function

Λm(z)1ϖz211gm(z,μ)zμ(1μ2)|m|/2dμ
and get the discrete paired real eigenvalues ±vjm(1jM) which are determined by the roots ofΛm(±vjm)=0, Mis the number of positive eigenvalues [29]. Λm(z)is a holomorphic function in the complex z-plane excluding a cut line −1 to 1 on the real axis. When zapproaches the cut line from above(+) and below(), by Plemelj’s formula [37], we get
[Λm(v)]±=1ϖv2P11gm(v,μ)vμ(1μ2)|m|/2dμ±iπϖv2gm(v,v)(1v2)|m|/2
Here P denotes the Cauchy principal value.

The CSEs for continuous spectrum v(1,1) are

ϕm(v,μ)=ϖv2Pgm(v,μ)vμ+λ(v)(1v2)|m|/2δ(vμ)
where

λ(v)=1ϖv2P11gm(v,μ)vμ(1μ2)|m|/2dμ

From Eq. (12), we get

[Λm(v)]++[Λm(v)]=2λ(v)
[Λm(v)]+[Λm(v)]=λ2(v)+[πϖv2gm(v,v)(1v2)|m|/2]2

2.2 Orthogonality and norms

The orthogonality and full/partial-range completeness properties of the discrete and continuous eigenfunctions for isotropic scattering [13], for linear anisotropic scattering [14, 15], and for general anisotropic scattering [16–18] are discussed in many early classical literatures. In our case the orthogonality and completeness are easy to be extended from the general anisotropic scattering case. However, the norms of the discrete and continuous eigenfunctions are rarely expressed explicitly.

The discrete and continuous eigenfunctions for specific m satisfy the following orthogonalities:

11ϕm(vim,μ)ϕm(vjm,μ)μdμ=N0m(vjm)δij
11ϕm(v,μ)ϕm(v,μ)μdμ=Nm(v)δ(vv)
11ϕm(vj,μ)ϕm(v,μ)μdμ=0
where vim,vjmare any discrete eigenvalues, v,vare in continuous spectrum,v,v(1,1).δij is the Kronecker delta, δij=1for i=j, and otherwiseδij=0.δ(vv) is the Dirac delta.N0m(vjm) is the norms of the discrete eigenfunctions
N0m(±vjm)=±ϖvjm2[ϖ(vjm)2l=|m|Lβlmglm(vjm)k=|m|Lβkmgkm(vjm)γlmδm0(vjm)21l=|m|Lβlm[glm(vjm)]2(1+δm0)vjml=|m|Lβlmglm(vjm)k=0[l12](2l4k1)gl2k1m(vjm)]
where δm0 is the Kronecker delta, δm0=1 for m=0, and otherwiseδm0=0. γlm=1 for l+k is even, and γlm=vj for l+k is odd. [(l1)/2]denotes the nearest integer less than or equal to(l1)/2 . The proof of N0m needs the integral relation for zthat is not on the cut line −1 to 1 on the real axis [38]
11Plm(μ)Pkm(μ)zμdμ={2Plm(z)Qkm(z)lk2Qlm(z)Pkm(z)l>k
and the expansion of the directive of Plm(x)
dPlm(x)dx=1+δm02k=0[l12](2l4k1)Pl2k1m(x)
Qlm(z)is the associated Legendre functions of the second kind. Whenm=0, it is easy to validate that Eq. (20) is really identical to the results by Mika [16] and by Van Den [17] for m-independent general anisotropic scattering. N0m(±vjm) can also be expressed as [39, 40]
N0m(±vjm)=±12ϖ(vjm)2gm(vjm,vjm)[(vjm)21]|m|/2dΛm(z)dz|z=vjm
Nm(v) in Eq. (18) is the norm of eigenfuncions for continuous spectrum [39]

Nm(v)=v[Λm(v)]+[Λm(v)](1v2)|m|

2.3 Singular eigenfunctions solution

Now, by Case’s elementary solution theory [13], we can write the singular eigenfunctions elementary solution for homogeneous equation of Eq. (4):

ψm(x,μ)=j=1Ma+mϕm(vjm,μ)exp(μtx/vjm)+j=1Mamϕm(vjm,μ)exp(μtx/vjm)+11Am(v)ϕm(v,μ)exp(μtx/v)dv

Considering the boundary condition

ψm(0+,μ)ψm(0,μ)=1μδ(x)δ(μμ0)eimφ0
and the subcritical tissue, henceϖ<1, the radiance must vanish at infinity:

lim|x|ψm(x,μ)=0

So the solution must be separated into two half spaces:

ψm(x,μ;μ0)={j=1Ma+mϕm(vjm,μ)exp(μtx/vjm)+01Am(v)ϕm(v,μ)exp(μtx/v)dv(x>0)j=1Mamϕm(vjm,μ)exp(μtx/vjm)10Am(v)ϕm(v,μ)exp(μtx/v)dv(x<0)

Using the orthogonalities Eq. (17) ~Eq. (19), the coefficients a±mand Am(v)can be easily deduced:

aj±m=1N0m(±vjm)ϕm(±vjm,μ0)eimφ0
A(v)=1Nm(v)ϕm(v,μ0)eimφ0

So in the infinite homogeneous tissue with m-dependent anisotropic scattering, the component of the Green’s function for specific m can be rewritten as

ψm(x,μ;μ0)=eimφ0{j=1Mϕm(vjm,μ0)ϕm(vjm,μ)N0m(vjm)exp(μtx/vjm)+01ϕm(v,μ0)ϕm(v,μ)Nm(v)exp(μtx/v)dv(x>0)j=1Mϕm(vjm,μ0)ϕm(vjm,μ)N0m(vjm)exp(μtx/vjm)10ϕm(v,μ0)ϕm(v,μ)Nm(v)exp(μtx/v)dv(x<0)

From Eq. (3) we can get the expected final result Green’ function for Eq. (1). This singular eigenfunctions solution is the comparative standard for the Fourier transform solution which will be discussed in our second accompanying paper.

3. Chandrasekhar polynomials

The Chandrasekhar polynomials glm(z)play a vital role in discrete eigenvalues finding and the numerical evaluation of the solution, so many researchers had studied them for decades. In this section we revisited the Chandrasekhar polynomials for m0 and analyzed the stability of various computational method.

3.1 Definitions and recurrences of polynomials

To avoid confusing, following the naming conventions for Legendre polynomials, gl(z) and ρl(z) are called the Chandrasekhar polynomials of the first and second kinds respectively whenm=0, glm(z) and ρlm(z) the associated Chandrasekhar polynomials of the first and second kinds respectively whenm>0.

Let’s back to the definition the associated Chandrasekhar polynomials of the first kind in Eq. (7). Obviously, whenl=m, we get the first term for specificm,gmm(z)=(2m1)!!。Due to the attributes ofPlm(x), glm(z) satisfies the following relations:

glm(z)=(1)m(lm)!(l+m)!glm(z)
glm(z)=(1)lmglm(z)

From Eq. (6), the famous three-term recurrence relation is emerged

zhlglm(z)(lm+1)gl+1m(z)(l+m)gl1m(z)=0
where hl=(2l+1)(1ϖfl),m0,lm. Whenm=0, Eq. (34) is the three-term recurrence relation for the associated Chandrasekhar polynomials of the first kind.

The matrix generated from Eq. (34) is not a real symmetric matrix form which is not convenient to compute the eigenvalues, so the normalized associated Chandrasekhar polynomials of the first kind are introduced

g^lm(z)=(lm)!(l+m)!glm(z)
and then Eq. (34) changes into
zhlg^lm(z)(l+1)2m2g^l+1m(z)l2m2g^l1m(z)=0
with the initial term

g^mm(z)=(2m1)!!(2m)!=(2m)!2mm!

Similarly we define the associated Chandrasekhar polynomials of the second kind ρlm(z) by three-term recurrence relations

zhlρlm(z)(lm+1)ρl+1m(z)(l+m)ρl1m(z)=ρm+1m(z)δlm
with initial two terms ρmm(z)=0 andρm+1m(z)=z/(2m1)!! . So the normalized associated Chandrasekhar polynomials of the second kind
ρ^lm(z)=(lm)!(l+m)!ρlm(z)
satisfy the following three-term recurrence relations

zhlρ^lm(z)(l+1)2m2ρ^l+1m(z)l2m2ρ^l1m(z)=2m+1ρ^m+1m(z)δlm

The Christoffel–Darboux (C-D) identity or Liouville-Ostrogradski (L-O) formula [27, 33, 34] between glm(z) and ρlm(z) when m0 are

gl1m(z)ρlm(z)glm(z)ρl1m(z)=(l+m1)!(lm)!ρm+1m(z)gmm(z)(2m)!

So for the normalized g^lm(z)and ρ^lm(z) we have

g^l1m(z)ρ^lm(z)g^lm(z)ρ^l1m(z)=(2m+1)ρ^m+1m(z)g^mm(z)l2m2

Equation (41) and Eq. (42) play an indispensable part in the following computation methods and in the derivation of the Fourier transform solution. It is noted that the initial terms of two kinds of polynomials, actually may be any nonzero constants. As shown in the following sections, we will letg^mm(z)=1 without loss of generality. The first term ofρ^lm(z),ρ^mm(z), is usually set to zero, and the initial nonzero term, ρ^m+1m(z), satisfies ρ^m+1m(z)g^mm(z)=z for convenience.

3.2 Determinant and eigenvalues of the characteristic matrix

The Chandrasekhar polynomials of the first and second kinds were expressed as the form of determinant by Inönü [27], Ganapol [23]. The explicit determinantal representation of the associated Chandrasekhar polynomials of first kind glm(z) was given by McCormick [32, 39]. For lower degrees or lower orders approximation evaluating the polynomials by determinant is still a good choice. However, they are inconvenient to compute and analyze due to their asymmetries. The associated Chandrasekhar polynomials of the first and second kinds which have got a real symmetric determinantal form can be expressed as

g^lm(z)=(2m)!g^mm(z)(lm)!(l+m)!|Al1m(z)|
ρ^lm(z)=(2m+1)!ρ^m+1m(z)(lm)!(l+m)!|Al1m+1(z)|
respectively, where matrix

Alm(z)=(zhm(m+1)2m2(m+1)2m2zhm+1zhl1l2m2l2m2zhl)

The subscript l denotes the maximum degree of h in the lower right corner of the matrix, and the superscript m denotes the minimum degree of h in the upper left corner of the matrix. |·| denotes the determinant of the matrix. It is obvious that g^lm(z) and ρ^lm(z) are both polynomials of lm degrees ofz.

Alm(z) is square matrix of orderlm+1, so

|Alm(z)|=(j=mlhj)|Blm+zI|
Where the characteristic matrix
Blm=(0bm+1mbm+1m0bm+2mbm+2m0blmblm0)
wherebjm=(j2m2)/(hj1hj),(j=m+1,m+2,,l) and I is lm+1 identity matrix. Because of the property of the real symmetry, the eigenvalues of Blm are real, simple, ± paired, andBlm can be easily converted to diagonal matrix.

Because the eigenvalues of characteristic matrix Blmwill be the key in our following evaluations and the discrete eigenvalues finding in Case’s method, here we focused on the distributions of eigenvalues. Firstly, we were interested in how the parameters ofBlm, such as the single particle albedo ϖ and anisotropy factorg, affect the greatest eigenvalues.

We set m from 0 to 30 with step 5 and built a 1001×1001 square matrix. Different from the neutron transport theory or the critical parameters given by Ganapol [33], in biological tissue, ϖandg are about 0.7 ~0.9(such as epidermis and dermis, g=0.78 [35]). So the 2-norms of the matrix were computed with anisotropy factor g from 0.7 to 1 with step 0.02 and ϖ=0.7,0.8,0.9 respectively.

As shown in Fig. 1, the greatest eigenvalues are very sensitive to the anisotropy factor g when g increases up to about 0.85. Furthermore, wheng0.85, the greatest eigenvalues are about 1 except inm=0 case. In fact, this may be a bad news. Inönü [27] and Machida [12] pointed out that the discrete eigenvalues in Case’s method are the zeroes of glm(z) asl. From another point of view, the discrete eigenvalues must be greater than 1. So in some biological tissue there is the potential danger that these expected discrete eigenvalues would not be found.

 figure: Fig. 1

Fig. 1 The 2-norms of the characteristic matrix vary with anisotropy factorg, the curves in each subfigure denote the 2-norms when m=0,5,10,15,20,25,30 from top to bottom

Download Full Size | PDF

Next, we benchmarked the distributions and variation tendency of the eigenvalues with the typical optical parameters of biological tissue. We letm=0, ϖ=0.8 and g=0.7,0.8,0.9without loss of generality, and then enlarged the characteristic matrix from 2×2square matrix to 500×500square matrix successively. That is to say, the degree of the polynomials varied from 2 to 500. For each square matrix, we computed its positive eigenvalues and sorted them in descending order. If the number of positive eigenvalues is less than 500, then fill the extra position with zeroes. Then we observed the variation tendency of the eigenvalues along the order of the square matrix increasing. Figure 2 shows the distributions and variation tendency of eigenvalues at some sampled positions, and each curve in each subfigure denotes the tendency of the eigenvalues with the same position but the different degrees of polynomials.

 figure: Fig. 2

Fig. 2 The distributions and variation tendency of the eigenvalues at some sampled positions, each curve in each subfigure corresponds to a certain sampled position

Download Full Size | PDF

Though for differentl, the number of eigenvalues of the matrix is, of course, different. But as shown in Fig. 2, the norms or the greatest eigenvalues (the top curve) will approach to a constant as l increases. So do the other eigenvalues. This may be a good news. For a very largel, letgl+1m(z)0, as shown in the following linear system of equations method, and then the eigenvalues of Bl+1m are just the zeroes of gl+1m(z) which are the discrete eigenvalues and can easily be found in the lower degrees of polynomials. This indicates that we can find Case’s discrete eigenvalues with a not much biggerl.

3.3 Computation of the polynomials

Similar to the case m=0 [33], there are various methods to compute the Chandrasekhar polynomials for high degrees or high orders. Here, we benchmarked the following four methods and compared the resulting sequences to evaluate the round off errors.

3.3.1 Computational method

  • (a) Determinant eigenvalues method

    Theoretically, Eq. (43), Eq. (44) can be used to evaluate g^lm(z) andρ^lm(z), but the factorials will soon lead to overflow or underflow for high orders or high degrees. We rewrite them as

    g^lm(z)=g^mm(z)j=0lm1hm+j(z+λg,jm)(j+1)(j+1+2m)
    ρ^lm(z)=ρ^m+1m(z)j=1lm1hm+j(z+λρ,jm)(j+1)(j+1+2m)

    where λg,jmis the jtheigenvalue of Bl1m and λρ,jm the jtheigenvalue ofBl1m+1. Equation (48) and Eq. (49) are optimized to fit the numerical computation because each term in the product is not too big or too small.

  • (b) Forward recurrence method

    The forward recurrence method is the most intuitive method of evaluating the polynomials because of the explicit form of the three-term recurrence relations. From Eq. (36) and Eq. (40), the forward recurrence formulas for g^lm(z) andρ^lm(z) are

    d^jm(z)=zhj1d^j1m(z)(j1)2m2d^j2m(z)j2m2

    where j=m+2,m+3,,N,m0 and d^jm(z) are g^jm(z) orρ^jm(z). With the appropriate initial terms the resulting sequences will be got without any difficulties.

  • (c) Backward recurrence method

    In the backward recurrence method, we set the degree l from m to a certain large integer N and force, ag^m+N+1m(z)0nd then evaluated all the g^lm(z) in the reverse direction. From Eq. (36), we get

    g^jm(z)=zhj+1g^j+1m(z)(j+2)2m2g^j+2m(z)(j+1)2m2

    wherej=m,m+1,,m+N1. Since g^Nm(z) is unknown, we define

    g^jm(z)=γjmg^m+Nm(z)

    So the coefficients γjm satisfy

    γjm(z)=zhj+1γj+1m(z)(j+2)2m2γj+2m(z)(j+1)2m2

    andγm+Nm(z)=1,γm+N+1m(z)=0. When the backward recurrence iterated toj=m without overflows or underflows, we get

    g^mm(z)=γmm(z)g^Nm(z)

    So all the terms will be got

    g^jm(z)=γjmγmm(z)g^mm(z)

    Since we get the resulting sequence of the associated Chandrasekhar polynomials of the first kindg^lm(z), the associated Chandrasekhar polynomials of the second kind ρ^lm(z) may be got by C-D identity Eq. (42) along forward direction. If we are mandatory to apply the backward recurrence toρ^lm(z) ignoring the reasonability and validity, the iteration should be end at j=m+1 becauseρ^mm(z)=0.

  • (d) Linear system of equations method

    In linear system of equations method, we still set the degreel from mto a certain large integer m+N and assembled the N+1 equations into a linear system

    [Bm+Nm(z)zI]Gm+Nm(z)=G¯m+N+1m(z)

    where vectorGm+Nm(z)=(am,am+1,,am+N)T, and vectorG¯m+N+1m(z)=(0,0,,cam+N+1)T. Here

    aj=hjg^jm(z),j=m,m+1,,m+N+1

    and c is a certain coefficient.

    Now, letg^m+N+1m(z)0, so Eq. (56) changes to homogeneous linear equations

    [Bm+NmzI]Gm+Nm(z)=0

    So, according to the theories of linear algebra, Eq. (58) has a nontrivial solution only when the determinant matrix of the coefficient equals to zero

    |Bm+NmzI|=0

    Then zi,(i=0,1,,N) that satisfy Eq. (59) are just the paired eigenvalues ofBm+Nm and corresponding Gm+Nm(zi) are eigenvectors. For variouszi, the ratio of aj to hjg^jm(zi) keeps a constant which can be determined by the known initial termg^mm(zi), so we get the approximation

    g^jm(zi)=hmhjajamg^mm(zi)

    Different from the other three methods, here we only get the approximation values of g^lm(z) at the specific eigenvaluezi.

3.3.2 Error measurements

From above four computational methods, the resulting sequences of the associated Chandrasekhar polynomials of the first and second kinds can be evaluated. However, as many other similar polynomials, such as Plm(z) orQlm(z), each computational method represents different stability for different z because of the inescapable round off errors.

To measure the errors when g^lm(z) and ρ^lm(z) are evaluated independently, such as method (a) and (b), we define the first residual vector

R1,jm(z)g^j1m(z)ρ^jm(z)g^jm(z)ρ^j1m(z)2m+1ρ^m+1m(z)g^mm(z)j2m2
where j=m+1,m+2,,l. If the computational method is stable, g^lm(z) and ρ^lm(z) will be evaluated precisely, and then they should confirm well to Eq. (42) and R1,jm(z) should be equal or very close to zero.

When g^lm(z) is evaluated firstly and then ρ^lm(z) by Eq. (42), we can define the second residual vector by the three-term recurrence of ρ^lm(z)

R2,jm(z)zhjρ^jm(z)(j+1)2m2ρ^j+1m(z)j2m2ρ^j1m(z)
where j=m+1,m+2,,l1. Similarly, when ρ^lm(z) is evaluated firstly and theng^lm(z)by Eq. (42), we define the third residual vector by the three-term recurrence of g^lm(z)

R3,jm(z)zhjg^jm(z)(j+1)2m2g^j+1m(z)j2m2g^j1m(z)

For the same kind of resulting sequences of g^lm(z) or ρ^lm(z)from different computational method, we define the absolute error vector

R3,jm(z)d^j,1m(z)d^j,2m(z)
where d^j,1m(z),d^j,2m(z)denotes the values of g^jm(z) or ρ^jm(z) from different computational method, and the relative error vector

R4,jm(z)d^j,1m(z)d^j,2m(z)d^j,1m(z)

After the residual or error vectors are got, we can measure the average of errors by

Dim(z)log|Ri,jm(z)|
or
D¯im(z)log{Ri,jm(z)[Ri,jm(z)]T}
wherei=1,2,3,4.

3.3.3 Numerical evaluation results

We have done a plenty of numerical experiments for the associated Chandrasekhar polynomials of the first and second kinds based on the previous computational methods and measured the errors by the previous measurements. In our numerical experiments, we set g^mm(z)1 andρ^mm(z)z.

  • 1) Method (a) and method (b) benchmark

    Based on the experience in the casem=0, firstly, we set a certain|z|<1. We letN=1000,m=15 then l=15,16,,1015 and let anisotropy factorg=0.8, single particle albedoϖ=0.9, and z0=0.48, and then the results are shown in Fig. 3.

As shown in Fig. 3 (a), the sequence of glm(z0), and (b), the sequence of ρlm(z0), even for a larger N, the two methods can complete successfully and the envelopes of two kinds of polynomials decrease as the degree l increases. Figure 3(c) shows the logarithmic absolute errors D3m(z0) of the two sequences of glm(z0) from method (a) and method (b) respectively. Figure 3(d) shows the D3m(z0) forρlm(z0).The extremely small errors show that the resulting sequences of the associated Chandrasekhar polynomials of the first/second kind computed by method (a) and (b) are almost identical. Figure 3(e) and (f) are logarithmic residual errors D1m(z0) of glm(z0) and ρlm(z0) respectively, where each polynomials is evaluated by method (a) and (b). It is apparent that they agree with C-D formula perfectly.

 figure: Fig. 3

Fig. 3 The result sequences and errors of the first and second kinds of Chandrasekhar polynomials as the degree (x-axis) increases whenz=0.48by method (a) or (b).

Download Full Size | PDF

Next, since when|z|<1, the method (a) and method (b) seem to be stable, we iterated z from 0.01 to 0.99 with step 0.02 and m from 0 to 30 with step 5. To save the computational time, we set N=599 and focused on the average errors measurement. For each pair of z andm, we computed glm(z)and ρlm(z) with the degree l from m to m+N by method (a) and (b), and then got four resulting sequences. For each method, we got the residual vector R1,jm(z),j=m,m+1,,m+N, and then the average residual errors . Figure 4(a) and (b) describe the changes of the average residual errors with z andm. When m=0, the errors is still extremely small even whenz1. So in Ref [33], Ganapol can safely conclude that the forward recurrence is stable for gl(z) when|z|<1. However, as shown in Fig. 4 (a) and (b), for a specificm, the errors will be increasing as zis increasing, and for a not very big orderm, the logarithmic average residual errors will deteriorate rapidly to under zero. That is to say, the resulting sequences do not confirm to the C-D formula at all. Similarly, we computed the absolute errorsR3,jm(z) by Eq. (64), and then the average absolute errors D¯3m(z) shown in Fig. 4 (c) and (d). The changes of the average absolute errors are similar to the average residual errors.

 figure: Fig. 4

Fig. 4 The average errors of the first (left) and second (right) kinds of Chandrasekhar polynomials as z from 0 to 1 with step 0.2 and m from 0 to 30 with step 5 when N=599

Download Full Size | PDF

As shown in Fig. 4, when |z|<1 and z is not approaching±1, for a not very bigm, we can conclude that method (a) and method (b) are stable for two kinds of polynomials. But we must be also clearly aware that in certain aforementioned cases the stabilities of these two methods are broken remarkably and we can’t determine that which method is unstable or both methods are unstable. Then the backward recurrence must come on stage. And when|z|>1, the two methods will soon lead to overflow, and whether they are stable is to be determined.

Since in a non-strict sense the method (a) and (b) are identical and stable when|z|<1, we can easily evaluate the Chandrasekhar polynomials with variable z and observe the complex curves to verify the abstract attributes of zeroes that summarized by Inönü [27] which are common to all orthogonal polynomials in general [36]. For example, as shown in Fig. 5, the zeroes of the five curves are interlacing, i.e., between two consecutive zeroes of g^lm(z), there must be one and only one zero of g^l1m(z), and vice versa.

 figure: Fig. 5

Fig. 5 The curves of the first (left) and second (right) kinds of Chandrasekhar polynomials as zfrom −1 to 1.g(l,m) denotesg^lm(z), and ρ(l,m) denotes ρ^lm(z)

Download Full Size | PDF

  • 2) Method (b) and method (c) benchmark

    When|z|<1, since in most cases the forward recurrence methods for two kinds of polynomials are stable and reliable, we introduced the backward recurrence and applied to the two kinds of polynomials to determine the stability of method (c) when|z|<1. We iteratedz from 0.01 to 0.99 with step 0.02 and m from 0 to 30 with step 5 as previous parameters and computed, for each kind of polynomials, the average logarithmic absolute errors of the two sequences from method (b) and (c) by Eq. (64) and Eq. (66). As shown in Fig. 6, for two kinds of polynomials, the average absolute errors are too large to be accepted even in the casem=0. So the backward recurrence method is not suitable for the two kinds of polynomials when|z|<1.

Thus, when|z|<1, for each kind of polynomials, we got two series of resulting sequences, one is from forward recurrence, and the other is from backward recurrence. So the average residual errors can be evaluated by Eq. (61) and Eq. (66) to determine how they confirm to C-D formula. The average residual errors for g^lm(z) and ρ^lm(z) which are both from method (a) are discussed before. The other combinations are shown in Fig. 7. As shown in Fig. 7, if the backward recurrence is applied to the polynomials of the second kind, the errors are still very large. It can also be seen that the backward recurrence method is not suitable for the second kind of polynomials when|z|<1. Though Fig. 7 (c) shows that the errors are very small, considering the results of Fig. 6 and comparing the resulting sequences, we still believe that the backward recurrence method is not suitable for the first kind of polynomials when|z|<1.

 figure: Fig. 6

Fig. 6 The average logarithmic absolute errors of the two sequences of the first (left) and second (right) kind of Chandrasekhar polynomials as z from 0 to 1 with step 0.2 and m from 0 to 30 with step 5 when N=599.

Download Full Size | PDF

 figure: Fig. 7

Fig. 7 The average logarithmic residual errors of the first and second kind of Chandrasekhar polynomials as z from 0 to 1 with step 0.2 and m from 0 to 30 with step 5 when N=599.

Download Full Size | PDF

Thus, it is obvious that the forward recurrence method is suitable for two kinds of polynomials when|z|<1, but the backward recurrence method is not. We went on making the benchmarks when|z|>1. When|z|>1, the benchmarks will tend to suffer from the parameters severely. When the backward recurrence is applied to the first kind of polynomials the evaluation will soon exceed the machine precision. So we must adjust the appropriate parameters, such asm,ϖ,z, and decrease N. We letN=251, z from 1.25 to 9.00 with step 0.25 and m from 0 to 30 with step 5 as previous parameters.

As shown in Fig. 8 (a), the backward recurrence is applied to g^lm(z) and ρ^lm(z), the very large errors show that the backward recurrence is not simultaneously suitable for two kinds of polynomials. So as shown in Fig. 8 (b), the combination, that the backward recurrence is applied to g^lm(z) and the forward recurrence is applied toρ^lm(z), is the best choice because of the extremely small errors.

 figure: Fig. 8

Fig. 8 The average logarithmic residual errors of the first and second kind of Chandrasekhar polynomials as z from 1.25 to 9.00 with step 0.25 and m from 0 to 30 with step 5 when N=251.

Download Full Size | PDF

Thus, we can safely conclude (not show) that (1) in the casem=0, the forward recurrence method is stable for the polynomials of the first kind only when |z|<1 and the backward recurrence is suitable for them only when|z|>1, (2) in the casem0, the forward recurrence method is only valid for the polynomials of the first kind when|z|<1 and not close to±1, and it should be aware that the biggerm, the worse the stability as zis approaching±1, so when |z|>1 or close to±1, the backward recurrence should be applied to them, (3) the forward recurrence is stable for the polynomials of the second kind for anyz and anym, and the backward recurrence is not suitable for the polynomials of the second kind.

Since the forward recurrence is a perfect evaluation method for the polynomials of the second kind, we may compute them firstly, and then compute the polynomials of the first kind by C-D formula. After numerical experiments, we found that when|z|<1, this method is equivalent to the method (a) and method (b), and when |z|>1it is equivalent to the method (c). But due to the rapid overflow or underflow, we don’t recommend this method.

  • 3) Method (d) benchmark

    From Eq. (59), the main task in method (d) is to find the eigenvalues of Bm+Nm, and the approximate values of the polynomials of the first kind at each eigenvalue zi(i=0,1,,N) can be evaluate by Eq. (60). The distributions and variation tendency of the eigenvalues are discussed in section 3.2. Here, the question is whether these approximate values are accurate for |zi|>1 or|zi|<1.

    Firstly, we noticed that the parameters N,m,ϖ,g had quite an influence on the arithmetic or the eigenvalues and eigenvectors of characteristic matrix. In some cases there are no eigenvalues greater than 1, and in other cases there are eigenvalues greater than 1 or less than 1, as indicated in section 3.2. After the eigenvectors were computed, we found that in many cases, when the eigenvalues are approaching the maximum and minimum eigenvalue, the corresponding eigenvectors have some zero components (or very close to zero), and the number of the zero components increases as the eigenvalues increase. These zero components is caused by the fact, as discussed in section 3.2, that many of the eigenvalues with the same position will tend to a constant as the order of the characteristic matrix increases. So the corresponding components, or polynomials with different degrees, have common zeroes.

    Next, we tested many characteristic matrix with various parameters like before and got their eigenvalues and eigenvectors. On one hand, by method (d), we evaluated the values of the first kind of polynomials for the corresponding degrees and orders. On the other hand, for every eigenvalue we applied forward recurrence to the polynomials of the first and second kinds with the same parameters and without considering whether the eigenvalues is less than 1 or not. Then we find that the method (d), in fact, is equivalent to the forward recurrence method, especially when the eigenvalue is close to 0 orm=0. In a typical case, we let ϖ=0.9,g=0.8,m=0, andN=251. There are four positive eigenvalues greater than 1. For every eigenvalue, to verify if the value vector of the polynomials of first kind from method (d) satisfies the C-D formula, we checked the logarithmic residual errors by Eq. (61) and Eq. (67).

As shown in Fig. 9, the x-axis, eigenvalue index, is the position number of the positive eigenvalues in ascending order, and the y-axis, error vector index, is the index of the residual errors computed by Eq. (61) and Eq. (67). In fact, though when the eigenvalue is greater than 1 the errors is not much small, as shown in the Fig. 9 (a), the results of method (d) confirm well to the C-D formula. Then, we checked the logarithmic absolute errors between the matrix of the polynomials from eigenvectors matrix by method (d) and the matrix of the polynomials by method (a). As shown in Fig. 9 (b), it is obvious that when the eigenvalue is less than 1, the errors are very small. When the eigenvalue is greater than 1, the forward recurrence is unstable, so it is no meaning to focus on it.

 figure: Fig. 9

Fig. 9 The logarithmic residual errors and absolute errors in method (d).

Download Full Size | PDF

So method (d) is also a feasible method to evaluate the polynomials of the first kind, especially when the other methods are not available.

4. Conclusion

The Case’s method is used to find the radiance distributions of the light transport in the infinite homogeneous tissue with m-dependent anisotropic scattering kernel, and the final analytical solution is expanded by the contributions of the discrete eigenfunctions and the continuous singular eigenfunctions. In our second accompanying paper, we will apply the Fourier transform to obtain another analytical solution of the light transport equation and prove the consistence between the two solutions in a mathematical manner.

Considering the vital importance of the two kinds of m-dependent Chandrasekhar orthogonal polynomials, we made great efforts on them. After the benchmark, the following guiding rules may be helpful:

  • 1) The determinant eigenvalues method is identical to the forward recurrence method for two kinds of Chandrasekhar polynomials.
  • 2) In the case m=0, the forward recurrence method is stable for the polynomials of the first kind only when |z|<1, and the backward recurrence method is suitable for them only when |z|>1.
  • 3) In the case m0, the forward recurrence method is only valid for the polynomials of the first kind when |z|<1 and not close to ±1, and it should be noticed that the bigger m, the worse the stability as zis approaching ±1, when |z|>1 or close to ±1, the backward recurrence should be applied to them.
  • 4) The forward recurrence is stable for the polynomials of the second kind for anyz and any m, and the backward recurrence is not suitable for the polynomials of the second kind.
  • 5) The linear system of equations method is also a feasible method for the larger degrees of the polynomials of the first kind, it is not suitable for the polynomials of the second kind.

Based on the conclusions, our future wok is to establish the measurable standards or formulas for the stabilities in m-dependent case. Since the Chandrasekhar polynomials can be evaluated effectively, the solutions of the light transport integro-differential equations in one/two/three-dimension full space, half space, and slab space etc. with general scattering kernel may be benchmarked to seek the more “exact” radiance or heat distributions in various tissue.

Funding

National Major Special Program of Scientific Instrument & Equipment Development of China (No. 2012YQ160203); Talent Plan 3551 of Wuhan City [No. WuXinGuan-(2010) 210], the Technology Innovation Fund of the National Science and Technology Ministry of China (No. 08C26224211088).

Acknowledgments

The main results of our two papers were obtained by extending derivation and benchmarks for the azimuth-independent Chandrasekhar polynomials by B D Ganapol to the m-dependent case and adapting these theories to biological optics.

Disclosures

The authors declare that there are no conflicts of interest related to this article.

References and links

1. S. Wright, M. Schweiger, and S. R. Arridge, “Reconstruction in optical tomography using the PN approximations,” Meas. Sci. Technol. 18(1), 247–260 (2007). [CrossRef]  

2. P. S. Mohan, T. Tarvainen, M. Schweiger, A. Pulkkinen, and S. R. Arridge, “Variable order spherical harmonic expansion scheme for the radiative transport equation using finite elements,” J. Comput. Phys. 230(19), 7364–7383 (2011). [CrossRef]  

3. A. D. Klose and E. W. Larsen, “Simplified spherical harmonics methods for modeling light transport in biological tissue,” in Biomedical Optics, Technical Digest (CD) (Optical Society of America, 2006), paper MH3. [CrossRef]  

4. A. Liemert and A. Kienle, “Analytical solutions of the simplified spherical harmonics equations,” Opt. Lett. 35(20), 3507–3509 (2010). [CrossRef]   [PubMed]  

5. H. Zheng and W. Han, “On simplified spherical harmonics equations for the radiative transfer equation,” J. Math. Chem. 49(8), 1785–1797 (2011). [CrossRef]  

6. A. Liemert and A. Kienle, “Comparison between radiative transfer theory and the simplified spherical harmonics approximation for a semi-infinite geometry,” Opt. Lett. 36(20), 4041–4043 (2011). [CrossRef]   [PubMed]  

7. M. Machida, G. Y. Panasyuk, J. C. Schotland, and V. A. Markel, “Diffusion approximation revisited,” J. Opt. Soc. Am. A 26(5), 1291–1300 (2009). [CrossRef]   [PubMed]  

8. A. Liemert and A. Kienle, “Light diffusion in a turbid cylinder. I. Homogeneous case,” Opt. Express 18(9), 9456–9473 (2010). [CrossRef]   [PubMed]  

9. A. Liemert and A. Kienle, “Light diffusion in a turbid cylinder. II. Layered case,” Opt. Express 18(9), 9266–9279 (2010). [CrossRef]   [PubMed]  

10. A. Kienle and A. Liemert, “Light diffusion in turbid media of different geometries in the steady-state, frequency, and time domains,” in Biomedical Optics and 3-D Imaging, OSA Technical Digest (CD) (Optical Society of America, 2010), paper BSuD41. [CrossRef]  

11. C. Devaux and C. E. Siewert, “The FN method for radiative transfer problems without azimuthal symmetry,” Z. Angew. Math. Phys. 31(5), 592–604 (1980). [CrossRef]  

12. M. Machida, “An FN method for the radiative transport equation in three dimensions,” J. Phys. A Math. Theor. 48(32), 325001 (2015). [CrossRef]  

13. K. M. Case, “Elementary solutions of the transport equation and their applications,” Ann. Phys. 9(1), 1–23 (1960). [CrossRef]  

14. N. J. Mccormick and N. Joseph, “One-speed neutron transport problems in plane geometry,” Univ. of Michigan Thesis, (1965).

15. N. J. McCormick and I. Kuščer, “Half-space neutron transport with linearly anisotropic scattering,” J. Math. Phys. 6(12), 1939–1945 (1965). [CrossRef]  

16. J. R. Mika, “Neutron transport with anisotropic scattering,” Nucl. Sci. Eng. 11(4), 415–427 (1961). [CrossRef]  

17. G. Van den Eynde, “Neutron transport with anisotropic scattering theory and applications,” Thesis, (2005).

18. N. V. Sultanov, “Elementary solution of the neutron transport equation with anisotropic scattering,” Sov. Atom Energ. 32(6), 539–546 (1972). [CrossRef]  

19. A. D. Kim, “Transport theory for light propagation in tissues,” in Biomedical Topical Meeting, OSA Technical Digest (Optical Society of America, 2004), paper SD4.

20. A. Liemert and A. Kienle, “Light transport in three-dimensional semi-infinite scattering media,” J. Opt. Soc. Am. A 29(7), 1475–1481 (2012). [CrossRef]   [PubMed]  

21. M. Machida, “Singular eigenfunctions for the three-dimensional radiative transport equation,” J. Opt. Soc. Am. A 31(1), 67–74 (2014). [CrossRef]   [PubMed]  

22. K. M. Case and R. D. Hazeltine, “Fourier transform methods in linear transport theory,” J. Math. Phys. 12(9), 1970–1980 (1971). [CrossRef]  

23. B. D. Ganapol, “A consistent theory of neutral particle transport in an infinite medium,” Transp. Theory Stat. Phys. 29(1–2), 43–68 (2000). [CrossRef]  

24. B. D. Ganapol, “Fourier transform transport solutions in spherical geometry,” Transp. Theory Stat. Phys. 32(5–7), 587–605 (2003). [CrossRef]  

25. B. D. Ganapol, “The Fourier transform solution for the Green’s function of monoenergetic neutron transport theory,” in 2014 Hawaii University International Conferences on Science, Technology, Engineering, Math & Education (2014).

26. B. D. Ganapol, “The infinite medium Green’s function of monoenergetic neutron transport theory via Fourier transform,” Nucl. Sci. Eng. 180(2), 224–246 (2015). [CrossRef]  

27. Inönü and Erdal, “Orthogonality of a set of polynomials encountered in neutron transport and radiative transfer theories,” J. Math. Phys. 11(2), 568–577 (1970). [CrossRef]  

28. K. M. Case, “Orthogonal polynomials from the viewpoint of scattering theory,” J. Math. Phys. 15(12), 2166–2174 (1974). [CrossRef]  

29. K. M. Case, “Scattering theory, orthogonal polynomials, and the transport equation,” J. Math. Phys. 15(7), 974–983 (1974). [CrossRef]  

30. R. D. M. Garcia and C. E. Siewert, “On discrete spectrum calculations in radiative transfer,” J. Quant. Spectrosc. Radiat. Transf. 42(5), 385–394 (1989). [CrossRef]  

31. R. D. M. Garcia and C. E. Siewert, “On computing the Chandrasekhar polynomials in high order and high degree,” J. Quant. Spectrosc. Radiat. Transf. 43(3), 201–205 (1990). [CrossRef]  

32. C. E. Siewert and N. J. Mccormick, “Some identities for Chandrasekhar polynomials,” J. Quant. Spectrosc. Radiat. Transf. 57(3), 399–404 (1997). [CrossRef]  

33. B. D. Ganapol, “Chandrasekhar polynomials and the solution to the transport equation in an infinite medium,” J. Comput. Theor. Trans. 43(1–7), 433–473 (2014). [CrossRef]  

34. M. Machida, “The Green’s function for the three-dimensional linear Boltzmann equation via Fourier transform,” J. Phys. A Math. Theor. 49(17), 175001 (2016). [CrossRef]  

35. A. J. Welch and M. J. C. V. Gemert, Optical-Thermal Response of Laser-Irradiated Tissue (Springer Netherlands, 2011).

36. Z. X. Wang and D. R. Guo, Special Functions (World Scientific, 1989).

37. N. I. Muskhelishvili, Singular Integral Equations: Boundary Problems of Function Theory and Their Application to Mathematical Physics (Dover Publications, 2011).

38. D. Zwillinger and V. Moll, Table of Integrals, Series, and Products (Eighth Edition) (Academic Press, 2015).

39. I. Kuščer and N. J. McCormick, “Some analytical results for radiative transfer in thick atmospheres,” J. Comput. Theor. Trans. 20(5–6), 351–381 (1974).

40. N. J. McCormick and I. Kuščer, “Bi-orthogonality relations for solving half-space transport problems,” J. Math. Phys. 7(11), 2036–2045 (1966). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (9)

Fig. 1
Fig. 1 The 2-norms of the characteristic matrix vary with anisotropy factor g , the curves in each subfigure denote the 2-norms when m = 0 , 5 , 10 , 15 , 20 , 25 , 30 from top to bottom
Fig. 2
Fig. 2 The distributions and variation tendency of the eigenvalues at some sampled positions, each curve in each subfigure corresponds to a certain sampled position
Fig. 3
Fig. 3 The result sequences and errors of the first and second kinds of Chandrasekhar polynomials as the degree (x-axis) increases when z = 0.48 by method (a) or (b).
Fig. 4
Fig. 4 The average errors of the first (left) and second (right) kinds of Chandrasekhar polynomials as z from 0 to 1 with step 0.2 and m from 0 to 30 with step 5 when N = 599
Fig. 5
Fig. 5 The curves of the first (left) and second (right) kinds of Chandrasekhar polynomials as z from −1 to 1. g ( l , m ) denotes g ^ l m ( z ) , and ρ ( l , m ) denotes ρ ^ l m ( z )
Fig. 6
Fig. 6 The average logarithmic absolute errors of the two sequences of the first (left) and second (right) kind of Chandrasekhar polynomials as z from 0 to 1 with step 0.2 and m from 0 to 30 with step 5 when N = 599 .
Fig. 7
Fig. 7 The average logarithmic residual errors of the first and second kind of Chandrasekhar polynomials as z from 0 to 1 with step 0.2 and m from 0 to 30 with step 5 when N = 599 .
Fig. 8
Fig. 8 The average logarithmic residual errors of the first and second kind of Chandrasekhar polynomials as z from 1.25 to 9.00 with step 0.25 and m from 0 to 30 with step 5 when N = 251 .
Fig. 9
Fig. 9 The logarithmic residual errors and absolute errors in method (d).

Equations (67)

Equations on this page are rendered with MathJax. Learn more.

μ ψ ( x , μ , φ ; μ 0 , φ 0 ) x + μ t ψ ( x , μ , φ ; μ 0 , φ 0 ) = μ s 0 2 π 1 + 1 p ( cos γ ) ψ ( x , μ , φ ; μ 0 , φ 0 ) d μ d φ + δ ( x ) δ ( μ μ 0 ) δ ( φ φ 0 )
p ( cos γ ) = 1 4 π l = 0 L ( 2 l + 1 ) f l m = l l ( l m ) ! ( l + m ) ! P l m ( μ ) P l m ( μ ) e i m ( φ φ )
ψ ( x , μ , φ ; μ 0 , φ 0 ) = 1 2 π m = ψ m ( x , μ ; μ 0 ) e i m φ
μ ψ m ( x , μ ; μ 0 ) x + μ t ψ m ( x , μ ; μ 0 ) = μ s 2 l = | m | L β l m P l m ( μ ) 1 + 1 P l m ( μ ) ψ m ( x , μ ; μ 0 ) d μ + δ ( x ) δ ( μ μ 0 ) e i m φ 0
ψ m ( x , μ ) exp ( μ t x / v ) ϕ m ( v , μ )
( 1 μ v ) ϕ m ( v , μ ) = ϖ 2 l = | m | L β l m P l m ( μ ) 1 + 1 P l m ( μ ) ϕ m ( v , μ ) d μ
g l m ( v ) ( 1 ) m 1 + 1 P l m ( μ ) ϕ m ( v , μ ) d μ
ϕ m ( v , μ ) = ϖ v 2 g m ( v , μ ) v μ
1 1 ϕ m ( v , μ ) ( 1 μ 2 ) | m | / 2 d μ = 1
g m ( v , μ ) = l = | m | L β l m g l m ( v ) P l m ( μ )
Λ m ( z ) 1 ϖ z 2 1 1 g m ( z , μ ) z μ ( 1 μ 2 ) | m | / 2 d μ
[ Λ m ( v ) ] ± = 1 ϖ v 2 P 1 1 g m ( v , μ ) v μ ( 1 μ 2 ) | m | / 2 d μ ± i π ϖ v 2 g m ( v , v ) ( 1 v 2 ) | m | / 2
ϕ m ( v , μ ) = ϖ v 2 P g m ( v , μ ) v μ + λ ( v ) ( 1 v 2 ) | m | / 2 δ ( v μ )
λ ( v ) = 1 ϖ v 2 P 1 1 g m ( v , μ ) v μ ( 1 μ 2 ) | m | / 2 d μ
[ Λ m ( v ) ] + + [ Λ m ( v ) ] = 2 λ ( v )
[ Λ m ( v ) ] + [ Λ m ( v ) ] = λ 2 ( v ) + [ π ϖ v 2 g m ( v , v ) ( 1 v 2 ) | m | / 2 ] 2
1 1 ϕ m ( v i m , μ ) ϕ m ( v j m , μ ) μ d μ = N 0 m ( v j m ) δ i j
1 1 ϕ m ( v , μ ) ϕ m ( v , μ ) μ d μ = N m ( v ) δ ( v v )
1 1 ϕ m ( v j , μ ) ϕ m ( v , μ ) μ d μ = 0
N 0 m ( ± v j m ) = ± ϖ v j m 2 [ ϖ ( v j m ) 2 l = | m | L β l m g l m ( v j m ) k = | m | L β k m g k m ( v j m ) γ l m δ m 0 ( v j m ) 2 1 l = | m | L β l m [ g l m ( v j m ) ] 2 ( 1 + δ m 0 ) v j m l = | m | L β l m g l m ( v j m ) k = 0 [ l 1 2 ] ( 2 l 4 k 1 ) g l 2 k 1 m ( v j m ) ]
1 1 P l m ( μ ) P k m ( μ ) z μ d μ = { 2 P l m ( z ) Q k m ( z ) l k 2 Q l m ( z ) P k m ( z ) l > k
d P l m ( x ) d x = 1 + δ m 0 2 k = 0 [ l 1 2 ] ( 2 l 4 k 1 ) P l 2 k 1 m ( x )
N 0 m ( ± v j m ) = ± 1 2 ϖ ( v j m ) 2 g m ( v j m , v j m ) [ ( v j m ) 2 1 ] | m | / 2 d Λ m ( z ) d z | z = v j m
N m ( v ) = v [ Λ m ( v ) ] + [ Λ m ( v ) ] ( 1 v 2 ) | m |
ψ m ( x , μ ) = j = 1 M a + m ϕ m ( v j m , μ ) exp ( μ t x / v j m ) + j = 1 M a m ϕ m ( v j m , μ ) exp ( μ t x / v j m ) + 1 1 A m ( v ) ϕ m ( v , μ ) exp ( μ t x / v ) d v
ψ m ( 0 + , μ ) ψ m ( 0 , μ ) = 1 μ δ ( x ) δ ( μ μ 0 ) e i m φ 0
lim | x | ψ m ( x , μ ) = 0
ψ m ( x , μ ; μ 0 ) = { j = 1 M a + m ϕ m ( v j m , μ ) exp ( μ t x / v j m ) + 0 1 A m ( v ) ϕ m ( v , μ ) exp ( μ t x / v ) d v ( x > 0 ) j = 1 M a m ϕ m ( v j m , μ ) exp ( μ t x / v j m ) 1 0 A m ( v ) ϕ m ( v , μ ) exp ( μ t x / v ) d v ( x < 0 )
a j ± m = 1 N 0 m ( ± v j m ) ϕ m ( ± v j m , μ 0 ) e i m φ 0
A ( v ) = 1 N m ( v ) ϕ m ( v , μ 0 ) e i m φ 0
ψ m ( x , μ ; μ 0 ) = e i m φ 0 { j = 1 M ϕ m ( v j m , μ 0 ) ϕ m ( v j m , μ ) N 0 m ( v j m ) exp ( μ t x / v j m ) + 0 1 ϕ m ( v , μ 0 ) ϕ m ( v , μ ) N m ( v ) exp ( μ t x / v ) d v ( x > 0 ) j = 1 M ϕ m ( v j m , μ 0 ) ϕ m ( v j m , μ ) N 0 m ( v j m ) exp ( μ t x / v j m ) 1 0 ϕ m ( v , μ 0 ) ϕ m ( v , μ ) N m ( v ) exp ( μ t x / v ) d v ( x < 0 )
g l m ( z ) = ( 1 ) m ( l m ) ! ( l + m ) ! g l m ( z )
g l m ( z ) = ( 1 ) l m g l m ( z )
z h l g l m ( z ) ( l m + 1 ) g l + 1 m ( z ) ( l + m ) g l 1 m ( z ) = 0
g ^ l m ( z ) = ( l m ) ! ( l + m ) ! g l m ( z )
z h l g ^ l m ( z ) ( l + 1 ) 2 m 2 g ^ l + 1 m ( z ) l 2 m 2 g ^ l 1 m ( z ) = 0
g ^ m m ( z ) = ( 2 m 1 ) ! ! ( 2 m ) ! = ( 2 m ) ! 2 m m !
z h l ρ l m ( z ) ( l m + 1 ) ρ l + 1 m ( z ) ( l + m ) ρ l 1 m ( z ) = ρ m + 1 m ( z ) δ l m
ρ ^ l m ( z ) = ( l m ) ! ( l + m ) ! ρ l m ( z )
z h l ρ ^ l m ( z ) ( l + 1 ) 2 m 2 ρ ^ l + 1 m ( z ) l 2 m 2 ρ ^ l 1 m ( z ) = 2 m + 1 ρ ^ m + 1 m ( z ) δ l m
g l 1 m ( z ) ρ l m ( z ) g l m ( z ) ρ l 1 m ( z ) = ( l + m 1 ) ! ( l m ) ! ρ m + 1 m ( z ) g m m ( z ) ( 2 m ) !
g ^ l 1 m ( z ) ρ ^ l m ( z ) g ^ l m ( z ) ρ ^ l 1 m ( z ) = ( 2 m + 1 ) ρ ^ m + 1 m ( z ) g ^ m m ( z ) l 2 m 2
g ^ l m ( z ) = ( 2 m ) ! g ^ m m ( z ) ( l m ) ! ( l + m ) ! | A l 1 m ( z ) |
ρ ^ l m ( z ) = ( 2 m + 1 ) ! ρ ^ m + 1 m ( z ) ( l m ) ! ( l + m ) ! | A l 1 m + 1 ( z ) |
A l m ( z ) = ( z h m ( m + 1 ) 2 m 2 ( m + 1 ) 2 m 2 z h m + 1 z h l 1 l 2 m 2 l 2 m 2 z h l )
| A l m ( z ) | = ( j = m l h j ) | B l m + z I |
B l m = ( 0 b m + 1 m b m + 1 m 0 b m + 2 m b m + 2 m 0 b l m b l m 0 )
g ^ l m ( z ) = g ^ m m ( z ) j = 0 l m 1 h m + j ( z + λ g , j m ) ( j + 1 ) ( j + 1 + 2 m )
ρ ^ l m ( z ) = ρ ^ m + 1 m ( z ) j = 1 l m 1 h m + j ( z + λ ρ , j m ) ( j + 1 ) ( j + 1 + 2 m )
d ^ j m ( z ) = z h j 1 d ^ j 1 m ( z ) ( j 1 ) 2 m 2 d ^ j 2 m ( z ) j 2 m 2
g ^ j m ( z ) = z h j + 1 g ^ j + 1 m ( z ) ( j + 2 ) 2 m 2 g ^ j + 2 m ( z ) ( j + 1 ) 2 m 2
g ^ j m ( z ) = γ j m g ^ m + N m ( z )
γ j m ( z ) = z h j + 1 γ j + 1 m ( z ) ( j + 2 ) 2 m 2 γ j + 2 m ( z ) ( j + 1 ) 2 m 2
g ^ m m ( z ) = γ m m ( z ) g ^ N m ( z )
g ^ j m ( z ) = γ j m γ m m ( z ) g ^ m m ( z )
[ B m + N m ( z ) z I ] G m + N m ( z ) = G ¯ m + N + 1 m ( z )
a j = h j g ^ j m ( z ) , j = m , m + 1 , , m + N + 1
[ B m + N m z I ] G m + N m ( z ) = 0
| B m + N m z I | = 0
g ^ j m ( z i ) = h m h j a j a m g ^ m m ( z i )
R 1 , j m ( z ) g ^ j 1 m ( z ) ρ ^ j m ( z ) g ^ j m ( z ) ρ ^ j 1 m ( z ) 2 m + 1 ρ ^ m + 1 m ( z ) g ^ m m ( z ) j 2 m 2
R 2 , j m ( z ) z h j ρ ^ j m ( z ) ( j + 1 ) 2 m 2 ρ ^ j + 1 m ( z ) j 2 m 2 ρ ^ j 1 m ( z )
R 3 , j m ( z ) z h j g ^ j m ( z ) ( j + 1 ) 2 m 2 g ^ j + 1 m ( z ) j 2 m 2 g ^ j 1 m ( z )
R 3 , j m ( z ) d ^ j , 1 m ( z ) d ^ j , 2 m ( z )
R 4 , j m ( z ) d ^ j , 1 m ( z ) d ^ j , 2 m ( z ) d ^ j , 1 m ( z )
D i m ( z ) log | R i , j m ( z ) |
D ¯ i m ( z ) log { R i , j m ( z ) [ R i , j m ( z ) ] T }
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.