Skip to main content
Log in

Spatial Second-Order Positive and Asymptotic Preserving Unified Gas Kinetic Schemes for Radiative Transfer Equations

  • Published:
Journal of Scientific Computing Aims and scope Submit manuscript

Abstract

We propose two spatial second-order schemes for linear radiative transfer equations by using the idea of the unified gas kinetic scheme (UGKS) to construct the numerical boundary fluxes, and show that the proposed schemes are both positive and asymptotic preserving. The UGKS was proposed by Xu and Huang (J Comput Phys 229:7747–7764, 2010) for continuum and rarefied flows firstly, and was then applied to a linear radiative transfer equation by Mieussens in (J Comput Phys 253:138–156, 2013) where the asymptotic preserving property of UGKS is shown. Although it is asymptotic preserving, UGKS can not always keep the positivity of solutions. We first apply UGKS to discretize a linear radiative transfer equation to have a spatial second-order scheme. Then, by a detailed analysis of the numerical boundary fluxes, we are able to find the reasons why the positive preserving property of UGKS fails. Finally, we carefully employ a linear scaling limiter and a flux correction to make UGKS positive-preserving but still asymptotic-preserving. Consequently, we propose two spatial second-order positive and asymptotic preserving unified gas kinetic schemes for the linear radiative transfer equation, thus improving the earlier work (J Comput Phys 444:110546, 2021) where only a first-order positive and asymptotic scheme is developed. The proposed schemes can well capture the solution of the diffusion limit equation in optically thick regions without requiring the cell size being smaller than the photon’s mean free path, while the solution in optically thin regions can also be well resolved in a natural way. To our best knowledge, this is the first time that a spatial second-order positive and asymptotic preserving gas kinetic scheme for linear radiative transfer equations is constructed. Several numerical experiments are included to validate the spatial second-order accuracy, positive- and asymptotic-preserving properties of the proposed schemes.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

Data Availability

Enquiries about data availability should be directed to the authors.

References

  1. Azmy, Y., Sartori, E.: Nuclear Computational Science: A Century in Review. Springer Science & Business Media, Berlin (2010)

    Book  Google Scholar 

  2. Castro, R.O., Trelles, J.P.: Spatial and angular finite element method for radiative transfer in participating media. J. Quant. Spectrosc. Radiat. Transf. 157, 81–105 (2015)

    Article  Google Scholar 

  3. Chai, J.C., Patankar, S.V., Lee, H.S.: Evaluation of spatial differencing practices for the discrete-ordinates method. J. Thermophys. Heat Transf. 8, 140–144 (1994)

    Article  Google Scholar 

  4. Coelho, P.J.: Advances in the discrete ordinates and finite volume methods for the solution of radiative heat transfer problems in participating media. J. Quant. Spectrosc. Radiat. Transf. 145, 121–146 (2014)

    Article  Google Scholar 

  5. Gentile, N.A.: Implicit Monte Carlo diffusion - an acceleration method for Monte Carlo time-dependent radiative transfer simulations. J. Comput. Phys. 172, 543–571 (2001)

    Article  MATH  Google Scholar 

  6. Hu, J.W., Shu, R.W.: A second-order asymptotic-preserving and positivity-preserving exponential Runge-Kutta method for a class of stiff kinetic equations. Multiscale Model. Simul. 17, 1123–1146 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  7. Guermond, J.L., Popov, B., Ragusa, J.: Positive and asymptotic preserving approximation of the radiation transport equation. SIAM J. Numer. Anal. 58, 519–540 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  8. Jin, S.: Efficient asymptotic-preserving (AP) schemes for some multiscale kinetic equations. SIAM J. Sci. Comput. 21, 441–454 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  9. Jin, S., Pareschi, L., Toscani, G.: Uniformly accurate diffusive relaxation schemes for multiscale transport equations. SIAM J. Numer. Anal. 38, 913–936 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  10. Klar, A.: An asymptotic-induced scheme for nonstationary transport equations in the diffusive limit. SIAM J. Numer. Anal. 35, 1073–1094 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  11. Larsen, E.W., Morel, J.E., Miller, W.F., Jr.: Asymptotic solutions of numerical transport problems in optically thick, diffusive regimes. J. Comput. Phys. 69, 283–324 (1987)

    Article  MathSciNet  MATH  Google Scholar 

  12. Larsen, A.W., Morel, J.E.: Asymptotic solutions of numerical transport problems in optically thick, diffusive regimes. II. J. Comput. Phys. 83(1), 212–236 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  13. Laboure, V.M., McClarren, R.G., Hauck, C.D.: Implicit filtered \(P_N\) for high-energy density thermal radiation transport using discontinuous Galerkin finite elements. J. Comput. Phys. 321, 624–643 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  14. Lathrop, K.D., Carlson, B.G.: Discrete ordinates angular quadrature of the neutron transport equation, LA-3186, Los Alamos Scientific Laboratory (1965)

  15. Laiu, M.P., Hauck, C.D.: Positivity limiters for filtered spectral approximations of linear kinetic transport equations. J. Sci. Comput. 78(2), 918–950 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  16. Laiu, M.P., Frank, M., Hauck, C.D.: A positive asymptotic-preserving scheme for linear kinetic transport equations. SIAM J. Sci. Comput. 41(3), A1500–A1526 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  17. Laiu, M.P.: Positive filtered PN method for linear transport equations and the associated optimization algorithm. Phd dissertation (2016)

  18. Leck, J.F., Ummings, J.C.: An implicit Monte Carlo scheme for calculating time and frequency dependent nonlinear radiation transport. J. Comput. Phys. 8, 313–342 (1971)

    Article  MathSciNet  Google Scholar 

  19. Ling, D., Cheng, J., Shu, C.W.: Conservative high order positivity-preserving discontinuous Galerkin methods for linear hyperbolic and radiative transfer equations. J. Sci. Comput. 77, 1801–1831 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  20. Mieussens, L.: On the asymptotic preserving property of the unified gas kinetic scheme for the diffusion limit of linear kinetic models. J. Comput. Phys. 253, 138–156 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  21. Peng, Z.C., Cheng, Y.D., Qiu, J.M., Li, F.Y.: Stability-enhanced AP IMEX-LDG schemes for linear kinetic transport equations under a diffusive scaling. J. Comput. Phys. 415, 109485 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  22. Radice, D., Abdikamalov, E., Rezzolla, L., Ott, C.D.: A new spherical harmonics scheme for multi-dimensional radiation transport I. Static matter configurations. J. Comput. Phys. 242, 648–669 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  23. Wareing, T.A.: An exponential discontinuous scheme for discrete-ordinate calculations in Cartesian geometries. In: Proceedings of the Joint International Conference on Mathemat- ical Methods and Supercomputing for Nuclear Applications, Saratoga Springs, NY (1997)

  24. Sun, W.J., Jiang, S., Xu, K.: An asymptotic preserving unified gas kinetic scheme for gray radiative transfer equations. J. Comput. Phys. 285, 265–279 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  25. Sun, W.J., Jiang, S., Xu, K.: An asymptotic preserving unified gas kinetic scheme for frequency-dependent radiative transfer equations. J. Comput. Phys. 302, 222–238 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  26. Sun, W.J., Jiang, S., Xu, K.: An asymptotic preserving implicit unified gas kinetic scheme for frequency-dependent radiative transfer equations. Int. J. Numer. Anal. Model. 15, 134–153 (2018)

    MathSciNet  MATH  Google Scholar 

  27. Sun, W.J., Jiang, S., Xu, K.: An implicit unified gas kinetic scheme for radiative transfer with equilibrium and non-equilibrium diffusive limits. Commun. Comput. Phys. 22, 889–912 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  28. Sun, W.J., Jiang, S., Xu, K.: A multidimensional unified gas-kinetic scheme for radiative transfer equations on unstructured mesh. J. Comput. Phys. 351, 455–472 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  29. Yuan, D.M., Cheng, J., Shu, C.W.: High order positivity-preserving discontinuous Galerkin methods for radiative transfer equations. SIAM J. Sci. Comput. 38, A2987–A3019 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  30. Xu, K., Huang, J.C.: A unified gas-kinetic scheme for continuum and rarefied flows. J. Comput. Phys. 229, 7747–7764 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  31. Xu, K.: Direct Modeling for Computational Fluid Dynamics: Construction and Application of Unified Gas-kinetic Schemes. World Scientific Press, Singapore (2015)

    Book  MATH  Google Scholar 

  32. Xu, X.J., Sun, W.J., Jiang, S.: An asymptotic preserving angular finite element based unified gas kinetic scheme for gray radiative transfer equations. J. Quant. Spectrosc. Radiat. Transf. 243, 106808 (2020)

    Article  Google Scholar 

  33. Xu, X.J., Jiang, S., Sun, W.J.: A positive and asymptotic preserving filtered \(P_N\) method for the gray radiative transfer equations. J. Comput. Phys. 444, 110546 (2021)

    Article  MATH  Google Scholar 

  34. Zhang, X.X., Shu, C.W.: On maximum-principle-satisfying high order schemes for scalar conservation laws. J. Comput. Phys. 229, 3091–3120 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  35. Adams, M.L., Larsen, E.W.: Fast iterative methods for discrete-ordinates particle transport calculations. Prog. Nucl. Energy 40, 3–159 (2002)

    Article  Google Scholar 

Download references

Acknowledgements

The current research is supported by NSFC (Grant No. 12001451) for Xu, and by National Key R &D Program (2020YFA0712200), National Key Project (GJXM92579), and NSFC (Grant No. 11631008), the Sino-German Science Center (Grant No. GZ 1465) for Jiang, and by NSFC (Grant No. 12292981, 12292982) for Sun.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Wenjun Sun.

Ethics declarations

Conflict of interest

The authors have not disclosed any competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix

Appendix

1.1 Derivation of the Time Step Constraint (4.5)

Denoting

$$\begin{aligned} f(x) =(1+e^{-x})-\frac{2}{x}(1-e^{-x}), \end{aligned}$$

we see that the coefficient function \({\tilde{d}}(\Delta t, \epsilon , \sigma _s,\sigma _a)\) can be rewritten as

$$\begin{aligned} {\tilde{d}}(\Delta t, \epsilon , \sigma _s,\sigma _a) = -\frac{\sigma _s}{2\pi \epsilon ^4 \nu ^2}f( \nu \Delta t ), \end{aligned}$$

where we recall \( \nu = \frac{\sigma _s}{\epsilon ^2}+\sigma _a\). When \(\nu \) is small (i.e., both \(\sigma _s\) and \(\sigma _a\) are small), we have by the Taylor expansion that

$$\begin{aligned} f(\nu \Delta t) \le \frac{1}{6} (\nu \Delta t)^2. \end{aligned}$$

Therefore,

$$\begin{aligned} -{\tilde{d}}(\Delta t, \epsilon , \sigma _s,\sigma _a) \le \frac{\sigma _s}{12\pi \epsilon ^4}(\Delta t)^2. \end{aligned}$$

Thus, for the optically thin regime, in order to make (4.4) hold, we can approximately require \(\Delta t\) to satisfy

$$\begin{aligned} \max _{i,j} \left( \frac{\sigma _{s,i\pm 1/2,j\pm 1/2}(\Delta t)^2 }{12\pi \epsilon ^4 \sigma _{s,i,j}} \right) \le \frac{ (\Delta x \Delta y)^2 }{ 4\pi \epsilon ^2 \left( (\Delta x)^2+(\Delta y)^2 \right) }, \end{aligned}$$

that is,

$$\begin{aligned} \Delta t \le \sqrt{ \min _{i,j} \frac{\sigma _{s,i,j}}{\sigma _{s,i\pm 1/2,j\pm 1/2}} \frac{ 3\epsilon ^2 (\Delta x \Delta y)^2}{ (\Delta x)^2+(\Delta y)^2}} \le \epsilon \min _{i,j} \sqrt{ \frac{\sigma _{s,i,j}}{\sigma _{s,i\pm 1/2,j\pm 1/2}} } \cdot \frac{\sqrt{6}}{2} \min \{\Delta x, \Delta y \}. \end{aligned}$$

The time step constraint (4.5) is now verified.

1.2 Truncation Error Estimation of the UGKS

This subsection is devoted to estimating the truncation errors of the original spatial second-order UGKS. For simplicity, we make the following assumption:

Assumption 8.1

We assume that the discretization of the angular variable is exact, that is, the angular error is ignored. In addition, we assume that \(\sigma _s\), \(\sigma _a\) are constants, and that the solution of (2.4) is sufficiently smooth.

In the following, for convenience, we add subscript \('{_h}'\) to the previous numerical solution and fluxes so as to distinguish them from the exact ones. Therefore, the previous discretized micro and macro Eqs. (3.17)–(3.16) can be rewritten as follows:

$$\begin{aligned} I^{n+1}_{i,j,m,h}= & {} I^{n}_{i,j,m,h}+\frac{\Delta t}{\Delta x}\left( F_{i-1/2,j,m,h}-F_{i+1/2,j,m,h}\right) +\frac{\Delta t}{\Delta y}\left( H_{i,j-1/2,m,h}-H_{i,j+1/2,m,h}\right) \nonumber \\{} & {} + \Delta t \frac{\sigma _{s}}{\epsilon ^2}\left( \frac{1}{2\pi }\rho ^{n+1}_{i,j,h} - I^{n+1}_{i,j,m,h}\right) -\sigma _{a}\Delta t I^{n+1}_{i,j,m,h}, \end{aligned}$$
(8.1)
$$\begin{aligned} \rho ^{n+1}_{i,j,h}= & {} \rho ^{n}_{i,j,h}+\frac{\Delta t}{\Delta x} \left( \Phi ^{n+1}_{i-1/2,j,h}-\Phi ^{n+1}_{i+1/2,j,h}\right) +\frac{\Delta t}{\Delta y} \left( \Psi ^{n+1}_{i,j-1/2,h}-\Psi ^{n+1}_{i,j+1/2,h}\right) \nonumber \\{} & {} \quad -\sigma _{a} \Delta t\rho ^{n+1}_{i,j,h}. \end{aligned}$$
(8.2)

In order to estimate the truncation error, we first derive an alternative form of the equation (2.4), from which we can have a more intuitive understanding on the construction of the UGKS scheme. More importantly, the truncation error estimate will be based on the alternative form. For this, we integrate the Eq. (2.4) over a spatial cell (ij) and time interval \([t_n, t_{n+1}]\), and also divide it by \(\Delta x\Delta y\). For the term \(\partial _x I\), it holds that

$$\begin{aligned}{} & {} \frac{1}{\Delta x\Delta y }\int _{t_n}^{t_{n+1}}\int _{y_{j-\frac{1}{2}}}^{y_{j+\frac{1}{2}}}\int _{x_{i-\frac{1}{2}}}^{x_{i+\frac{1}{2}}} \partial _xI(x,y,{\varvec{\Omega }},t) dxdydt\\{} & {} \quad =\frac{1}{\Delta x\Delta y}\int _{t_n}^{t_{n+1}} \int _{y_{j-\frac{1}{2}}}^{y_{j+\frac{1}{2}}} \int _{x_{i-\frac{1}{2}}}^{x_{i+\frac{1}{2}}} \partial _x I(x,y_j,{\varvec{\Omega }},t) +(y-y_j)\partial ^2_{xy}I(x,y_j,{\varvec{\Omega }},t) +O((\Delta y)^2) dxdydt\\{} & {} \quad = \frac{1}{\Delta x}\left( \int _{t_n}^{t_{n+1}} I(x_{i+1/2},y_j,{\varvec{\Omega }},t)dt- \int _{t_n}^{t_{n+1}}I(x_{i-1/2},y_j,{\varvec{\Omega }},t)dt \right) + O\left( (\Delta y)^2\Delta t\right) . \end{aligned}$$

And the term \( \partial _yI\) can be handled similarly. Hence, we conclude that

$$\begin{aligned} I^{n+1}_{i,j}= & {} I^{n}_{i,j}+\frac{\xi }{\epsilon \Delta x} \int _{t_n}^{t_{n+1}} I(x_{i-1/2},y_j,{\varvec{\Omega }},t) - I(x_{i+1/2},y_j,{\varvec{\Omega }},t) dt \nonumber \\{} & {} + \frac{\eta }{\epsilon \Delta y} \int _{t_n}^{t_{n+1}} I(x_i,y_{j-1/2},{\varvec{\Omega }},t)dt- I(x_i,y_{j+1/2},{\varvec{\Omega }},t)dt\nonumber \\{} & {} + \Delta t \frac{\sigma _s}{2\pi \epsilon ^2} \breve{\rho }_{i,j} -\Delta t \nu \breve{I}_{i,j} + O\left( \frac{\eta }{\epsilon } (\Delta x)^2\Delta t + \frac{\xi }{\epsilon }(\Delta y)^2 \Delta t\right) , \end{aligned}$$
(8.3)

where \(\breve{\rho }, \breve{I}\) are defined in the same way as in (3.6).

Now, in order to determine \(I(x_{i\pm 1/2},y_j,{\varvec{\Omega }},t)\), we get back to (2.4). Take \(y=y_j\) in (2.4), then the resulting equation can be written as

$$\begin{aligned} \frac{d}{dt} \left( e^{\nu (t-t_n)}I\left( x+\frac{\xi }{\epsilon }t, y_j, {\varvec{\Omega }}, t\right) \right) = e^{\nu (t-t_n)}\left( \frac{\sigma _s }{2\pi \epsilon ^2} \rho \left( x+\frac{\xi }{\epsilon } t, y_j, t\right) -\frac{\eta }{\epsilon }\partial _y I\left( x+\frac{\xi }{\epsilon } t, y_j, {\varvec{\Omega }}, t\right) \right) . \end{aligned}$$

Integrating the above equation from \(t_n\) to t for some \(t\in [t_n,t_{n+1}]\), and replacing \(x+\frac{\xi }{\epsilon } t\) by x, we get

(8.4)

with

(8.5)

Letting \(x=x_{i\pm 1/2}\) in (8.5) and utilizing the differential mean value theorem, we find that there is a \(\zeta _x\in (x_{i-1/2}, x_{i+1/2})\), such that

$$\begin{aligned}{} & {} \int _{t_n}^{t_{n+1}}\frac{\eta }{\epsilon } \int _{t_n}^te^{-\nu (t-s)}\left( \partial _y I \left( x_{i+1/2}-\frac{\xi }{\epsilon } (t-s),y_j, {\varvec{\Omega }}, s\right) \right. \\{} & {} \qquad - \left. \partial _y I \left( x_{i-1/2}-\frac{\xi }{\epsilon } (t-s), y_j, {\varvec{\Omega }}, s\right) \right) dsdt \\{} & {} \quad = \frac{\eta }{\epsilon } \int _{t_n}^{t_{n+1}} \int _{t_n}^te^{-\nu (t-s)} \Delta x \partial ^2_{xy} I \left( \zeta _x-\frac{\xi }{\epsilon } (t-s), y_j, {\varvec{\Omega }}, s\right) dsdt\\{} & {} \quad = O(\eta ~{\tilde{g}}~ \Delta x), \end{aligned}$$

where

$$\begin{aligned} {\tilde{g}}=\frac{1}{\epsilon \nu }( \frac{1}{\nu }e^{-\nu \Delta t}+\Delta t -\frac{1}{\nu } ). \end{aligned}$$

Thus, one finds the following relation

(8.6)

In the same manner, the relation for \(I(x_i,y_{j\pm 1/2},{\varvec{\Omega }},t)\) and can be obtained. Substituting (8.6) into (8.3), we obtain the desired form, which we summarize in the following lemma.

Lemma 8.1

Under Assumptions 8.1, we have the following alternative form of the Eq. (2.4) in a spatial cell (ij) and time interval \([t_n, t_{n+1}]\):

$$\begin{aligned} I^{n+1}_{i,j}= & {} I^{n}_{i,j}+\frac{\Delta t}{\Delta x} (F_{i-1/2,j}-F_{i+1/2,j})+ \frac{\Delta t}{\Delta y}(H_{i,j-1/2}-H_{i,j+1/2})\nonumber \\{} & {} + \Delta t \frac{\sigma _s}{2\pi \epsilon ^2} \breve{\rho }_{i,j} -\Delta t \nu \breve{I}_{i,j} + O\left( \frac{\eta }{\epsilon } (\Delta x)^2\Delta t + \frac{\xi }{\epsilon }(\Delta y)^2 \Delta t + \frac{\xi \eta }{\epsilon } {\tilde{g}} \right) , \end{aligned}$$
(8.7)

where

(8.8)

Integrating (8.7) with respect to the angular variable, we obtain the macroscopic equation:

$$\begin{aligned} \rho ^{n+1}_{i,j}= & {} \rho ^{n}_{i,j}+\frac{\Delta t}{\Delta x} (\Phi _{i-1/2,j}^{n+1}-\Phi _{i+1/2,j}^{n+1})+ \frac{\Delta t}{\Delta y}(\Psi _{i,j-1/2}^{n+1}-\Psi _{i,j+1/2}^{n+1})\nonumber \\{} & {} -\Delta t \sigma _{a,i,j} \breve{\rho }_{i,j} + O\left( (\Delta x)^2\Delta t + (\Delta y)^2 \Delta t + {\tilde{g}} \right) , \end{aligned}$$
(8.9)

where

(8.10)

In view of (8.1), (8.2), and (8.7) and (8.9), we conclude that the original spatial second-order UGKS is obtained immediately by taking the discrete angular directions \(\{{\varvec{\Omega }}_m = (\xi _m,\eta _m),\; m=1,2,\ldots ,M\}\), and constructing the approximation to \(\breve{\rho }\) and \(\breve{I}\), and also taking the piecewise linear reconstruction for both \( I(\ldots ,t_n)\) and \(\rho \) in (8.5).

With (8.7) and (8.9) in hand, we are able to derive the truncation error. Denoting by \(\tau _I, \tau _{\rho }\) the truncation errors of the scheme (8.1), (8.2), i.e., the remainders obtained by inserting the exact solution I and \(\rho \) into (8.1) and (8.2), we get

$$\begin{aligned}{} & {} \frac{ I^{n+1}_{i,j,m} - I^{n}_{i,j,m} }{\Delta t} + \frac{F_{i+1/2,j,m,h}(I,\rho )-F_{i-1/2,j,m,h}(I,\rho )}{\Delta x}\nonumber \\{} & {} \qquad +\frac{H_{i,j+1/2,m,h}(I,\rho )-H_{i,j-1/2,m,h}(I,\rho )}{\Delta y}\nonumber \\{} & {} \quad =\frac{\sigma _s}{2\pi \epsilon ^2} \rho _{i,j}^{n+1} - \nu I_{i,j,m}^{n+1} + \frac{\tau _I}{\epsilon ^2}, \end{aligned}$$
(8.11)
$$\begin{aligned}{} & {} \frac{ \rho ^{n+1}_{i,j} - \rho ^{n}_{i,j} }{\Delta t} + \frac{\Phi _{i+1/2,j,h}(I,\rho )^{n+1}-\Phi _{i-1/2,j,h}^{n+1}(I,\rho )}{\Delta x}\nonumber \\{} & {} \qquad +\frac{\Psi _{i,j+1/2,h}^{n+1} (I,\rho )-\Psi _{i,j-1/2,h}^{n+1}(I,\rho )}{\Delta y} \nonumber \\{} & {} \quad = -\sigma _a \rho _{i,j}^{n+1} + \tau _{\rho }, \end{aligned}$$
(8.12)

where we have used the notations \(F_{i+1/2,j,m,h}(I,\rho )\), \( H_{i,j+1/2,m,h}(I,\rho )\) and \(\Phi _{i+1/2,j,h}^{n+1}(I,\rho )\), \(\Psi _{i,j+1/2,h}^{n+1}(I,\rho )\) to highlight that the numerical solutions \(I_h,\rho _h\) in the numerical fluxes are replaced by the exact solutions \(I, \rho \). Then we can prove the following estimates of these truncation errors:

Theorem 8.1

Under Assumptions 8.1, then for the regime \(\epsilon =O(1)\) and the regime \(\epsilon \ll 1\), it holds that

$$\begin{aligned} \big | \tau _I \big | \le C \left( \Delta t+ (\Delta x)^2+(\Delta y)^2\right) ,\qquad \quad \big | \tau _{\rho } \big | \le C \left( \Delta t+ (\Delta x)^2+(\Delta y)^2\right) , \end{aligned}$$
(8.13)

where \(C\ge 0\) is a constant independent of \(\epsilon \), \(\Delta x\), \(\Delta y\) and \(\Delta t\).

Proof

Taking the angular direction \(\varvec{\Omega }_m\) in (8.7), and then subtracting (8.11) from (8.7), one has

$$\begin{aligned} \frac{\tau _I}{\epsilon ^2}= & {} \frac{ \left( F_{i+1/2,j,m} -F_{i+1/2,j,m,h}(I,\rho )\right) - \left( F_{i-1/2,j,m} -F_{i-1/2,j,m,h}(I,\rho )\right) }{\Delta x} \nonumber \\{} & {} + \frac{ \left( H_{i,j+1/2,m} -H_{i,j+1/2,m,h}(I,\rho )\right) - \left( H_{i,j-1/2,m} -H_{i,j-1/2,m,h}(I,\rho )\right) }{\Delta y} \nonumber \\{} & {} - \frac{\sigma _s}{2\pi \epsilon ^2}\left( \breve{\rho }_{i,j}-\rho _{i,j}^{n+1}\right) +\nu \left( \breve{I}_{i,j,m} - I_{i,j,m}^{n+1}\right) +O\left( \frac{\eta }{\epsilon } (\Delta x)^2 + \frac{\xi }{\epsilon }(\Delta y)^2 +\frac{\xi \eta }{\epsilon } \frac{{\tilde{g}}}{\Delta t} \right) .\nonumber \\ \end{aligned}$$
(8.14)

And also subtracting (8.12) from (8.9), we see that

$$\begin{aligned} \tau _{\rho }= & {} \frac{ \left( \Phi _{i+1/2,j}^{n+1}-\Phi _{i+1/2,j,h}^{n+1}(I,\rho )\right) - \left( \Phi _{i-1/2,j}^{n+1}-\Phi _{i-1/2,j,h}^{n+1}(I,\rho )\right) }{\Delta x} \nonumber \\{} & {} + \frac{ \left( \Psi _{i,j+1/2}^{n+1}-\Psi _{i,j+1/2,h}^{n+1}(I,\rho )\right) - \left( \Psi _{i,j-1/2}^{n+1} - \Psi _{i,j-1/2,h}^{n+1}(I,\rho )\right) }{\Delta y} \nonumber \\{} & {} +\sigma _a\left( \breve{\rho }_{i,j} - \rho _{i,j}^{n+1}\right) +O\left( (\Delta x)^2 +(\Delta y)^2 + \frac{{\tilde{g}}}{\Delta t} \right) .\nonumber \\ \end{aligned}$$
(8.15)

Recalling the definition of \(\breve{\rho }\) and \(\breve{I}\), and by employing the Taylor expansion, we infer that

$$\begin{aligned} \breve{\rho }_{i,j}- \rho _{i,j}^{n+1} = O(\Delta t + (\Delta x)^2 +(\Delta y)^2 ),\qquad \breve{I}_{i,j,m}- I_{i,j,m}^{n+1} = O(\Delta t + (\Delta x)^2 +(\Delta y)^2). \end{aligned}$$
(8.16)

Since under Assumptions 8.1, the macro fluxes are exact integrals of the corresponding micro fluxes with respect to the angular variable. Hence, we shall focus on the terms of the fluxes in (8.14). For the first term in (8.14), we consider the case \(\xi _m<0\) only, while \(\xi _m>0\) can be handled similarly. Keeping in mind that \(F_{i-1/2,j,m}\) is defined in (8.8) with angular direction taken as \(\varvec{\Omega }_m\). We take Taylor’s expansions to \(I\left( x_{i-1/2}-\frac{\xi _m}{\epsilon }(t-t_n),y_j, {\varvec{\Omega }}, t_n \right) \) at \(x=x_i\), and \(\rho \left( x_{i-1/2}-\frac{\xi _m}{\epsilon }(t-s),y_j,s\right) \) at \(x=x_{i-1/2}, s=t_{n+1}\) in (8.5), and then substitute (8.5) into (8.8). After a straightforward calculation, we obtain

(8.17)

where

$$\begin{aligned} \breve{F}_{i-1/2,j,m}= & {} {\tilde{\alpha }}~ \xi _m \left( I^n_{i,j,m}-\frac{\Delta x}{2}\left( \frac{\partial I}{\partial x}\right) _{i,j,m}^n\right) +{\tilde{b}}~\xi _m^2 \left( \frac{\partial I}{\partial x}\right) _{i,j,m}^n +{\tilde{c}}~\xi _m\rho _{i-1/2,j}^{n+1} + {\tilde{d}}~\xi _m^2 ~ \left( \frac{\partial \rho }{\partial x}\right) _{i,j}^{n+1}, \nonumber \\ R_{i-1/2,j,m}= & {} {\tilde{e}}\left( \frac{\partial ^2 I}{\partial x^2}\right) ^n_{i,j,m} + {\tilde{f}}\left( \frac{\partial \rho }{\partial t}\right) _{i-1/2,j}^{n+1} +{\tilde{h}}\left( \frac{\partial ^2 \rho }{\partial x^2}\right) _{i-1/2,j}^{n+1} + {\tilde{\beta }}\left( \frac{\partial ^2 \rho }{\partial x \partial t}\right) _{i-1/2,j}^{n+1} \nonumber \\{} & {} + {\tilde{\gamma }}\left( \frac{\partial ^2 \rho }{\partial t^2}\right) _{i-1/2,j}^{n+1}+ higher\text {-}order~terms, \end{aligned}$$
(8.18)

with

$$\begin{aligned} {\widetilde{e}}= & {} -\frac{\xi _m}{2\epsilon \nu \Delta t} \left\{ \left( \frac{\Delta x}{2}+\frac{\xi _m}{\epsilon }\Delta t\right) ^2 e^{-\nu \Delta t}- \frac{(\Delta x)^2}{4} + \frac{2\xi _m}{\epsilon \nu } \left( \frac{\Delta x}{2}+\frac{\xi _m}{\epsilon }\Delta t\right) e^{-\nu \Delta t} \right. \\{} & {} \left. - \frac{\xi _m\Delta x}{\epsilon \nu } +\frac{2\xi _m^2}{\epsilon ^2 \nu ^2} e^{-\nu \Delta t} - \frac{2\xi _m^2}{\epsilon ^2 \nu ^2} \right\} ,\\ {\tilde{f}}= & {} \frac{\xi _m\sigma _s }{2\pi \epsilon ^3\Delta t} \left( \frac{1}{\nu ^2} \left( \Delta t + \frac{1}{\nu }\right) (1- e^{-\nu \Delta t}) -\frac{1}{2\nu }(\Delta t)^2 - \frac{1}{\nu ^2}\Delta t \right) , \\ {\tilde{h}}= & {} \frac{\xi _m^3\sigma _s }{4\pi \epsilon ^5\Delta t} \left\{ \frac{1}{\nu ^2} (\Delta t)^2 e^{-\nu \Delta t}+ \frac{4}{\nu ^3}\Delta te^{-\nu \Delta t}+ \frac{6}{\nu ^4} e^{-\nu \Delta t} + \frac{2}{\nu ^3}\Delta t - \frac{6}{\nu ^4} \right\} ,\\ {\tilde{\beta }}= & {} \frac{\xi _m^2\sigma _s }{2\pi \epsilon ^4\Delta t} \left\{ \frac{1}{\nu ^2} (\Delta t)^2e^{-\nu \Delta t} + \frac{3}{\nu ^3}\Delta t e^{-\nu \Delta t} + \frac{3}{\nu ^4} e^{-\nu \Delta t} + \frac{(\Delta t)^2}{2\nu ^2} - \frac{3}{\nu ^4} \right\} ,\\ {\widetilde{\gamma }}= & {} \frac{\xi _m\sigma _s }{4\pi \epsilon ^3\Delta t} \left\{ \frac{(\Delta t)^3}{3\nu } + \frac{(\Delta t)^2}{\nu ^2} e^{-\nu \Delta t} +\frac{2}{\nu ^3} \Delta t e^{-\nu \Delta t}-\frac{2}{\nu ^4} + \frac{2}{\nu ^4}e^{-\nu \Delta t} \right\} . \end{aligned}$$

Based on (8.17), we make the following decomposition

$$\begin{aligned}{} & {} \frac{ \left( F_{i+1/2,j,m} -F_{i+1/2,j,m,h}(I,\rho )\right) - \left( F_{i-1/2,j,m} -F_{i-1/2,j,m,h}(I,\rho )\right) }{\Delta x}\nonumber \\{} & {} \quad = \frac{ \left( \breve{F}_{i+1/2,j,m} -F_{i+1/2,j,m,h}(I,\rho )\right) - \left( \breve{F}_{i-1/2,j,m} -F_{i-1/2,j,m,h}(I,\rho )\right) }{\Delta x} \nonumber \\{} & {} \qquad + \frac{ R_{i+1/2,j,m} - R_{i-1/2,j,m} }{\Delta x},\nonumber \\ \end{aligned}$$
(8.19)

where we further have

$$\begin{aligned} \breve{F}_{i-1/2,j,m}-F_{i-1/2,j,m,h}(I,\rho )= & {} \left( -{\tilde{\alpha }}~ \xi _m~\frac{\Delta x}{2} + {\tilde{b}}~\xi _m^2\right) ~ \left( \left( \frac{\partial I}{\partial x}\right) _{i,j,m}^n - \delta _x I_{i,j,m}^n \right) \\{} & {} +{\tilde{c}}~\xi _m \left( \rho _{i-1/2,j}^{n+1}-{\overline{\rho }}_{i-1/2,j}^{n+1} \right) + {\tilde{d}}~\xi _m^2 ~ \left( \left( \frac{\partial \rho }{\partial x}\right) _{i-1/2,j}^{n+1}-\delta _x \rho _{i-1/2,j}^{n+1} \right) , \end{aligned}$$

with \( {\overline{\rho }}_{i-1/2,j}^{n+1} = \frac{\rho ^{n+1}_{i-1,j}+\rho ^{n+1}_{i,j} }{2}\).

Since the solutions are smooth enough, we find that

$$\begin{aligned} \left( \frac{\partial I}{\partial x}\right) _{i,j,m}^n - \delta _x I_{i,j,m}^n= & {} -\frac{1}{3!} \left( \frac{\partial ^3 I}{\partial x^3} \right) _{i,j,m}^n (\Delta x)^2 + O( (\Delta x)^4), \\ \rho _{i-1/2,j}^{n+1}-{\overline{\rho }}_{i-1/2,j}^{n+1}= & {} -\frac{1}{2}\left( \frac{\partial ^2 \rho }{\partial x^2}\right) _{i-1/2,j}^{n+1}\left( \frac{\Delta x}{2}\right) ^2 + O\left( (\Delta x)^4\right) , \\ \left( \frac{\partial \rho }{\partial x}\right) _{i-1/2,j}^{n+1}-\delta _x \rho _{i-1/2,j}^{n+1}= & {} -\frac{1}{3!}\left( \frac{\partial ^3 \rho }{\partial x^3}\right) _{i-1/2,j}^{n+1}\left( \frac{\Delta x}{2}\right) ^2 + O\left( (\Delta x)^4 \right) . \end{aligned}$$

Thus, the first term on the right-hand side of (8.19) can be written as follows.

$$\begin{aligned}{} & {} \frac{ \left( \breve{F}_{i+1/2,j,m} -F_{i+1/2,j,m,h}(I,\rho )\right) - \left( \breve{F}_{i-1/2,j,m} -F_{i-1/2,j,m,h}(I,\rho )\right) }{\Delta x} \nonumber \\{} & {} \quad = \left( -{\tilde{\alpha }}~ \xi _m~\frac{\Delta x}{2} + {\tilde{b}}~\xi _m^2\right) ~ \left( \frac{1}{3!} \left( \left( \frac{\partial ^3 I}{\partial x^3} \right) _{i,j,m}^n - \left( \frac{\partial ^3 I}{\partial x^3} \right) _{i+1,j,m}^n \right) \Delta x + O( (\Delta x)^3) \right) \nonumber \\{} & {} \qquad +{\tilde{c}}~\xi _m \left( \frac{1}{2} \left( \left( \frac{\partial ^2 \rho }{\partial x^2}\right) _{i-1/2,j}^{n+1}- \left( \frac{\partial ^2 \rho }{\partial x^2}\right) _{i+1/2,j}^{n+1}\right) \frac{\Delta x}{4} + O\left( (\Delta x)^3\right) \right) \nonumber \\{} & {} \qquad + {\tilde{d}}~\xi _m^2 ~ \left( \frac{1}{3!}\left( \left( \frac{\partial ^3 \rho }{\partial x^3}\right) _{i-1/2,j}^{n+1}- \left( \frac{\partial ^3 \rho }{\partial x^3}\right) _{i+1/2,j}^{n+1} \right) \frac{\Delta x}{4} + O\left( (\Delta x)^3 \right) \right) \nonumber \\{} & {} \quad = \left( -{\tilde{\alpha }}~ \xi _m~\frac{\Delta x}{2} + {\tilde{b}}~\xi _m^2\right) ~ O\left( (\Delta x)^2 \right) + {\tilde{c}}~\xi _m~ O\left( (\Delta x)^2 \right) + {\tilde{d}}~\xi _m^2 ~O\left( (\Delta x)^2\right) , \end{aligned}$$
(8.20)

where the differential mean value theorem is used. As for the second term on the right-hand side of (8.19), we use the differential mean value theorem again to obtain

$$\begin{aligned} \frac{ R_{i+1/2,j,m} - R_{i-1/2,j,m} }{\Delta x}= & {} {\tilde{e}}~O(1) +{\tilde{f}}~\xi _m~O(1) +{\tilde{h}}~\xi _m^3~O(1) +{\tilde{\beta }}~\xi _m^2~O(1)\nonumber \\{} & {} + {\tilde{\gamma }}~\xi _m~O(1) + higher\text {-}order~terms. \end{aligned}$$
(8.21)

We remark that here the “higher-order terms” can be handled in the same manner, and we shall not repeat to mention it in what follows.

For the regime \(\epsilon = O(1)\), we take the Taylor expansion to \(e^{-\nu \Delta t}\) in the coefficients to obtain

$$\begin{aligned}{} & {} {\tilde{\alpha }}=O(1), \qquad {\tilde{b}}=O\left( \Delta t\right) , \qquad {\tilde{c}} = O(\Delta t),\qquad {\tilde{d}} = O\left( (\Delta t)^2)\right) ,\quad {\tilde{e}} = O\left( (\Delta x)^2 + \Delta x\Delta t)\right) ,\qquad \\{} & {} \qquad {\tilde{f}} = O\left( (\Delta t)^2\right) ,\qquad {\tilde{h}} = O\left( (\Delta t)^3\right) , \qquad {\tilde{\gamma }} = O\left( (\Delta t)^3\right) , \qquad {\tilde{\beta }} = O\left( (\Delta t)^3\right) . \end{aligned}$$

Hence, combining (8.19) and (8.20) with (8.21), we conclude that

$$\begin{aligned} \Big |\frac{ \left( F_{i+1/2,j,m} -F_{i+1/2,j,m,h}(I,\rho )\right) - \left( F_{i-1/2,j,m} -F_{i-1/2,j,m,h}(I,\rho )\right) }{\Delta x} \Big | \le C\left( (\Delta t)^2 +(\Delta x)^2 \right) .\nonumber \\ \end{aligned}$$
(8.22)

Similarly, one can obtain

$$\begin{aligned} \Big |\frac{ \left( H_{i,j+1/2,m} -H_{i,j+1/2,m,h}(I,\rho )\right) - \left( H_{i,j-1/2,m} -H_{i,j-1/2,m,h}(I,\rho )\right) }{\Delta y} \Big | \le C\left( (\Delta t)^2 +(\Delta y)^2\right) .\nonumber \\ \end{aligned}$$
(8.23)

Substituting (8.16), (8.22) and (8.23) into (8.14), and keeping in mind that \({\tilde{g}} = O\left( (\Delta t)^2\right) \) in the regime \(\epsilon =O(1)\), we deduce that \(\big | \tau _I \big |\le C\left( \Delta t+ (\Delta x)^2+(\Delta y)^2\right) \). And the estimate \( \big |\tau _{\rho }\big | \le C\left( \Delta t+ (\Delta x)^2+(\Delta y)^2\right) \) follows immediately.

On the other hand, for the regime \(\epsilon \ll 1\), we can verify that

$$\begin{aligned}{} & {} {\tilde{\alpha }}=O(\epsilon ), \qquad {\tilde{b}}=O(\epsilon ^2), \qquad {\tilde{c}}=O(\frac{1}{\epsilon }),\qquad {\tilde{d}}=O(1),\qquad {\tilde{e}}=O(\epsilon ),\qquad \nonumber \\{} & {} \quad {\tilde{f}}=O(\frac{1}{\epsilon }),\qquad {\tilde{h}} = O\left( \epsilon \right) , \qquad {\tilde{\gamma }} = O\left( \frac{1}{\epsilon }\right) , \qquad {\tilde{\beta }} = O\left( \Delta t\right) . \end{aligned}$$
(8.24)

Thus,

$$\begin{aligned} \epsilon ^2 \Big |\frac{ \left( F_{i+1/2,j,m} -F_{i+1/2,j,m,h}(I,\rho )\right) - \left( F_{i-1/2,j,m} -F_{i-1/2,j,m,h}(I,\rho )\right) }{\Delta x} \Big | \le C\left( \Delta t +(\Delta x)^2\right) , \end{aligned}$$

where we have used the fact that \(\epsilon < \Delta t\) in the regime \(\epsilon \ll 1\). The estimate for the \(y-\)direction flux can be obtained in the same manner. Therefore, we have \(\big | \tau _I \big | \le C\left( \Delta t+ (\Delta x)^2+(\Delta y)^2\right) \).

Now, it remains to prove the estimate for \(\tau _{\rho }\). Using (3.1), (8.20) and (8.21), one gets

$$\begin{aligned}{} & {} \frac{ \left( \Phi _{i+1/2,j}^{n+1}-\Phi _{i+1/2,j,h}^{n+1}(I,\rho )\right) - \left( \Phi _{i-1/2,j}^{n+1}-\Phi _{i-1/2,j,h}^{n+1}(I,\rho )\right) }{\Delta x}\\{} & {} \quad = \left( -{\tilde{\alpha }}~\frac{\Delta x}{2} + {\tilde{b}} \right) ~ O\left( (\Delta x)^2 \right) + {\tilde{d}} ~O\left( (\Delta x)^2\right) + {\tilde{e}}~O(1) +{\tilde{\beta }}~O(1) + higher\text {-}order~terms. \end{aligned}$$

In view of (8.24) and the fact that \({\tilde{g}}=O(\epsilon \Delta t)\) in the regime \(\epsilon \ll 1\), we conclude \( \big |\tau _{\rho }\big | \le C\left( \Delta t+ (\Delta x)^2+(\Delta y)^2\right) \). This completes the proof. \(\square \)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Xu, X., Jiang, S. & Sun, W. Spatial Second-Order Positive and Asymptotic Preserving Unified Gas Kinetic Schemes for Radiative Transfer Equations. J Sci Comput 96, 94 (2023). https://doi.org/10.1007/s10915-023-02305-3

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10915-023-02305-3

Keywords

Mathematics Subject Classification

Navigation