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

Dimension-variable invariant imbedding (DVIIM) T-matrix computational method for the light scattering simulation of atmospheric nonspherical particles

Open Access Open Access

Abstract

The invariant imbedding (IIM) T-matrix method has shown great potential in light scattering field. However, the T-matrix need to be calculated through the matrix recurrence formula derived from the Helmholtz equation, thus its computational efficiency is much lower than Extended Boundary Condition Method (EBCM). To alleviate this problem, the Dimension-Variable Invariant Imbedding (DVIIM) T-matrix method is presented in this paper. Compared with the traditional IIM T-matrix model, the dimensions of the T-matrix and relevant matrices are gradually increasing as the iteration performed step by step, thus the unnecessary operations of large matrices can be avoided in early iterations. To optimally determine the dimension of these matrices in each iterative calculation, the spheroid-equivalent scheme (SES) is also proposed. The effectiveness of the DVIIM T-matrix method is validated from the modeling accuracy and calculation efficiency. The simulation results show that compared with traditional T-matrix method, its modeling efficiency can be improved notably, especially for the particles with large size and aspect ratio, where for the spheroid with a aspect ratio of 0.5, the computational time is cut down by 25%. Though the dimension of the T matrix is cut down in the early iterations, the computational precision of DVIIM T-matrix model is not decreased notably, and a good agreement is achieved between the calculation results of DVIIM T-matrix method, IIM T-matrix method and other well-validated models (like EBCM and DDACSAT), where the relative errors of the integral scattering parameters (e.g., extinction, absorption, scattering cross sections) are generally less than 1%.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

To improve the simulation accuracy of atmospheric radiative transfer, the light scattering properties of aerosols and ice crystals should be accurately simulated [17]. However, due to their irregular shapes and inhomogeneous compositions [810], their light scattering properties haven’t been adequately understood, thus causing substantial uncertainties in radiative transfer simulation [11,12]. In recent years, with the development of polarized remote sensing, more and more researchers have found that the nonspherical shapes of aerosol and ice crystals have a significant influence on the polarized components of the radiation [2,12,13]. As a result, how to obtain the light scattering properties have become an important issue in the field of atmospheric remote sensing [6,1416].

To obtain the scattering properties of nonspherical particles, many scattering models are developed, such as GOA/IGOA (Geometrical Optics Approximation/Improved Geometrical Optics Approximation) [1719], PGHOM (Physical Geometric Optics Hybrid Method, a geometric optical model using a beam-splitting technique) [20,21], T-matrix method [2225], DDA (Discrete Dipole Approximation) [2628] and time domain methods (FDTD, PSTD and MRTD, etc.) [2935]. The semi-empirical geometric-optics method GOA/IGOA is suitable to the scattering simulation of large particles but it fails for small particles because the “ray” concept is not valid when the light wavelength is similar to particle size [20]. DDA and the time domain models are applicable to particles with moderate size parameters However, they only can calculate the light scattering properties of the particle with specific size and orientation in one wave-particle interaction [11]. Since the scattering parameters in the Radiative Transfer Models (RTMs) are volume-averaged values (namely, the scattering properties should be integrated over their size distributions and orientations [3,11]), in order to obtain the scattering parameters of the randomly oriented particles, the computational process should be repeated many times for particles with different orientations, and the process will become very time consuming.

In contrast, T matrix method has the incomparable advantage. Once the T matrix of the nonspherical particle is obtained, the scattering properties of randomly oriented particles can be calculated analytically [36,37]. Thus, it will be much more precise and efficient than other computational models. The T matrix can be calculated by the surface integral and volume integral scheme [38]. The surface integral scheme, which is also called EBCM (Extended Boundary Condition Method), is firstly proposed by P. C. Waterman [39,40]. After the improvement of Mishchenko and other scholars [23,37,4144], it has become a well-tested scattering model and widely applied in the fields of atmospheric radiation, bio-optics, astrophysics, etc. [38,45]. However, due to the difficulty in solving the surface integral equations, the EBCM T-matrix method is mainly good at the light scattering simulation by rotationally symmetric particles (like spheroidal and cylindrical particles, etc.) [42]. To extend the application scope of the T-matrix methods, the invariant imbedding (IIM) technique is also introduced into the calculation of the T matrix [38,4648]. Because the IIM T-matrix method is essentially derived from the Helmholtz volume integral equation, it can be applied to particles with arbitrary shapes and inhomogeneous compositions. After the efficient implementation of Bi L. and Yang P. [38,4749], the IIM T-matrix method has become a powerful tool for the light scattering simulation of nonspherical particles. In recently years, referring to the work of Bi L., we further derived the symmetrical properties in T-matrix method and proposed an efficient realization scheme of this model [5052]. Based on those works, a FORTRAN code of the IIM T-matrix method has also developed by our team [50,5254].

In principle, with the increasing of dimension of T-matrix, the higher accuracy can be achieved in the light scattering simulation. Generally, to guarantee the information content of the T matrix, the dimensions of T matrix and other matrices involved are fixed as lmax=k·Nmax(Nmax+2) (Nmax is the expansion order of the electric fields, which is dependent on the size of the nonspherical particle, and k=2 or 3) [38,53,55]. Therefore, for particles with large sizes, all the matrices involved in the computations are supermatrices, and most of the time is consumed on their inversion and multiplication processes. Besides, the T matrix should be updated step by step along the radial direction, all these operations of large matrices should be repeated in every iteration. Thus its computational burden is much heavier than that of EBCM T-matrix method [48]. Owning to this reason, how to improve the modeling efficiency has become an important issue of the IIM T-matrix method. From the principle of IIM T-matrix model, it can be found that in the early iterations, only a part of the particle (the portion in the nth spherical shell) is used. That is to say, the effective size for computation is gradually increased with the iteration performed. According to the T matrix theory, the smaller the particle size is, the smaller dimension of T matrix is needed. Therefore, we consider whether the dimension of the matrices can be reduced in early iterations. From this idea, a dimension-variable invariant imbedding T-Matrix method is proposed in this paper. In this model, the dimension of T matrix is gradually increasing with the increasing of the radius of spherical shell rather than fixed as specific value, i.e., 2Nmax (Nmax+2). In this way, the computational amount can be cut down notably, especially for particles with large size and aspect ratio.

This paper is organized as follows: in section 2, the basic principle of the Dimension-Variable IIM (DVIIM) T-Matrix method is introduced. Then, to optimally determine the dimension of the supermatrices in each iterative step, a spheroid-equivalent scheme is proposed in section 3. Further, the modeling accuracy is further validated for particles with different shapes and sizes, and computational efficiency of the DVIIM and IIM T-matrix methods are compared quantitatively. Section 7 is a summary of our work.

2. Basic principles of DVIIM T-matrix method

In the DVIIM T-matrix method, the nonspherical particle is also regarded as an inhomogeneous sphere (as shown is Fig. 1) and is discretized into several inhomogeneous spherical shells in the spherical coordinate system. Then, the T Matrix of the nonspherical particle is solved by combining the invariant imbedding technique with the Lorenz-Mie theory or EBCM, where the T matrix of inscribed sphere of the nonspherical particles is firstly calculated by Lorenz-Mie theory or EBCM (the yellow part in Fig. 1(a)), and then, in the spherical mantle region, the principle of the IIM is applied to obtain the T-matrix of a larger sphere of n layers based on the T-matrix of a smaller sphere with n–1 layers (see Fig. 1(c)). The iterative equation for the T matrix computation can be written as

$$\overline{\overline{\mathbf{T}}} ({r_n}) = \overline{\overline{\mathbf{Q}}} _{11}({r_n}) + (\overline{\overline{\mathbf{I}}} + \overline{\overline{\mathbf{Q}}} _{12}({r_n})){[{\overline{\overline{\mathbf{I}}} - \overline{\overline{\mathbf{T}}} ({r_{n - 1}}){\overline{\overline{\mathbf{Q}} }_{22}}({r_n})} ]^{ - 1}}\overline{\overline{\mathbf{T}}} ({r_{n - 1}})[{\overline{\overline{\mathbf{I}}} + {\overline{\overline{\mathbf{Q}} }_{21}}({r_n})} ],$$
where R0 is the radius of the inscribed sphere, $\overline{\overline{\mathbf{T}}} ({r_n})$ $\overline{\overline{\mathbf{T}}} ({r_n})$ and $\overline{\overline{\mathbf{T}}} ({r_{n - 1}})$ are the T matrices of the spheres with n layers and n-1 layers, $\overline{\overline{\mathbf{Q}}}_{11}\left(r_n\right), \overline{\overline{\mathbf{Q}}}_{12}\left(r_n\right), \overline{\overline{\mathbf{Q}}}_{21}\left(r_n\right)$ \;\text{and}\; $\overline{\overline{\mathbf{Q}}}_{22}\left(r_n\right)$ are the optical matrices of the nth spherical shell.

 figure: Fig. 1.

Fig. 1. A schematic diagram of the discretization of nonspherical particles and the computational regions for IIM technique and Lorenz-Mie theory (or EBCM).

Download Full Size | PDF

In principle, with the increasing of particle size, there will be more scattering information existed in the T matrix, and the dimension of matrices should be increased as well. Therefore, in traditional IIM T-matrix method, the dimensions of the T and Qij (i, j=1,2) are fixed as lmax=2·Nmax(Nmax+2) in all iterations to guarantee the convergence of the computation, where, Nmax is the expansion order of the electric fields, a parameter dependent on the size parameter of the nonspherical particle. According to the work of Wiscombe [56], Nmax generally can be obtained by Eq. (2) [41]

$${N_{\max }}(x) = x\textrm{ + }4.05{x^{1/3}} + 8;$$
where x=2πr/λ is size parameter of the nonspherical particle (λ is the light wavelength, r is the radius of the circumscribed sphere).

However, the effective size of the particle in the initial iterative steps is not as large as that of the last iteration (the effective size for computation is become larger and larger with the iteration process carried out, as shown in Fig. 2). Therefore, according to the T matrix theory, the expansion order of the electric field is not needed to be taken as large as Nmax in early iterations, and the dimensions of the T matrix can be reduced in the initial iterative process. Referring to this idea, a dimension-variable IIM T-matrix method is proposed. The basic principle of the method is shown in the Fig. 2. In the computation, the expansion order of the electric field is gradually increasing as effective size of the inner sphere becomes larger. By using this method, both the computational amount and memory consumption can be saved. In the implementation of this model, the expansion order of the electric field in the ith iteration, i.e., Ni, is firstly determined by the Spheroidal Equivalent Scheme (SES, see section 3 for detail), then the dimension of T, U, Q and Qij (i, j=1, 2) can be defined as

$$\textrm{For }\mathbf{T}\,\textrm{and}\,{\mathbf{Q}_{ij}}\textrm{matrices}:{l_i} = 2{N_i} \cdot ({N_i} + 2);$$
$$\textrm{For}\,\mathbf{U}\,\textrm{and}\,\mathbf{Q}\,\textrm{matrices}\,\textrm{:}\,{l_i} = 3{N_i} \cdot ({N_i} + 2). $$

 figure: Fig. 2.

Fig. 2. The basic principle of DVIIM T-Matrix method.

Download Full Size | PDF

After the dimension of these matrices is determined, the T matrix can be updated layer by layer by using Eq. (1).

3. Optimal determination of the T-matrix dimension and Gaussian integral points by spheroidal equivalent scheme (SES)

From section 2, we can know that it is very important to determine Ni in each iteration since it is the parameter to determine the dimension of the T, U, Q and Qij matrices. However, not only that, Ni is also an important parameter to determine the number of Gaussian discrete points of the azimuth and zenith integrals in the U matrix calculation, which will influence the modeling accuracy of the DVIIM T-matrix method. Due to these reasons, the expansion order Ni should be optimally determined considering both the computational efficiency and accuracy.

In the DVIIM T-matrix method, the computational accuracy of U matrix is very critical since the irregular shape of particles and their optical properties is manifested by the matrix, the expression of U matrix can be written as:

$$\scalebox{0.93}{$\displaystyle\begin{array}{l} \overline{\overline{\mathbf{U}}} _{mn{m^{\prime}}{n^{\prime}}}^{}(r) = \\ = {k^2}{r^2}{( - 1)^{m + {m^{\prime}}}}{\left[ {\frac{{2n + 1}}{{4\pi n(n + 1)}}} \right]^{1/2}}{\left[ {\frac{{2{n^{\prime}} + 1}}{{4\pi {n^{\prime}}({n^{\prime}} + 1)}}} \right]^{1/2}}\int\limits_0^{2\pi } {d\varphi \int\limits_0^\pi {d\theta \sin \theta \exp [{ - i(m - {m^{\prime}})\varphi } ]} } [{{\varepsilon_r}(r,\theta ,\varphi ) - 1} ]\\ \times \left( {\begin{array}{ccc} {{\pi_{mn}}(\theta ){\pi_{{m^{\prime}}{n^{\prime}}}}(\theta ) + {\tau_{mn}}(\theta ){\tau_{{m^{\prime}}{n^{\prime}}}}(\theta )}&{ - i[{{\pi_{mn}}(\theta ){\tau_{{m^{\prime}}{n^{\prime}}}}(\theta ) + {\tau_{mn}}(\theta ){\pi_{{m^{\prime}}{n^{\prime}}}}(\theta )} ]}&0\\ {i[{{\pi_{mn}}(\theta ){\tau_{{m^{\prime}}{n^{\prime}}}}(\theta ) + {\tau_{mn}}(\theta ){\pi_{{m^{\prime}}{n^{\prime}}}}(\theta )} ]}&{{\pi_{mn}}(\theta ){\pi_{{m^{\prime}}{n^{\prime}}}}(\theta ) + {\tau_{mn}}(\theta ){\tau_{{m^{\prime}}{n^{\prime}}}}(\theta )}&0\\ 0&0&{\frac{{\sqrt {n(n + 1){n^{\prime}}({n^{\prime}} + 1)} d_{0m}^nd_{0{m^{\prime}}}^{{n^{\prime}}}}}{{{\varepsilon_r}(r,\theta ,\varphi )}}} \end{array}} \right) \end{array}$}$$
where r is radius of the nth spherical layer, k is wavenumber, $\theta$ and $\varphi$ are the zenith angle and azimuth angle respectively, ${\varepsilon _r}(r,\theta ,\varphi )$ is the permittivity of the medium; $d_{0m}^n$ is the Wigner-d function, ${\pi _{mn}}(\theta )$ and ${\tau _{mn}}(\theta )$ are the angular functions. To guarantee both the simulation accuracy and efficiency of the computation, similar to the way adopted by Mishchenko, the number of the Gaussian discrete points for the integral over the zenith and azimuth angles are taken as [41,42]
$${N_{ZEN}} = 2 \cdot NDGS \cdot {N_i};$$
$${N_{AZI}} = 2 \cdot {N_{ZEN}} = 4 \cdot NDGS \cdot {N_i};$$
where ${N_{ZEN}}$ and ${N_{AZI}}$ are the number of the discrete points for zenith and azimuth integrals (since the variation range of the azimuth angle is twice that of zenith angle, therefore ${N_{AZI}}$ is set as $2 \cdot {N_{ZEN}}$ in our IIM T-Matrix method). From the equations above, on the premise of ensuring the simulation accuracy, not only the expansion order of the electric field should be optimally determined, but also NDGS need to be determined properly.

To reasonably determine Ni and NDGS in each iterative process, a Spheroid-Equivalent Scheme (SES) is proposed. The basic principle of SES is illustrated as follows: the particle in the ith spherical shell is firstly equivalent as a spheroidal particle according to its size and shape. Then the optimal expansion order Ni and NDGS is obtained from the convergent condition obtained by the EBCM T-matrix method for the equivalent spheroid. Taking the ith iterative step as the example, the implementation of SES can be divided into 2 steps, as shown in Fig. 3.

 figure: Fig. 3.

Fig. 3. The basic principle of spheroid-equivalent scheme. The key point of this method is to select a spheroid which is similar to the particle in the spherical shell and determine the expansion order using EBCM.

Download Full Size | PDF

Step I: determining the shape parameter of the equivalent spheroidal particle. According to the shape of the particle in the ith spherical layer, the maximum aspect ratio and equivalent size parameter is extracted firstly, where the aspect ratio is regarded as the ratio of radius of the ith spherical layer and that of the inscribed sphere, and the equivalent size parameter is set as that of the inscribed sphere, written as

$${\varepsilon _{eq}} = {r_i}/{R_0};$$
$${X_{eq}} = 2\pi {r_i}/\lambda;$$
where ${\varepsilon _{eq}}$ and ${X_{eq}}$ are the equivalent aspect ratio and size parameter, respectively. ${r_i}$ and ${R_0}$ are the radius of the ith spherical shell and the inscribed sphere.

Step II: the optimal determination of Ni and NDGS. To facilitate the determination of the expansion order of electric field and Gaussian discrete points, a Lookup Table (LUT) of Ni and NDGS is established for spheroidal particles with different size parameters, aspect ratios and refractive indices by using the EBCM T-matrix method provided by Mishchenko. The convergence condition of the EBCM model is set as

$$\frac{{|{{C_{ext}}({N_{\max }},NDGS) - {C_{ext}}({N_{\max }} - 1,NDGS)} |}}{{{C_{ext}}({N_{\max }})}} < 0.001,$$
where the size parameter of the spheroid is calculated by Xs=2πb/λ (b is the half length of long axis), and part of the LUT is presented in Table 1. For spheroid particles with large size and refractive index, the EBCM might not be convergent in the computation. In this case, Ni is taken as Nmax, namely, the expansion order determined by Eq. (2).

Tables Icon

Table 1. The optimal value of Nmax and NDGS for spheroidal particles with different sizes, aspect ratios and refractive indices.

After the LUT is established, then NDGS for particles with an equivalent size parameter and aspect ratio of ${\varepsilon _{eq}}$ and ${X_{eq}}$ can be obtained by

$$NDGS = \max \{{LUT({\varepsilon_{eq}},{X_{eq}},m),\textrm{ }4} \};$$
where $LUT({\varepsilon _{eq}},{X_{eq}},m)$ denotes the linear interpolation in the established lookup table.

For the expansion order Ni, it is obtained in the following manner, expressed as

$${N_i} = \max \{{LUT({\varepsilon_{eq}},{X_{eq}},m),{N_r}} \};$$
$${N_r}\textrm{ = }\left( {\frac{{2\pi {r_i}}}{\lambda }} \right)\textrm{ + }4.05{\left( {\frac{{2\pi {r_i}}}{\lambda }} \right)^{1/3}} + 8;$$
where ${r_i}$ is the radius of the ith spherical shell, λ is the wavelength, ${N_r}$ is the expansion order for the sphere in the ith iteration using Wiscombe’s rule [56].

4. Implementation of DVIIM T-matrix computational code

Based on the principle proposed above, we have developed the FORTRAN code of the DVIIM T-matrix method. As shown in Fig. 4, the computational process of DVIIM T-matrix method can be concluded as following 3 steps:

  • Step I: The initialization of the model. This function of this part is to do the necessary initialization of the light scattering simulation. In the part, the refractive index of the nonspherical particle should be firstly converted into the permittivity, and radius of the inscribed and circumscribed sphere of the nonspherical particle should be determined. Then, according to the size of the particle, the expansion order of the electric field and the dimensions of matrices T, U, Q, Qij (i, j=1,2) are initially determined.
  • Step II: T-matrix computational process. In this part, the T matrix of the inscribed sphere is firstly calculated by the Lorenz-Mie theory and EBCM, and then the T matrix of the nonspherical particle is updated layer by layer by Eq. (1). In the computational process, the expansion order of the electric field is firstly determined by the SES technique; and based on the results, the dimensions of the matrices and the Gaussian discrete points for the zenith and azimuth integrals can also be determined (see section 3 for detail). The optical properties of the spherical shells, namely, the U matrix, is firstly calculated based on the particle’s geometry, and then the Q, Q11, Q12, Q21 can be Q22 could be calculated based on U matrix. To save the computer memory, the memory consumed by U, Q, Qij and $\mathbf{J}({r_n})$ should be deallocated before the scattering parameter computational process.
  • Step III: The scattering parameter computational process. This process is to calculate the scattering parameters based on the T-matrix obtained by the DVIIM T-Matrix method. In this process, both the scattering parameter of particles with specified orientation and randomly orientations can be obtained. The scattering parameters calculated in this model includes the extinction, absorption and scattering cross sections, single scattering albedo, asymmetric factor, and scattering phase matrix.

 figure: Fig. 4.

Fig. 4. The computational flowchart of the implementation of DVIIM T-matrix method.

Download Full Size | PDF

Different from traditional invariant imbedding T-matrix method, the most important in the whole calculation process is how to update the dimension of the matrix in step II, which will seriously affect the calculation accuracy of DVIIM T-matrix method. And it should be noted that it is not necessary to allocate and deallocate the computational memory for the large matrices (e.g., T, U and Q matrices, etc.) in each iterative step. We can allocate the memory according to the highest order of electric field Nmax in model’s initial process; and in the computational process, only the sub-matrix required in computation is applied, while the remaining matrix elements in are set to be zero. In this way, the repeated memory allocation processes can be avoided.

5. Model validation and results analysis

To validate the modeling accuracy of DVIIM T-matrix method, its simulation results (including the scattering phase matrices and integral scattering parameters such as extinction, scattering and absorption cross sections) are compared with those obtained by EBCM T-matrix method [36,57], DDASCAT [27,58], and IIM T-matrix method [51].

5.1 Spheroidal particle case

In this case, the computational results of DVIIM T-matrix method are compared with those of EBCM T-matrix method for three spheroidal particles with different sizes, and the results are presented in Fig. 5 and Fig. 6. Figure 5 shows the scattering phase matrices of particles with specific orientations, and Fig. 6 are the scattering matrices of randomly oriented particles. In this test, the light wavelength is taken as 0.5µm, the refractive index is set as m= 1.20+0.0008i, the half-lengths of the horizontal and rotational axis are set as (a, b)=(1µm, 2µm), (2µm, 4µm) and (5µm, 7µm), respectively.

 figure: Fig. 5.

Fig. 5. The comparisons of scattering phase matrices obtained by EBCM and DVIIM T-matrix methods for the single-oriented spheroidal particles with different sizes (the incident light is parallel to the rotational axis).

Download Full Size | PDF

 figure: Fig. 6.

Fig. 6. The comparisons of scattering phase matrices obtained by EBCM and DVIIM T-matrix methods for the randomly oriented spheroidal particles with different sizes.

Download Full Size | PDF

As can be seen from these figures, a good consistency is obtained between the results of the EBCM and DVIIM T-matrix method for both the particles with specific orientation and random orientations. For particles with specific orientation, the relative errors of the scattering phase functions are all with 10% in scattering angles ranging from 0° to 90°, and the absolute errors of F12/F11 are generally smaller than 0.1, which indicating that the DVIIM T-matrix method can simulate the scattering parameters of spheroidal particles with a high precision. For the particles with random orientations, the modeling accuracy of DVIIM T-matrix method is much higher than particles with single orientation, where the relative errors of the F11 are smaller than 1% in most scattering directions, and the absolute errors of F12/F11 are generally less than 0.03. On the whole, the simulation accuracy of the DVIIM T-matrix method in forward directions is higher than that in large scattering angle, which similar to that of the traditional IIM T-matrix method.

To further illustrate whether the decreasing of the dimension of T matrix in early iterations will influence the simulation accuracy, the scattering phase matrix calculated by DVIIM T-matrix method is also compared with that of the traditional IIM T-matrix method quantitatively, the results are shown in the Table 2.

Tables Icon

Table 2. The scattering phase matrices obtained by DVIIM and IIM T-matrix methods for the spheroidal particle with (a, b) = (2µm, 4µm).

As can be seen, a good agreement is achieved between the results of DVIIM and IIM T-matrix methods. For the scattering phase function, F11, the relative differences of the results are generally smaller than 1%, the absolute differences of F12/F11, F33/F11 and F34/F11 are generally smaller than 0.01. Therefore, it can be concluded that the decreasing of the T matrix dimension in the early iterative steps will not influence the modeling accuracy of the T matrix model notably for spheroidal particles.

Moreover, the integral parameters calculated by IIM and DVIIM T-matrix methods are also compared with each other, the results are shown in Table 3. From the table, it can be found that the results of DVIIM T-matrix method show a good agreement with those obtained by IIM T-matrix method, indicating that the DVIIM T-matrix method can achieve similar computational accuracy of the IIM T-matrix method for spheroidal particles.

Tables Icon

Table 3. The integral scattering parameters obtained by DVIIM and IIM T-matrix methods for the spheroidal particle with different sizes.

5.2 Cylinder particle case

The modeling accuracy of the DVIIM T-matrix method is also tested for cylindrical particles, the scattering phase matrices calculated by different models are shown in Fig. 7 and Fig. 8. In this simulation, three cylinders with different sizes are used for model validation, their diameters and lengths are set as (D, H)=(2µm, 2µm), (4µm, 4µm) and (8µm, 8µm), respectively.

 figure: Fig. 7.

Fig. 7. The comparisons of scattering phase matrices obtained by EBCM and DVIIM T-matrix methods for the single-oriented cylindrical particles with different sizes (the incident light is parallel to the rotational axis).

Download Full Size | PDF

 figure: Fig. 8.

Fig. 8. The comparisons of scattering phase matrices obtained by EBCM and DVIIM T-matrix methods for the randomly oriented cylindrical particles with different sizes.

Download Full Size | PDF

From the results, it can be seen that the curves of the matrix elements obtained by DVIIM T-matrix method is almost coincided with those of EBCM, where for F11, the relative differences are generally smaller than 10%, and for F12/ F11, their absolute errors are with 0.1 for most scattering directions. Similar to the spheroidal particle case, the simulation accuracy of DVIIM is much higher for the randomly oriented particles, and the relative differences of the scattering phase functions are smaller than 2% in most scattering angles. From the discussion above, it can be found that the high modeling accuracy can be achieved by the DVIIM T-matrix method in the light scattering simulation of cylinders.

Similar to the spheroidal particles, the scattering phase matrix calculated by DVIIM T-matrix method is also compared with the traditional IIM T-matrix method, and the results are presented in Table 4. From the results, it can be found that the difference between the results of DVIIM and IIM T-matrix methods is small, where the relative differences of F11 are generally less than 1%, and the absolute differences of F12/F11 are smaller than 0.01 for most scattering angles. From the results, it can be concluded that the spheroid-equivalent scheme will not cause the decreasing of the modeling accuracy remarkably in spite of the decreasing of dimension of the T matrix in the early iterative computations.

Tables Icon

Table 4. The scattering phase matrices obtained by DVIIM and IIM T-matrix methods for the cylindrical particle with (D, L) = (2µm, 2µm).

Besides, the integral scattering parameters of DVIIM and IIM T-matrix methods are also compared with each other, and the results are shown in Table 5. It can be seen that the cross sections calculated by DVIIM T-matrix method show a good agreement with those of the traditional one, where the relative differences of Cext, Cab, and Csc are generally smaller than 1%, which further validates both the modeling accuracy of the DVIIM T-matrix method.

Tables Icon

Table 5. The integral scattering parameters obtained by DVIIM and IIM T-matrix methods for the cylindrical particle with different sizes.

5.3 Hexagonal ice crystals case

To further test the modeling accuracy of the DVIIM T-matrix method for particles with non-rotational symmetrical geometry, the simulation results is also validated against DDASCAT for hexagonal particles, and the results of the random-orientation case are presented in Fig. 9. In this case, the wavelength of the incident light is set as λ=0.5µm, the refractive index of the particles are all set as m=1.22+0.0008i, the shape parameters (a, L) are set as (1µm, 2µm) and (2µm, 4µm), respectively.

 figure: Fig. 9.

Fig. 9. The comparisons of scattering phase matrices obtained by EBCM and DVIIM T-matrix methods for the random oriented hexagonal particles with different sizes.

Download Full Size | PDF

It can be found that a good consistency is also achieved between the results of DVIIM T-matrix method and DDASCAT, and the curves of the phase matrix elements obtained by different models are very close to each other. For the F11, their relative differences are generally smaller than 5% in forward scattering directions, though the differences in large scattering angles are slightly larger, but they are with 15% for most directions. The high agreement indicates that the DVIIM T-matrix method can simulate the light scattering process of ice crystals with a high precision.

To further investigate the spheroid-equivalent scheme whether will cause the decreasing of the modeling accuracy of invariant imbedding T-matrix method for hexagonal ice crystals, the scattering phase matrices obtained by DVIIM and IIM T-matrix methods are also compared with each other, the results are illustrated in Table 6.

Tables Icon

Table 6. The computational results for the IIM and DVIIM T-Matrix models for ice crystal particle (R = 1µm, L = 2µm).

As can be seen, the computational results of DVIIM T-matrix method agree with those of IIM T-matrix methods well, where for F12/F11 and F44/F11, their absolute differences are generally smaller than 0.01, which validates the effectiveness of the spheroid-equivalent scheme and the modeling accuracy of the DVIIM T-matrix method.

6. Discussion of model efficiency

To further investigate the modeling efficiency of the DVIIM T-matrix method, the computational time required by IIM and DVIIM T-matrix methods are also compared for particles with different sizes and shapes. In these simulations, the light wavelength is set as 0.5µm, and the scatters are set as spheroidal particles. All the scattering processes are simulated on the same computer (32bit 3.1 GHz, Windows operation system), and the computational time of the IIM and DVIIM T-Matrix codes is presented in Table 7 and Table 8.

Tables Icon

Table 7. The computational time needed for DVIIM and IIM T-Matrix models for the particles with different aspect ratios (b = 3µm)

Tables Icon

Table 8. The computational time needed for the traditional and improved IIM T-Matrix models for particles with different sizes (a/b = 2/3)

As can been seen, on the whole, the computational time needed for DVIIM T-matrix method is less than that required by the IIM T-matrix method. With the increasing of the aspect ratio (a/b), the computational efficiency of DVIIM T-matrix method becomes higher as well. It can also be found that with the increasing of particle size, the improvement of the computational efficiency becomes much more remarkable. The discussion above indicates that the Spheroid-Equivalent Scheme is effective, and can improve the computational efficiency notably.

7. Conclusion

With the development of the computational technology, the IIM T-matrix method has become one of the most promising models for the light scattering simulation. Because this model is based on volume-integral form of the Helmholtz equation, it is suitable for the particles with arbitrary shapes, which greatly expands the application scope of the T-matrix method. However, since the IIM T-matrix method should update the T-matrix layer by layer, how to improve the computational efficiency and stability has become a key problem to be solved. To further improve the modeling efficiency of the IIM T-matrix method, a Dimension-Variable Invariant Imbedding (DVIIM) T-matrix method is designed in this paper, in which the dimension of the T matrix can be changed according to the actual size and shape in the recurrent iteration. To optimally determine the expansion order of the electric field, the Spheroid-Equivalent Scheme is proposed. To validate the computational accuracy of DVIIM T-matrix method, its calculation results are compared with those of EBCM, traditional IIM T-matrix method and DDASCAT. Moreover, the calculation efficiency of the model is also quantitatively compared with the traditional IIM T-matrix method. The main conclusion of this paper can be drawn as follows:

  • (1) The calculation results obtained by the DVIIM T-matrix method show a good agreement with the EBCM, DDASCAT and IIM T-matrix method, which indicates that the computational precision of DVIIM T-matrix method is not decreased notably though the dimension of the T matrix is reduced in the early recurrent iterations.
  • (2) The modeling efficiency of DVIIM T-matrix method is higher than that of traditional IIM–matrix method. With the increasing of particle size and aspect ratio, the computational time consumed can be cut down more notably.

It should also be noted that, the expansion order of the electric field and NDGS is determined from the spheroid-equivalent scheme (SES), namely, we taken the particles in the spherical shells equivalent as spheroid, therefore it mainly suitable for convex particles with smooth surface, while for particles with very complex geometry or sharp edges, its computational accuracy might not be so high as the particles calculated in this paper. Besides, because the spheroid-equivalent scheme is established based on the convergence of the EBCM T-matrix method, the DVIIM T-matrix computational method might not be very good at light scattering of particles with very large axial ratio.

Funding

National Natural Science Foundation of China (42175154, 62105367); Natural Science Foundation of Hunan Province (2020JJ4662, 2021JJ40666).

Acknowledgments

Thanks for the Tmatrix code provided by Prof. M. I. Mishchenko. Thanks for the DDSCAT code provided by Prof. Bruce T. Draine and Piotr J. Flatau.

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. M. I. Mishchenko and L. D. Travis, “Satellite retrieval of aerosol properties over the ocean using polarization as well as intensity of reflected sunlight,” J. Geophys. Res. 102(D14), 16989–17013 (1997). [CrossRef]  

2. O. Dubovik, A. Sinyuk, T. Lapyonok, B. N. Holben, M. Mishchenko, P. Yang, T. F. Eck, H. Volten, O. Munoz, B. Veihelmann, W. J. v, D. Zande, J. F. Leon, M. Sorokin, and I. Slutsker, “Application of spheroid models to account for aerosol particle nonsphericity in remote sensing of desert dust,” J. Geophys. Res. 111(D11), D11208 (2006). [CrossRef]  

3. K. N. Liou, An Introduction to Atmospheric Radiation (Academic Press, 2003), Chap. 1.

4. F. Zhang, K. Wu, J. Li, Q. Yang, J. Q. Zhao, and J. Li, “Analytical Infrared Delta-Four-Stream Adding Method from Invariance Principle,” J. Atmos. Sci. 73(10), 4171–4188 (2016). [CrossRef]  

5. F. Zhang and J. Li, “Doubling–Adding Method for Delta-Four-Stream Spherical Harmonic Expansion Approximation in Radiative Transfer Parameterization,” J. Atmos. Sci. 70(10), 3084–3101 (2013). [CrossRef]  

6. H. Letu, K. Yang, T. Y. Nakajima, H. Ishimoto, T. M. Nagao, J. Riedi, A. J. Baran, R. Ma, T. Wang, H. Shang, P. Khatri, L. Chen, C. Shi, and J. Shi, “High-resolution retrieval of cloud microphysical properties and surface solar radiation using Himawari-8/AHI next-generation geostationary satellite,” Remote Sens Environ. 239, 111583 (2020). [CrossRef]  

7. H. Letu, T. Y. Nakajima, and T. N. Matsui, “Development of an ice crystal scattering database for the global change observation mission/second generation global imager satellite mission: investigating the refractive index grid system and potential retrieval error,” Appl. Opt. 51(25), 6172–6178 (2012). [CrossRef]  

8. H. Xiao, F. Zhang, Q. He, P. Liu, F. Yan, L. Miao, and Z. Yang, “Classification of Ice Crystal Habits Observed From Airborne Cloud Particle Imager by Deep Transfer Learning,” Earth Space Sci. 6(10), 1877–1886 (2019). [CrossRef]  

9. K.-N. Liou and P. Yang, light scattering by ice crystals (Cambridge University, 2016), pp. 1–14.

10. M. Bailey and J. Hallett, “Growth Rates and Habits of Ice Crystals between -20° and -70°C,” J. Atmos. Sci. 61(5), 514–544 (2004). [CrossRef]  

11. M. I. Mishchenko, J. W. Hovenier, and L. D. Travis, light scattering by nonspherical particles, Thoery, Measurements, and Application (Academic Press, 2000), Chap. 1.

12. T. H. Cheng, X. F. Gu, T. Yu, and G. L. Tian, “The reflection and polarization properties of non-spherical aerosol particles,” J. Quant. Spectrosc. Radiat. Transfer 111(6), 895–906 (2010). [CrossRef]  

13. M. Herman, J. L. Deuzé, A. Marchand, B. Roger, and P. Lallart, “Aerosol remote sensing from POLDER//ADEOS over the ocean:Improved retrieval using a nonspherical particle model,” J. Geophys. Res. 110(D10), D10S02 (2005). [CrossRef]  

14. J. Q. Zhao and Y. Q. Hu, “Bridging technique for calculating the extinction efficiency of arbitrary shaped particles,” Appl. Opt. 42(24), 4937–4945 (2003). [CrossRef]  

15. J. Q. Zhao, G. Shi, H. Che, and G. Cheng, “Approximation of scattering phase function of particles,” Adv. Atmos. Sci. 23(5), 802–808 (2006). [CrossRef]  

16. H. Letu, H. Ishimoto, J. Riedi, T. Y. Nakajima, L. C. Labonnote, A. J. Baran, T. M. Nagao, and M. Sekiguchi, “Investigation of ice particle habits to be used for ice cloud remote sensing for the GCOM-C satellite mission,” Atmos. Chem. Phys. 16(18), 12287–12303 (2016). [CrossRef]  

17. A. Macke, M. I. Mishchenko, K. Muinonen, and B. E. Carlson, “Scattering of light by large nonspherical particles: ray-tracing approximation versus T-matrix method,” Opt. Lett. 20(19), 1934 (1995). [CrossRef]  

18. P. Yang and K. N. Liou, “Geometric-optics–integral-equation method for light scattering by nonspherical ice crystals,” Appl. Opt. 35(33), 6568–6584 (1996). [CrossRef]  

19. Q. Cai and K.-N. Liou, “Polarized light scattering by hexagonal ice crystals: theory,” Appl. Opt. 21(19), 3569–3580 (1982). [CrossRef]  

20. L. Bi, P. Yang, G. W. Kattawar, Y. Hu, and B. A. Baum, “Scattering and absorption of light by ice particles: Solution by a new physical-geometric optics hybrid method,” J. Quant. Spectrosc. Radiat. Transfer 112(9), 1492–1508 (2011). [CrossRef]  

21. 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]  

22. M. I. Mishchenko, “Electromagnetic scattering by nonspherical particles:A tutorial review,” J. Quant. Spectrosc. Radiat. Transfer 110(11), 808–832 (2009). [CrossRef]  

23. M. I. Mishchenko, “Calculation of the amplitude matrix for a nonspherical particle in a fixed orientation,” Appl. Opt. 39(6), 1026–1031 (2000). [CrossRef]  

24. A. Quirantes, “A T-matrix method and computer code for randomly oriented, axially symmetric coated scatterers,” J. Quant. Spectrosc. Radiat. Transfer 92(3), 373–381 (2005). [CrossRef]  

25. L. Liu, M. I. Mishchenko, and B. Cairns, “Modeling single-scattering properties of small cirrus particles by use of a size-shape distribution of ice spheroids and cylinders,” J. Quant. Spectrosc. Radiat. Transfer 101(3), 488–497 (2006). [CrossRef]  

26. M. A. Yurkin and A. G. Hoekstra, “The discrete-dipole-approximation code ADDA: Capabilities and known limitations,” J. Quant. Spectrosc. Radiat. Transfer 112(13), 2234–2247 (2011). [CrossRef]  

27. B. T. Draine and P. J. Flatau, “Discrete-dipole approximation for scattering calculations,” J. Opt. Soc. Am. A 11(4), 1491–1499 (1994). [CrossRef]  

28. J. J. Goodman, B. T. Draine, and P. J. Flatau, “Application of fast-Fourier-transform techniques to the discrete-dipole approximation,” Opt. Lett. 16(15), 1198–1200 (1991). [CrossRef]  

29. P. Yang and K. N. Liou, “Finite-difference time domain method for light scattering by small ice crystals in three-dimensional space,” J. Opt. Soc. Am. A 13(10), 2072–2085 (1996). [CrossRef]  

30. C. Liu, R. L. Panetta, and P. Yang, “Application of the pseudo-spectral time domain method to compute particle single-scattering properties for size parameters up to 200,” J. Quant. Spectrosc. Radiat. Transfer 113(13), 1728–1740 (2012). [CrossRef]  

31. S. Hu, T. Gao, L. Liu, H. Li, M. Chen, and B. Yang, “Application of the weighted total field-scattering field technique to 3D-PSTD light scattering model,” J. Quant. Spectrosc. Radiat. Transfer 209, 58–72 (2018). [CrossRef]  

32. S. Hu, T. Gao, H. Li, L. Liu, M. Chen, and B. Yang, “Light-scattering model for aerosol particles with irregular shapes and inhomogeneous compositions using a parallelized pseudo-spectral time-domain technique,” Chin. Phys. B 27(5), 054215 (2018). [CrossRef]  

33. S. Hu, T. Gao, H. Li, B. Yang, F. Zhang, M. Chen, and L. Liu, “Light scattering computation model for nonspherical aerosol particles based on multi-resolution time-domain scheme: model development and validation,” Opt. Express 25(2), 1463–1686 (2017). [CrossRef]  

34. S. Hu, T. Gao, H. Li, B. Yang, Z. Jiang, L. Liu, and M. Chen, “Application of Convolution Perfectly Matched Layer in MRTD scattering model for non-spherical aerosol particles and its performance analysis,” J. Quant. Spectrosc. Radiat. Transfer 200, 1–11 (2017). [CrossRef]  

35. W. Sun and Q. Fu, “Finite-difference time-domain solution of light scattering by dielectric particles with large complex refractive indices,” Appl. Opt. 39(30), 5569–5578 (2000). [CrossRef]  

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

37. M. I. Mishchenko, L. D. Travis, and A. Macke, “Scattering of light by polydisperse, randomly oriented, finite circular cylinders,” Appl. Opt. 35(24), 4927–4940 (1996). [CrossRef]  

38. 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. Transfer 116, 169–183 (2013). [CrossRef]  

39. P. C. Waterman, “Symmetry, Unitarity, and Geometry in Electromagnetic Scattering,” Phys. Rev. D 53(3), 825–839 (1971). [CrossRef]  

40. P. C. Waterman, “Matrix Formulation of Electromagnetic Scattering,” Proc. IEEE 53(8), 805–812 (1965). [CrossRef]  

41. M. I. Mishchenko, L. D. Travis, and A. A. Lacis, Scattering, Absorption, and Emission of Light by Small Particles (Cambridge University, 2002), Chap. 5.

42. M. I. Mishchenko and L. D. Travis, “Capabilities and Limitations of a Current Fortran Implementation of the T-Martrix Method for Randomly Oriented,Rotationally Symmetric Scatterers,” J. Quant. Spectrosc. Radiat. Transfer 60(3), 309–324 (1998). [CrossRef]  

43. 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]  

44. S. Havemann and A. J. Baran, “Extension of T-matrix to scattering of electromagnetic plane waves by non-axisymmetric dielectric particles: application to hexagonal ice cylinders,” J. Quant. Spectrosc. Radiat. Transfer 70(2), 139–158 (2001). [CrossRef]  

45. L. Bi and P. Yang, “Modeling of light scattering by biconcave and deformed red blood cells with the invariant imbedding T-matrix method,” J. Biomed. Opt. 18(5), 055001 (2013). [CrossRef]  

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

47. 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. Transfer 138, 17–35 (2014). [CrossRef]  

48. L. Bi, P. Yang, G. W. Kattawar, and M. I. Mishchenko, “A numerical combination of extended boundary condition method and invariant imbedding method applied to light scattering by large spheroids and cylinders,” J. Quant. Spectrosc. Radiat. Transfer 123, 17–22 (2013). [CrossRef]  

49. B. Q. Sun, L. Bi, P. Yang, M. Kahnert, and X. Zhang, Invariant Imbedding T-matrix method for light scattering by nonspherical and inhomogeneous particles (Elsevier, 2019), Chap. 3.

50. S. Hu, L. Liu, Q. Zeng, T. Gao, and F. Zhang, “An investigation of the symmetrical properties in the invariant imbedding T-matrix method for the nonspherical particles with symmetrical geometry,” J. Quant. Spectrosc. Radiat. Transfer 259, 107401 (2021). [CrossRef]  

51. S. Hu, L. Liu, T. Gao, Q. Zeng, F. Zhang, and X. Liu, “Efficient design of the realization scheme of the invariant imbedding (IIM) T-matrix light scattering model for atmospheric nonspherical particles,” J. Quant. Spectrosc. Radiat. Transfer 251, 106999 (2020). [CrossRef]  

52. S. Hu, L. Liu, T. Gao, Q. Zeng, and X. Liu, “An efficient implementation of the light scattering simulation for random-oriented Non-Rotationally symmetric particles using Invariant Imbedding T-matrix method,” J. Quant. Spectrosc. Radiat. Transfer 241, 106734 (2020). [CrossRef]  

53. S. Hu, L. Liu, T. Gao, and Q. Zeng, “Design and Validation of the Invariant Imbedded T-Matrix Scattering Model for Atmospheric Particles with Arbitrary Shapes,” Appl. Sci. 9(20), 4423 (2019). [CrossRef]  

54. S. Hu, “Research on the Numerical Computational Models and Application of the scattering Properties of Nonspherical Atmospheric Particles,” in College of Meteorology and Oceanography (National University of Defense Technology, 2018), chap. 5.

55. S. Zhai, R. L. Panetta, and P. Yang, “Improvements in the computational efficiency and convergence of the Invariant Imbedding T-matrix method for spheroids and hexagonal prisms,” Opt. Express 27, A1441–A1457 (2019). [CrossRef]  

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

57. M. I. Mishchenko and L. D. Travis, “T-Martrix computations of light scattering by large spheriodal particles,” Opt. Commun. 109(1-2), 16–21 (1994). [CrossRef]  

58. B. T. Draine, “Discrete-dipole approximation and its application to interstellar graphite grains,” Astrophys. J. 333, 848–872 (1988). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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. A schematic diagram of the discretization of nonspherical particles and the computational regions for IIM technique and Lorenz-Mie theory (or EBCM).
Fig. 2.
Fig. 2. The basic principle of DVIIM T-Matrix method.
Fig. 3.
Fig. 3. The basic principle of spheroid-equivalent scheme. The key point of this method is to select a spheroid which is similar to the particle in the spherical shell and determine the expansion order using EBCM.
Fig. 4.
Fig. 4. The computational flowchart of the implementation of DVIIM T-matrix method.
Fig. 5.
Fig. 5. The comparisons of scattering phase matrices obtained by EBCM and DVIIM T-matrix methods for the single-oriented spheroidal particles with different sizes (the incident light is parallel to the rotational axis).
Fig. 6.
Fig. 6. The comparisons of scattering phase matrices obtained by EBCM and DVIIM T-matrix methods for the randomly oriented spheroidal particles with different sizes.
Fig. 7.
Fig. 7. The comparisons of scattering phase matrices obtained by EBCM and DVIIM T-matrix methods for the single-oriented cylindrical particles with different sizes (the incident light is parallel to the rotational axis).
Fig. 8.
Fig. 8. The comparisons of scattering phase matrices obtained by EBCM and DVIIM T-matrix methods for the randomly oriented cylindrical particles with different sizes.
Fig. 9.
Fig. 9. The comparisons of scattering phase matrices obtained by EBCM and DVIIM T-matrix methods for the random oriented hexagonal particles with different sizes.

Tables (8)

Tables Icon

Table 1. The optimal value of Nmax and NDGS for spheroidal particles with different sizes, aspect ratios and refractive indices.

Tables Icon

Table 2. The scattering phase matrices obtained by DVIIM and IIM T-matrix methods for the spheroidal particle with (a, b) = (2µm, 4µm).

Tables Icon

Table 3. The integral scattering parameters obtained by DVIIM and IIM T-matrix methods for the spheroidal particle with different sizes.

Tables Icon

Table 4. The scattering phase matrices obtained by DVIIM and IIM T-matrix methods for the cylindrical particle with (D, L) = (2µm, 2µm).

Tables Icon

Table 5. The integral scattering parameters obtained by DVIIM and IIM T-matrix methods for the cylindrical particle with different sizes.

Tables Icon

Table 6. The computational results for the IIM and DVIIM T-Matrix models for ice crystal particle (R = 1µm, L = 2µm).

Tables Icon

Table 7. The computational time needed for DVIIM and IIM T-Matrix models for the particles with different aspect ratios (b = 3µm)

Tables Icon

Table 8. The computational time needed for the traditional and improved IIM T-Matrix models for particles with different sizes (a/b = 2/3)

Equations (12)

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

T ¯ ¯ ( r n ) = Q ¯ ¯ 11 ( r n ) + ( I ¯ ¯ + Q ¯ ¯ 12 ( r n ) ) [ I ¯ ¯ T ¯ ¯ ( r n 1 ) Q ¯ ¯ 22 ( r n ) ] 1 T ¯ ¯ ( r n 1 ) [ I ¯ ¯ + Q ¯ ¯ 21 ( r n ) ] ,
N max ( x ) = x  +  4.05 x 1 / 3 + 8 ;
For  T and Q i j matrices : l i = 2 N i ( N i + 2 ) ;
For U and Q matrices : l i = 3 N i ( N i + 2 ) .
N Z E N = 2 N D G S N i ;
N A Z I = 2 N Z E N = 4 N D G S N i ;
ε e q = r i / R 0 ;
X e q = 2 π r i / λ ;
| C e x t ( N max , N D G S ) C e x t ( N max 1 , N D G S ) | C e x t ( N max ) < 0.001 ,
N D G S = max { L U T ( ε e q , X e q , m ) ,   4 } ;
N i = max { L U T ( ε e q , X e q , m ) , N r } ;
N r  =  ( 2 π r i λ )  +  4.05 ( 2 π r i λ ) 1 / 3 + 8 ;
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.