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

Improvements in the computational efficiency and convergence of the Invariant Imbedding T-matrix method for spheroids and hexagonal prisms

Open Access Open Access

Abstract

The invariant-imbedding T-matrix (II-TM) method is a numerical method for accurately computing the single-scattering properties of dielectric particles. Because of the linearity of Maxwell’s equations, the incident electric field and the scattered electric field can be related through a transition matrix (T-Matrix). The II-TM method computes the T-matrix through a matrix recurrence formula which stems from an electromagnetic volume integral equation. The recurrence starts with an inscribed sphere within the particle and ends with the circumscribed sphere of the particle. For each iteration, a matrix known as the U-matrix is computed with the Gauss Legendre (GL) quadrature, and matrix inversion is subsequently performed to obtain the T-matrix corresponding to the portion of the particle enclosed by the spherical shell. We modify a commonly used scheme to avoid applying the quadrature scheme to discontinuities. Moreover, we apply a new scheme to generate nodes and weights in conjunction with the GL quadrature. This leads to a considerable improvement on convergence and computational efficiency in the cases of hexagonal prisms and spheroids. The basic shapes of ice crystals in the atmosphere are hexagonal columns and plates. The single-scattering properties of hexagonal ice prisms are important to atmospheric optics, radiative transfer, and remote sensing. We demonstrate that the present approach can significantly accelerate the convergence in simulating light scattering by hexagonal ice crystals.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

For atmospheric and oceanic radiative transfer simulations and remote sensing implementations, accurate computation of the single-scattering properties, namely the extinction cross section, the single-scattering albedo, and the phase matrix, of non-spherical particles is essential [1,2]. Because of the linearity of the equations governing the light-scattering process, the transformation from the incident field to the scattered field through their expansions in terms of vector spherical wave functions can be described by the Transition Matrix (T-Matrix) [3,4]. The single-scattering properties of the scattering particle can be derived from the T-matrix. The T-Matrix is an inherent property of the particle; in particular, it depends only on the particle structure, size and refractive index. The T-matrix of a homogeneous sphere reduces to a diagonal form with elements on the diagonal corresponding to the Lorenz-Mie series expansion coefficients. For nonspherical particles, the T-matrix method is computationally efficient because it is feasible to analytically average the optical properties over random orientations, as illustrated in [5,6]. Several efficient computational methods have been developed to compute the T-matrix of nonspherical and inhomogeneous particles [7,8]. The invariant-imbedding T-matrix (II-TM) method is a robust method developed to compute the single-scattering properties of non-spherical and inhomogeneous dielectric particles [9–11].

The II-TM method computes the T-Matrix of a scattering particle in a radial recurrence form starting with a certain initial inscribed sphere inside the particle and ending with the circumscribed sphere of the particle. At each recurrence level, Gauss-Legendre (GL) quadrature and matrix inversions are performed. As a result, three parameters determine the accuracy of the aforementioned numerical procedures: the radial discretizationΔr; the number of GL quadrature nodesNq, and the T-matrix expansion order N. In the numerical implementation of II-TM, a quadrature scheme is applied to define the geometry of the scattering particle. Any numerical errors associated with the quadrature scheme may be magnified in the matrix inversions and radial recurrence.

The focus of this study is on improving the computational efficiency and convergence of the GL quadrature. For particles with sharp edges, discontinuities occur in the quadrature method; we modify a commonly used scheme to avoid performing quadrature over such discontinuities. The example of a particle with sharp edges we consider here is inherently of considerable interest in atmospheric single-scattering studies, but the modification we make can be more generally applied. In addition, we replace the GL node generating scheme with the fastest algorithm to date [12]. These improvements substantially enhance the computational efficiency of the II-TM method for spheroids and hexagonal columns.

2. The T-matrix method

The incident and scattered electric fields can be expanded with the vector spherical wave functions (VSWFs) Mmn(1,3)and Nmn(1,3) in the form [13]

Einc(r)=E0n=0m=nnamnMmn(1)(kr)+bmnNmn(1)(kr),
Esca(r)=E0n=0m=nnpmnMmn(3)(kr)+qmnNmn(3)(kr),
where k is the wavenumber, ris the position vector. Superscript 1 indicates regular wave functions in which spherical Bessel functions of the first kind are used to meet the boundary condition at coordinate center (particle center), and superscript 3 indicates radiating wave functions in which spherical Hankel functions are used to represent the field radiating from the source (particle center) to infinity. The T-matrix relates the expansion coefficients:
(pmnqmn)=E0n'=0m'=n'n'(Tmnm'n'11Tmnm'n'12Tmnm'n'21Tmnm'n'22)(am'n'bm'n').
If index n is truncated at nmax, the T-matrix is a nmax(nmax+2)×nmax(nmax+2) square matrix. The two column vectors on the left and right sides of Eq. (3) are nmax(nmax+2) column vectors. Denote the T-matrix as T and the column vectors as p and a. The above equation can be written in a concise super-matrix form:
p=Ta.
After the T-matrix is obtained, it is straightforward to compute the single-scattering properties (see [14] for mathematical details).

3. Improving the convergence of Gaussian quadrature in II-TM

In the II-TM method, the T-matrix of a particle is obtained recursively, starting with the T-matrix T(r0)of an inscribed sphere of the non-spherical particle, successively calculating T-matrices for a series of concentric spheres, and ending at the circumscribed sphere of the particle (r = R). The T-matrix of the inscribed sphere T(r0)can be obtained with the Lorenz-Mie theory [9]. The T-matrix recurrence formula that relates the T-matrix T(r0)for sphere of radius rpto the T-matrix T(rp1) for the sphere of smaller radius rp1 is given in super-matrix notation [9] as:

T(rp)=Q11(rp)+[I+Q12(rp)][IT(rp1)Q22(rp)]1·T(rp)[I+Q21(rp)].
Qij(rp),i,j=1,2are all nmax(nmax+2)×nmax(nmax+2)square super-matrices given by
Q11(rp)=ikJT(rp)Q(rp)J(rp),
Q12(rp)=ikJT(rp)Q(rp)H(rp),
Q21(rp)=ikHT(rp)Q(rp)J(rp),
Q22(rp)=ikHT(rp)Q(rp)H(rp),
where J(rp) and H(rp) are nmax(nmax+2)×nmax(nmax+2) diagonal square super-matrices whose diagonal elements are defined byJ¯¯n(rp)and H¯¯n(rp) [9]:
J¯¯n(rp)=(jn(krp)0001krrrjn(krp)n(n+1)jn(krp)krp),
H¯¯n(rp)=(hn(1)(krp)0001krrrhn(1)(krp)n(n+1)hn(1)(krp)krp),
where jnis the spherical Bessel function of the first kind, hn(1)is the spherical Hankel function of the first kind. The super-matrix Q(rp)in Eqs. (6)-(9) is given by,
Q(rp)=Δr[IΔrU(rp)g(rp,rp)]1U(rp).
In Eq. (12), the elements of g(rp,rp)contain spherical Bessel and Hankel functions of the first kind, the elements of U(rp) contain zenith and azimuth integrals on the spherical layers that are used to define the particle geometry (see Eqs. (23)-(25) of [9]). Two inversions must be performed to obtain the T-matrix T(rp) in each iteration (Eqs. (5) and (12)). Equation (5) was first derived through discretizing integral equations of T-matrix [13]. Figure 1 shows the geometries. The origin of the coordinate system is at the particle center.

 figure: Fig. 1

Fig. 1 Schematic of the radial recurrence. Particle in this schematic configuration is a particle with arbitrary geometry marked in green. Coordinate origin is on the geometric center of the particle. Two spherical shells are indicated by dashed circles. Arrows indicate the two radii rp andrp+1.

Download Full Size | PDF

The matrix elements of U(rp) for each iteration are surface integrals in zenith and azimuth directions on the spherical shell. The nonzero elements in the U-matrix can be written as [10]

Umnm'n'ij=Amnm'n'0πdθsinθFmm'(r,θ)Kmnm'n'ij(θ),i,j=1,2,
Umnm'n'33=A'mnm'n'0πdθsinθF'mm'(r,θ)d0mn(θ)d0m'n'(θ),
where the integrations in the azimuth direction Fmm'(r,θ) and F'mm'(r,θ) are
Fmm'(r,θ)=02πdφei(mm')φ[ε(r,θ,φ)1],
F'mm'(r,θ)=02πdφei(mm')φ[ε(r,θ,φ)1]ε(r,θ,φ),
and the constants Amnm'n' and A'mnm'n' are
Amnm'n'=(kr)2(1)m+m'[2n+14πn(n+1)]1/2[2n'+14πn'(n'+1)]1/2,
A'mnm'n'=nn'(n+1)(n'+1)Amnm'n'.
The terms Kmnm'n'ij(θ) appearing in the integrand of Eq. (13) are defined by
Kmnm'n'ij(θ)=(πmn(θ)πm'n'(θ)+τmn(θ)τm'n'(θ)i[πmn(θ)τm'n'(θ)+τmn(θ)πm'n'(θ)]i[πmn(θ)τm'n'(θ)+τmn(θ)πm'n'(θ)]πmn(θ)πm'n'(θ)+τmn(θ)τm'n'(θ)),
where πmn and τmn are obtained from the Wigner-d function d0mn(θ)
πmn(θ)=msinθd0mn(θ),
τmn(θ)=ddθd0mn(θ).
The U-matrix elements are different for different particles. Below we outline the formulations for the spheroid and hexagonal column. These shapes are used as basic particle structures in modelling atmospheric ice crystals and other aerosols. A spheroid has two important kinds of symmetries that allow computational simplifications: rotational symmetry with respect to the z-axis ε(r,θ,φ)=ε(r,θ) and mirror symmetry with respect to the xy-planeε(r,θ)=ε(r,πθ). Also,
ε(r,θ)1={0,θ2<θ,ε1,0<θ<θ2,
where θ2 is the angle at which the spherical shell intersects the spheroid (Fig. 2).

 figure: Fig. 2

Fig. 2 Procedures in the sequence of expanding spherical shells in the recurrence. In the case of the spheroid, θ1=0. θ2 is the angle where the spherical shell intersects the spheroid.

Download Full Size | PDF

The following properties of the Wigner-d functions allow us to simplify the formulations for spheroid [15]

[πmn(θ)πmn'(θ)+τmn(θ)τmn'(θ)]sinθ=ddθ(τmn'(θ)d0mn(θ)sinθ)+n'(n'+1)sinθd0mn(θ)d0mn'(θ)=ddθ(τmn(θ)d0mn'(θ)sinθ)+n(n+1)sinθd0mn(θ)d0mn'(θ),
[πmn(θ)τmn'(θ)+τmn(θ)πmn'(θ)]sinθ=mddθ(d0mn(θ)d0mn'(θ)).
Substituting Eqs. (23) and (24) into Eqs. (13) and (14), followed by integration by parts, we obtain the simplified formulations:
Umnm'n'11=Umnm'n'22=Amnm'n'2πδmm'[ε1][cnn'τnn'(θ2)d0mn(θ2)sinθ2+n'(n'+1)0θ2dθsinθd0mn(θ)d0mn'(θ)],cnn'=1+(1)n+n',
Umnm'n'33=A'mnm'n'2πδmm'ε1εcnn'0θ2dθsinθd0mn(θ)d0mn'(θ),
Umnm'n'12=Umnm'n'21=Amnm'n'2πδmm'[ε1]c'nn'(im)[d0mn(θ2)d0mn'(θ2)d0mn(0)d0mn'(0)],c'nn'=1+(1)n+n'+1,
The original integrals in the zenith direction in Eqs. (13) and (14) are reduced to 0θ2dθsinθd0mn(θ)d0mn'(θ) in Eqs. (25) and (26). The hexagonal column also has a rotational symmetry about the z-axis, but it is no longer continuous and is instead discrete. As a consequence, the azimuth integrals Fmm'(r,θ) and F'mm'(r,θ) cannot be decoupled from the zenith integrals. The values of the azimuth integrals are dependent on θ and r because the spherical shell intersects a hexagonal column in different ways as θ and r change. Figure 3 illustrates the three different situations as the recurrence progresses in the case of a long column.

 figure: Fig. 3

Fig. 3 Procedures in the sequence of expanding spherical shells in the recurrence for a hexagonal column. θ2is the angle where the spherical shell intersects the column.

Download Full Size | PDF

Using the mirror symmetry ofε(r,θ)=ε(r,πθ), we can simplify Eqs. (13) and (14):

Umnm'n'11=Umnm'n'22=Amnm'n'cnn'θ1θ2dθsinθFmm'(r,θ)Kmnm'n'ii(θ),cnn'=1+(1)n+n',
Umnm'n'33=A'mnm'n'cnn'θ1θ2dθsinθF'mm'(r,θ)d0mn(θ)d0m'n'(θ),
Umnm'n'12=Umnm'n'21=Amnm'n'c'nn'θ1θ2dθsinθFmm'(r,θ)Kmnm'n'ij(θ),c'nn'=1+(1)n+n'+1.
Fmm'(r,θ)and F'mm'(r,θ) in this case are piecewise smooth functions inθ. Figure 3 shows Fmm'(r,θ) of some index mm' at a certainrp. A discontinuity (in the first derivative) occurs when the spherical shell intersects the hexagonal surface. As a result, the U-matrix integrand Fmm'(r,θ)Kmnm'n'ii(θ) is no longer smooth in the interval[θ1,θ2]. The location of the discontinuity θ=θjumpis given in Eq. (16) of [10], whererpsinθjump=3a2.

 figure: Fig. 4

Fig. 4 An example Fmm'(r,θ) as function of θ for hexagonal column.

Download Full Size | PDF

The surface integrations in the zenith direction (Eqs. (25)~(30)) are naturally computed using Gauss-Legendre (GL) quadrature in the II-TM numerical implementations:

0π(dθsinθ)(Fmm'(r,θ)Kmnm'n'ij(θ))l=1NqwlF(r,θl)mm'Kmnm'n'ij(θl).
In addition to the T-matrix truncation order N and the radial discretizationΔr, the number of quadrature nodes Nq is an important parameter controlling the accuracy of the II-TM method. For a standard Gaussian Legendre quadrature such as Eq. (31), the optimal nodes for quadrature of order Nq are given by the roots of the Legendre polynomialsPNq(x)=PNq(cosθ). The GL quadrature is optimal if the integrand (Fmm'(r,θ)Kmnm'n'ij(θ)) can be well approximated by polynomials of order 2Nq1 or less. Generally, this means that the various orders of derivative of the integrand are smooth. Discontinuities or singularities in the integrand or its derivatives will reduce the smoothness of the integrand. Higher order polynomials are required to approximate the fine scale oscillations of the integrand thus Nq can go to very large value. We use Gaussian quadrature on the following two integrals to illustrate our point that convergence rate of the Gaussian quadrature depends on behaviour of the integrand.

01f1(x)dxn=1Nqwnf1(xn),f1(x)={1,0x0.5,2(1x),0.5<x1.
01f2(x)dxn=1Nqwnf2(xn),f2(x)={1,0x0.5,0.5[1+cos2π(x0.5)],0.5<x1.
 figure: Fig. 5

Fig. 5 f1(x) andf2(x).

Download Full Size | PDF

Function f2(x) is smooth inside the [0,1] interval while f1(x) possesses a discontinuity in its first-order derivative at x=0.5. The behaviour of f1(x) is similar to that of Fmm'(θ)in Fig. 4; they each have a discontinuity in the first-order derivative within their integration range. The two integrals of f1(x) and f2(x) have the same exact value (0.75), and we can demonstrate the convergence rate of Gaussian quadrature in terms of the rate of decrease of relative error in the numerical approximations. Table 1 shows the magnitude of the relative errors with respect to increasing the number of nodes Nq for the two integrands. Apparently f2(x) shows a faster rate of convergence.

Tables Icon

Table 1. Convergence rate of Gaussian quadrature for 2 different integrands.

Back to the II-TM method, the spheroid has the simplest integrand (in Eqs. (25) and (26)) among all other shapes:

0θ2dθsinθd0mn(θ)d0mn'(θ),
where the integrand is the product of Wigner-d functionsd0mn(θ)d0mn'(θ). We expect rapid convergence of the quadrature with respect toNq, because the integrand d0mn(θ)d0mn'(θ)can be relatively well approximated by polynomials. Figure 6 shows the convergence rate of the outputs (Qext, asymmetry factor (g factor), P11(180)) with respect to Nq for these different size parameters. Qextis the extinction efficiency, P11(180)is the phase function in the exact backscattering direction and g factor is the asymmetry factor. Nq=1000is chosen as the reference value and outputs of Nq=200,400,600,800 are compared against it to demonstrate the convergence rate. T-matrix truncation order N and radial discretization Δr are fixed. The T-matrix truncation order N for a particle with size parameter x is determined according to the formula
N=x+4.05x1/3+5,
in effect adding three more terms to the familiar choice used for spherical particles [16]
NMIE=x+4.05x1/3+2,
No explicit formula has been published for calculations of the T-matrix for non-spherical particles, and we simply have made a choice that appears to increase the reliability of the results without greatly increasing computational cost. The increment in radius Δr is determined according to the formula,
ΔrRr0=0.1,
r0is the radius of the homogeneous sphere inscribed inside the non-spherical particle, and is the starting radius for the radial recurrence. R is the radius of the circumscribed sphere. The value 0.1 has been found to be sufficient for particles without small scale features such as surface roughness; in the present cases where the particle shapes are regular spheroids and hexagonal columns, 0.1 should be sufficient.

The spheroid shape is indicated in the figures. x indicates the particle size parameter defined with respect to the radius of the circumscribed sphere of the particle.

 figure: Fig. 6

Fig. 6 Convergence rate of the outputs (Qext, asymmetry factor (g factor), P11(180)) with respect to number of quadrature nodesNq. Six panels correspond to 6 different spheroidal shapes indicated with size parameter x and aspect ratio a/b.

Download Full Size | PDF

The convergence rates of different sizes show mixed behaviors. Generally, we can conclude that the convergence rate for spheroid is fast. In most cases, the outputs agree to more than 12 significant digits. The obstacle for spheroid is the intense oscillation of high-order Wigner-d functions for larger spheroids. Once Nq gets large enough (>200), the sampling rate is sufficiently large and we have rapid convergence. The integrals in the hexagonal column formulation (in Eqs. (28)~(30)) take the following form:

θ1θ2dθsinθFmm'(r,θ)Kmnm'n'ij(θ).
It was pointed out in Fig. 4 that the function Fmm'(r,θ) in the above integral has a discontinuity in its first derivative when the spherical shell crosses the hexagonal column. We expect this discontinuity to undermine the convergence of the quadrature if we do the quadrature in the interval [θ1,θ2] which contains the discontinuity. Figure 6 shows the convergence rate of the outputs (the extinction efficiencyQext, the asymmetry factor (the g factor), P11(180o)with respect toNq. Nq=1000 is again used as the reference value. The hexagonal column shape is indicated in the figure and x is the size parameter.

 figure: Fig. 7

Fig. 7 Convergence rate of the outputs (Qext, asymmetry factor (g factor), P11(180)) with respect to number of quadrature nodesNq. Six panels correspond to 6 different hexagonal column shapes indicated with size parameter x and aspect ratio a/b.

Download Full Size | PDF

Compared to spheroids (Fig. 6), the convergence rate is much slower. The discontinuity in Fmm'(r,θ) reduces the convergence rate of the quadrature. Note that there is only one discontinuity θjump in [θ1,θ2]. A similar situation was encountered in the EBCM method for hexagonal columns [17]. To address the problem of discontinuities, a natural solution is to split the interval and perform quadrature on the two intervals where the integrand is smooth (Eq. (41).

θ1θ2dθsinθFmm'(r,θ)Kmnm'n'ij(θ)=θ1θjumpdθsinθFmm'(r,θ)Kmnm'n'ij(θ)+θjumpθ2dθsinθFmm'(r,θ)Kmnm'n'ij(θ).
After making this modification, the results are shown in Fig. 8. Compared to Fig. 6, the convergence rate is accelerated. At this rate, we can expect to reach convergence to 16 digits at about Nq=1000. This justifies the use Nq1000 in practical applications.

 figure: Fig. 8

Fig. 8 Convergence rate of the outputs (Qext, asymmetry factor (g factor), P11(180)) with respect to number of quadrature nodes Nq. Six panels correspond to 6 different hexagonal column shapes indicated with size parameter x and aspect ratio a/b. Unlike Fig. 7, the integration range is split into 2 smooth intervals so the convergence rate is accelerated (see the text).

Download Full Size | PDF

From Figs. 6 and 8, we can conclude that, with respect to the number of quadrature points Nq, to reach convergence for common particle sizes and aspect ratios, Nq1000 is sufficient for hexagonal columns and Nq600 is sufficient for spheroids.

4. Accelerating the node and weight generating scheme for Gaussian quadrature in II-TM

As stated above, the optimal nodes for Gaussian-Legendre quadrature of order L are the roots of Legendre PolynomialsPL(x). There are no closed formulas for location of the roots so they have to be generated numerically along with the corresponding weights. Different algorithms exist for generating the nodes and weights of Gauss-Legendre quadrature [18,19].

The current algorithm in the II-TM is based on the Newton’s method (Newton-Raphson) to find the roots of PL(x). Denote the L node-weight pairs {xL,k,wL,k},k[0,L] and they satisfy PL(xL,k)=0. The weights are given by wL,k=2(1xL,k2)[(n+1)PL+1(xL,K)]2. For large L, some initial value xL,k,0 for xL,k is given and then Newton-Raphson iterations are performed to obtain a true value for xL,k up to some precision. A key in the design of fast algorithms is proper choice of asymptotic expansions (large L) of the Legendre polynomials. Considerable research effort has gone into the search for suitable asymptotic expansions [12,19–21].

The state-of-the-art for Gaussian-Legendre quadrature is the Bogaert’s algorithm [12,22]. In their algorithm, the Newton-Raphson iteration is avoided by providing asymptotic expansions directly for xL,k and wL,k. This algorithm is much faster than previous methods. In the following discussions, a commonly used algorithm is referred to as “Newton” and the new algorithm is referred to as “Bogaert”. We first illustrate the difference in the performance of the algorithms in the simple case of the model function f1(x) of the previous section (Fig. 5 and Table 1). Table 2 shows the CPU time required to compute the numerical quadrature using the two algorithms. The Bogaert algorithm is clearly much faster. Obviously, the Bogaert algorithm is much faster.

Tables Icon

Table 2. CPU time required to run the same code section with Newton and Bogaert algorithms.

We now turn to implementation of the Bogaert method in conjunction with the II-TM. For L>100, the Bogaert asymptotic expansions are used; for L<100, tabulated values for nodes and weights are used. Table 3 shows the acceleration for hexagonal columns and spheroids. II-TM is implemented with the Massage Passing Interface (MPI) parallelization and 100 CPUs were used for the experiments. Significant acceleration is achieved for hexagonal columns of all sizes and aspect ratios. Moderate acceleration is achieved for spheroid in terms of CPU time. However, in two cases, the Bogaert algorithm slightly slows down the method in terms of wall clock time. As mentioned above, the spheroid and hexagon have different symmetries, and correspondingly the code structures of the II-TM implementations are different. As a result, the respective MPI parallelizations are different. The result is that the corresponding acceleration rates associated with the Bogaert algorithm are different. The optimization of the Bogaert algorithm in the II-TM MPI implementations for a wide range of particle geometries is subject of future work.

Tables Icon

Table 3. Acceleration for hexagonal column and spheroids of various sizes and aspect ratios.

With the accelerated code, we compute the single scattering-properties of large hexagonal columns with size parameter (defined with the circumscribed sphere radius) x = 225 and aspect ratio = 1. Two cases are considered. One has the index of refractionm=1.308+i1.43×109which corresponds to ice refractive index at 650 nm wavelength. Another one has the index of refraction m=1.276+i0.413which corresponds to ice refractive index at 12 μmwavelength. Random orientation is assumed. Figures 9 and 10 show comparison between the II-TM method and a physical geometric optics method (PGOM) [23]. Very good agreement is achieved between the two methods.

 figure: Fig. 9

Fig. 9 Phase matrix elements for hexagonal column with size parameter x = 225 (defined with the circumscribed sphere radius). Comparison is between PGOM (blue) and II-TM (red dashed). Particle refractive index ism=1.308+i1.43×109.

Download Full Size | PDF

 figure: Fig. 10

Fig. 10 Phase matrix elements for hexagonal column with size parameter x = 225 (defined with the circumscribed sphere radius). Comparison is between PGOM (blue) and II-TM (red dashed). Particle refractive index ism=1.276+i0.413.

Download Full Size | PDF

5. Summary

We studied the performance of the angular quadrature in the II-TM method for spheroid and hexagonal column. The precision of the angular quadrature is of fundamental importance to the II-TM method and we want to achieve convergence by choosing an appropriate number of quadrature nodes Nq. In the case of a hexagonal column, the previous formulation [9] is modified to avoid a discontinuity in the integration domain, and the convergence rate is largely accelerated. Based on convergence test results from spheroids and hexagonal columns with various sizes, we conclude that Nq600 for spheroid and Nq1000 for hexagonal column is sufficiently large for the improved numerical implementation. A new node and weight generating algorithm (Bogaert algorithm) is implemented in the II-TM method, and considerable acceleration in computation time is achieved for hexagonal columns and spheroids. With the improved II-TM method, we compute the single-scattering properties of two large hexagonal columns with size parameter x = 225 and compare the results against the physical geometric optics method [23]. Good agreement is found between the two methods.

Funding

National Science Foundation (AGS-1826936); Texas A&M David Bullock Harris Chair in Geosciences endowment funds (02-512231-10000).

Acknowledgment

We gratefully acknowledge the computational support of the Texas A&M High Performance Computing Center.

References

1. C. F. Bohren and D. R. Huffman, Absorption and Scattering of Light by Small Particles (John Wiley & Sons, 1983).

2. H. C. van de Hulst, Light Scattering by Small Particles. (Dover, 1981).

3. P. C. Waterman, “Matrix formulation of electromagnetic scattering,” Proc. IEEE 53(8), 805–812 (1965). [CrossRef]  

4. P. C. Waterman, “The T-matrix revisited,” J. Opt. Soc. Am. A 24(8), 2257–2267 (2007). [CrossRef]   [PubMed]  

5. N. G. Khlebtsov, “Orientational averaging of light-scattering observables in the J-matrix approach,” Appl. Opt. 31(25), 5359–5365 (1992). [CrossRef]   [PubMed]  

6. M. I. Mishchenko, “Light scattering by randomly oriented axially symmetric particles,” J. Opt. Soc. Am. A 8(6), 871–882 (1991). [CrossRef]  

7. T. Wriedt, “Review of the null-field method with discrete sources,” J. Quant. Spectrosc. Radiat. Transf. 106(1-3), 535–545 (2007). [CrossRef]  

8. M. I. Mishchenko and L. D. Travis, “Capabilities and limitations of a current FORTRAN implementation of the T-matrix method for randomly oriented, rotationally symmetric scatterers,” J. Quant. Spectrosc. Radiat. Transf. 60(3), 309–324 (1998). [CrossRef]  

9. L. Bi, P. Yang, G. W. Kattawar, and M. I. Mishchenko, “Efficient implementation of the invariant imbedding T-matrix method and the separation of variables method applied to large nonspherical inhomogeneous particles,” J. Quant. Spectrosc. Radiat. Transf. 116, 169–183 (2013). [CrossRef]  

10. L. Bi and P. Yang, “Accurate simulation of the optical properties of atmospheric ice crystals with the invariant imbedding T-matrix method,” J. Quant. Spectrosc. Radiat. Transf. 138, 17–35 (2014). [CrossRef]  

11. T. Wriedt and Y. Eremin, The Generalized Multipole Technique for Light Scattering. (Springer, 2018).

12. I. Bogaert, “Iteration-Free Computation of Gauss–Legendre Quadrature Nodes and Weights,” SIAM J. Sci. Comput. 36(3), A1008–A1026 (2014). [CrossRef]  

13. B. R. Johnson, “Invariant imbedding T matrix approach to electromagnetic scattering,” Appl. Opt. 27(23), 4861–4873 (1988). [CrossRef]   [PubMed]  

14. M. Mischenko, L. Travis, and A. Lacis, Scattering, absorption, and emission of light by small particles, (Cambridge University Press,1970).

15. W. R. C. Somerville, B. Auguié, and E. C. Le Ru, “Simplified expressions of the T-matrix integrals for electromagnetic scattering,” Opt. Lett. 36(17), 3482–3484 (2011). [CrossRef]   [PubMed]  

16. W. J. Wiscombe, “Improved Mie scattering algorithms,” Appl. Opt. 19(9), 1505–1509 (1980). [CrossRef]   [PubMed]  

17. F. M. Kahnert, J. J. Stamnes, and K. Stamnes, “Application of the extended boundary condition method to particles with sharp edges: a comparison of two surface integration approaches,” Appl. Opt. 40(18), 3101–3109 (2001). [CrossRef]   [PubMed]  

18. A. Glaser, X. Liu, and V. Rokhlin, “A Fast Algorithm for the Calculation of the Roots of Special Functions,” SIAM J. Sci. Comput. 29(4), 1420–1438 (2007). [CrossRef]  

19. G. H. Golub and J. H. Welsch, “Calculation of Gauss Quadrature Rules,” Math. Comput. 23(106), 221–230 (1969). [CrossRef]  

20. N. Hale and A. Townsend, “Fast and Accurate Computation of Gauss–Legendre and Gauss–Jacobi Quadrature Nodes and Weights,” SIAM J. Sci. Comput. 35(2), A652–A674 (2013). [CrossRef]  

21. A. Townsend, T. Trogdon, and S. Olver, “Fast computation of Gauss quadrature nodes and weights on the whole real line,” IMA J. Numer. Anal. 36(1), 337–358 (2014).

22. A. Townsend, “The race to compute high-order Gauss-Legendre quadature,” SIAM J. Sci. Comput. (2015).

23. B. Sun, P. Yang, G. W. Kattawar, and X. Zhang, “Physical-geometric optics method for large size faceted particles,” Opt. Express 25(20), 24044–24060 (2017). [CrossRef]   [PubMed]  

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 (10)

Fig. 1
Fig. 1 Schematic of the radial recurrence. Particle in this schematic configuration is a particle with arbitrary geometry marked in green. Coordinate origin is on the geometric center of the particle. Two spherical shells are indicated by dashed circles. Arrows indicate the two radii r p and r p+1 .
Fig. 2
Fig. 2 Procedures in the sequence of expanding spherical shells in the recurrence. In the case of the spheroid, θ 1 =0. θ 2 is the angle where the spherical shell intersects the spheroid.
Fig. 3
Fig. 3 Procedures in the sequence of expanding spherical shells in the recurrence for a hexagonal column. θ 2 is the angle where the spherical shell intersects the column.
Fig. 4
Fig. 4 An example F mm' (r,θ) as function of θ for hexagonal column.
Fig. 5
Fig. 5 f 1 (x) and f 2 (x).
Fig. 6
Fig. 6 Convergence rate of the outputs ( Q ext , asymmetry factor (g factor), P 11 ( 180 )) with respect to number of quadrature nodes N q . Six panels correspond to 6 different spheroidal shapes indicated with size parameter x and aspect ratio a/b.
Fig. 7
Fig. 7 Convergence rate of the outputs ( Q ext , asymmetry factor (g factor), P 11 ( 180 )) with respect to number of quadrature nodes N q . Six panels correspond to 6 different hexagonal column shapes indicated with size parameter x and aspect ratio a/b.
Fig. 8
Fig. 8 Convergence rate of the outputs ( Q ext , asymmetry factor (g factor), P 11 ( 180 )) with respect to number of quadrature nodes N q . Six panels correspond to 6 different hexagonal column shapes indicated with size parameter x and aspect ratio a/b. Unlike Fig. 7, the integration range is split into 2 smooth intervals so the convergence rate is accelerated (see the text).
Fig. 9
Fig. 9 Phase matrix elements for hexagonal column with size parameter x = 225 (defined with the circumscribed sphere radius). Comparison is between PGOM (blue) and II-TM (red dashed). Particle refractive index is m=1.308+i1.43× 10 9 .
Fig. 10
Fig. 10 Phase matrix elements for hexagonal column with size parameter x = 225 (defined with the circumscribed sphere radius). Comparison is between PGOM (blue) and II-TM (red dashed). Particle refractive index is m=1.276+i0.413.

Tables (3)

Tables Icon

Table 1 Convergence rate of Gaussian quadrature for 2 different integrands.

Tables Icon

Table 2 CPU time required to run the same code section with Newton and Bogaert algorithms.

Tables Icon

Table 3 Acceleration for hexagonal column and spheroids of various sizes and aspect ratios.

Equations (39)

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

E inc ( r )= E 0 n=0 m=n n a mn M mn (1) (k r )+ b mn N mn (1) (k r ),
E sca ( r )= E 0 n=0 m=n n p mn M mn (3) (k r )+ q mn N mn (3) (k r ),
( p mn q mn )= E 0 n ' =0 m ' = n ' n ' ( T mn m ' n ' 11 T mn m ' n ' 12 T mn m ' n ' 21 T mn m ' n ' 22 ) ( a m ' n ' b m ' n ' ).
p=Ta.
T( r p )= Q 11 ( r p )+[ I+ Q 12 ( r p ) ] [ IT( r p1 ) Q 22 ( r p ) ] 1 · T( r p )[ I+ Q 21 ( r p ) ].
Q 11 ( r p )=ik J T ( r p )Q( r p )J( r p ),
Q 12 ( r p )=ik J T ( r p )Q( r p )H( r p ),
Q 21 ( r p )=ik H T ( r p )Q( r p )J( r p ),
Q 22 ( r p )=ik H T ( r p )Q( r p )H( r p ),
J ¯ ¯ n ( r p )=( j n (k r p ) 0 0 0 1 kr r r j n (k r p ) n(n+1) j n (k r p ) k r p ),
H ¯ ¯ n ( r p )=( h n (1) (k r p ) 0 0 0 1 kr r r h n (1) (k r p ) n(n+1) h n (1) (k r p ) k r p ),
Q( r p )=Δr [ IΔrU( r p )g( r p , r p ) ] 1 U( r p ).
U mn m ' n ' ij = A mn m ' n ' 0 π dθsinθ F m m ' (r,θ) K mn m ' n ' ij (θ),i,j=1,2,
U mn m ' n ' 33 = A ' mn m ' n ' 0 π dθsinθ F ' m m ' (r,θ) d 0m n (θ) d 0 m ' n ' (θ),
F m m ' (r,θ)= 0 2π dφ e i(m m ' )φ [ ε(r,θ,φ)1 ] ,
F ' m m ' (r,θ)= 0 2π dφ e i(m m ' )φ [ ε(r,θ,φ)1 ] ε(r,θ,φ) ,
A mn m ' n ' = (kr) 2 (1) m+ m ' [ 2n+1 4πn(n+1) ] 1/2 [ 2 n ' +1 4π n ' ( n ' +1) ] 1/2 ,
A ' mn m ' n ' = n n ' (n+1)( n ' +1) A mn m ' n ' .
K mn m ' n ' ij (θ)= ( π mn (θ) π m ' n ' (θ)+ τ mn (θ) τ m ' n ' (θ) i[ π mn (θ) τ m ' n ' (θ)+ τ mn (θ) π m ' n ' (θ) ] i[ π mn (θ) τ m ' n ' (θ)+ τ mn (θ) π m ' n ' (θ) ] π mn (θ) π m ' n ' (θ)+ τ mn (θ) τ m ' n ' (θ) ),
π mn (θ)= m sinθ d 0m n (θ),
τ mn (θ)= d dθ d 0m n (θ).
ε(r,θ)1={ 0, θ 2 <θ, ε1,0<θ< θ 2 ,
[ π mn (θ) π m n ' (θ)+ τ mn (θ) τ m n ' (θ) ]sinθ = d dθ ( τ mn' (θ) d 0m n (θ)sinθ )+n'(n'+1)sinθ d 0m n (θ) d 0m n' (θ) = d dθ ( τ mn (θ) d 0m n' (θ)sinθ )+n(n+1)sinθ d 0m n (θ) d 0m n' (θ),
[ π mn (θ) τ m n ' (θ)+ τ mn (θ) π m n ' (θ) ]sinθ=m d dθ ( d 0m n (θ) d 0m n' (θ) ).
U mnm'n' 11 = U mnm'n' 22 = A mnm'n' 2π δ mm' [ε1] [ c nn' τ nn' ( θ 2 ) d 0m n ( θ 2 )sin θ 2 +n'(n'+1) 0 θ 2 dθsinθ d 0m n (θ) d 0m n' (θ) ], c nn' =1+ (1) n+n' ,
U mnm'n' 33 =A ' mnm'n' 2π δ mm' ε1 ε c nn' 0 θ 2 dθsinθ d 0m n (θ) d 0m n' (θ),
U mnm'n' 12 = U mnm'n' 21 = A mnm'n' 2π δ mm' [ε1]c ' nn' (im)[ d 0m n ( θ 2 ) d 0m n' ( θ 2 ) d 0m n (0) d 0m n' (0) ], c ' nn' =1+ (1) n+n'+1 ,
U mnm'n' 11 = U mnm'n' 22 = A mnm'n' c nn' θ 1 θ 2 dθsinθ F mm' (r,θ) K mnm'n' ii (θ), c nn' =1+ (1) n+n' ,
U mnm'n' 33 =A ' mnm'n' c nn' θ 1 θ 2 dθsinθF ' mm' (r,θ) d 0m n (θ) d 0m' n' (θ),
U mnm'n' 12 = U mnm'n' 21 = A mnm'n' c ' nn' θ 1 θ 2 dθsinθ F mm' (r,θ) K mnm'n' ij (θ), c ' nn' =1+ (1) n+n'+1 .
0 π (dθsinθ)( F mm' (r,θ) K mnm'n' ij (θ) ) l=1 N q w l F (r, θ l ) mm' K mnm'n' ij ( θ l ) .
0 1 f 1 (x)dx n=1 N q w n f 1 ( x n ) , f 1 (x)={ 1,0x0.5, 2(1x),0.5<x1.
0 1 f 2 (x)dx n=1 N q w n f 2 ( x n ) , f 2 (x)={ 1,0x0.5, 0.5[1+cos2π(x0.5)],0.5<x1.
0 θ 2 dθsinθ d 0m n (θ) d 0m n' (θ),
N=x+4.05 x 1/3 +5,
N MIE =x+4.05 x 1/3 +2,
Δr R r 0 =0.1,
θ 1 θ 2 dθsinθ F mm' (r,θ) K mnm'n' ij (θ).
θ 1 θ 2 dθsinθ F mm' (r,θ) K mnm'n' ij (θ) = θ 1 θ jump dθsinθ F mm' (r,θ) K mnm'n' ij (θ) + θ jump θ 2 dθsinθ F mm' (r,θ) K mnm'n' ij (θ) .
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.