1 Introduction
Significant advances in high-performance computing clusters (HPCs) have enabled three-dimensional (3D) whole-core high-fidelity heterogeneous transport simulations at the subpin level. This is becoming increasingly important for improving the reactor performance and safety as well as providing an efficient tool for more complex and novel designs. A numerical simulation method for solving the neutron transport equation, i.e., the method of characteristics (MOC), has garnered significant attention in whole-core high-fidelity simulations as the MOC can be applied to obtain detailed flux distributions without any restrictions on the geometry mesh. Hence, efficient performances can be achieved on various HPC systems. In fact, many famous high-fidelity neutron transport simulation codes were developed based on the MOC and have been used successfully in practical reactor applications, such as the CRX code [1], DeCART code [2,3], nTRACER [4], MPACT [5], NECP-X [6], and STREAM [7].
However, the computational feasibility of whole-core MOC calculation depends on the acceleration technique used. The neutron transport equation contains seven independent variables associated with space, angle, energy, and time. The key to accelerating high-fidelity transport calculations is to establish a low-order equivalent equation relevant to the space, angle, energy, or time while minimizing the computational cost. However, previous studies focused more on space and energy accelerations. An equivalent low-order equation relevant to a coarse mesh or a coarse energy interval has been established to accelerate high-order transport calculations. The coarse mesh finite difference (CMFD) method is a popular method, and different types of CMFD methods [8-15] are widely utilized to accelerate MOC calculations. Furthermore, researchers have proposed a transient multilevel (TML) method to reduce the number of high-order transport calculations significantly [16,17]. Using the TML method, acceleration was achieved on the time scale; in this regard, large time steps were used for the MOC calculations, whereas small time steps were used for the low-order CMFD and point kinetics equations.
In the high-fidelity core simulators established, the pin-homogenized multigroup CMFD (MG CMFD) method is widely used to accelerate the convergence of pin-resolved whole-core transport calculations. Excellent acceleration has been achieved. However, the MG CMFD method only updates the scalar flux and eigenvalue, and a bottleneck is encountered. Therefore, the MOC iterations will be reduced. Subsequently, the number of iterations cannot continue to decrease if only the MG CMFD is used for acceleration. In this case, the MOC calculations still incur a significant computational burden. Therefore, whole core pin-resolved MOC calculations require a more efficient acceleration technique. Fourier analysis [18,19] is a standard technique regularly used to investigate the stability of CMFD acceleration. Based on the Fourier analysis, the MG CMFD which based on flat source region or finer mesh can be used to further reduce the number of MOC iterations. Another potential acceleration technique is to directly update the incident angular flux and sources to further reduce the iterations of MOC calculations. Hence, an equivalent angular flux nonlinear finite difference (ANFD) equation must be established and applied to accelerate the MOC calculations. Furthermore, the domain decomposition technique is typically utilized for parallel computations, and the reactor core can be categorized into subdomains. Subsequently, the incident angular flux of each domain can be updated in time.
However, the angular flux acceleration technique has not yet been investigated; hence, we will attempt to establish an equivalent ANFD equation in this study. The even parity discrete ordinates method can be used to derive a low-order angular flux nonlinear finite difference equation. However, ensuring the equivalence between this low-order equation and the high-order pin-resolved transport equation results in new challenges. The introduction of discontinuity factors used in popular CMFD methods is a good concept. Through the strict definition and calculation of even parity and odd parity discontinuity factors based on flux information obtained from MOC calculations, a completely equivalent angular flux nonlinear finite difference equation can be established. As relevant studies have not yet been performed, more details regarding the angular flux nonlinear finite difference equation based on the even parity discrete ordinates method, as well as the definition and calculation scheme of even and odd parity discontinuity factors including details for addressing different boundary conditions are discussed herein. Furthermore, an equivalent angular flux nonlinear finite difference equation with the pin-resolved transport equation was established in this study.
This paper is organized as follows. In Sect. 2, we provide the necessary theory for establishing the equivalent low-order angular flux nonlinear finite difference equation. The numerical method to solve this equation is discussed in Sect. 3. A code based on this new method is developed, and two-dimensional (2D) pin, 2D assembly, and 2D C5G7 benchmark problems [20] are selected to verify the equivalence of the angular flux nonlinear finite difference method. The numerical results are discussed in Sect. 4, and important conclusions are provided. Sect. 5 presents the summary and conclusions.
2 Methodology of angular flux nonlinear finite difference equation
2.1 Even parity discrete ordinates method for neutron transport equation
To provide a detailed description of the angular flux nonlinear finite difference equation, we begin with the multigroup neutron transport equation in the Cartesian coordinate system, as follows:
where vector r is the spatial position relative to the reference point, and vector Ω is the unit direction of flying with respect to the reference. Subscript g denotes the energy group index using multigroup approximation. This equation can represent the steady-state or transient neutron transport equation. When it represents a transient equation, the time variable is eliminated for brevity, as follows:
where Ss, Sf, Sd, and Se are the scattering, fission, delayed, and external sources, respectively. The external source is often disregarded, except when it must be considered. The fission and delayed neutron sources are isotropic. For the steady state, the time derivative term is zero, and the delayed neutron source is included in the fission source. If we discretize the time derivative term
The scattering and fission source can be expressed in terms of the angular flux as follows:
where ψ, φ, Σ, χ, μ0, and keff are the angular flux, scalar flux, macro cross sections, fission spectrum, cosine of scattering angle in the center of mass system, and eigenvalue with a standard definition in nuclear reactor physics, respectively. Using the discrete ordinates method, we can represent the angular space with selected angles as follows:
where
The even and odd parity fluxes are expressed as follows:
The even and odd parity fluxes exhibit the following properties:
Using Eqs. (4) and (5), we can obtain the even and odd parity transport equations as follows:
where
As the fission sources are isotropic, we have
where
Here,
μ0 is the cosine of the scattering angle in the center of mass system, where
where Pk are the Legendre polynomials and
Combined with Eqs. (9) and (13), the even and odd parity sources can be expressed as follows:
Integrate the even parity transport equation in Eq. (7a) over a region with constant a cross section as follows:
where subscript r is the region index;
It is noteworthy that the divergence theorem was used for deriving Eq. (15).
The surface of region r is divided into nr segments, where nr is the number of neighbors. Each segment is the interface between region r and one of its neighbors. Define the average odd parity flux on each segment as follows:
where n is the neighbor node of node r, and
By rearranging Eq. (15), we can obtain the even parity flux equation as follows:
The even parity flux equation exactly preserves the neutron balance in each direction and in each coarse mesh. If the relationship between the average odd parity flux
Integrating the odd parity transport equation (Eq. (7b)) over each segment of the surface, we obtain
The left hand of Eq. (21) is the average derivative of the even parity flux on interface
Substituting Eq. (22) into Eq. (21), we have
Through a finite difference approximation for
where
The average even parity flux on the surface exhibits following priority:
2.2 Definition and application of discontinuity factor
To render all
An algorithm was designed to determine the even and odd parity discontinuity factors. This algorithm begins by evaluating the following conditional:
If the condition in Eq. (29) is true, then the even and odd parity discontinuity factors can be calculated from the MOC solutions as follows:
If the condition is false, then different equations are designed to calculate the even and odd parity discontinuity factors as follows:
Using this approach, the even and odd parity discontinuity factors will be positive at all times. Rearranging Eq. (28) yields
If neighbor node n is within the computational domain, then a similar equation is obtained for the other side of the interface
Combining Eqs. (32) and (33) and the continuity condition
As for the vacuum boundary surface, the incident angular flux is zero. Therefore, we have the equation
For the reflective boundary surface, the angular fluxes in the reflect directions are equal. Therefore, we have the continuity condition
-202012/1001-8042-31-12-011/alternativeImage/1001-8042-31-12-011-F001.jpg)
Combined with Eq. (32) for both the angle i and i*, we obtain
In general, the relationships between the surface average odd parity flux and node average even parity flux for both the interior surface and boundary surface are established through Eqs. (34), (35), and (36). Moreover, these equations satisfy the high-order MOC solution through the strict definitions of the even and odd parity discontinuity factors.
2.3 Construction of angular flux nonlinear finite difference linear system
Based on Eqs. (34), (35), and (36), the odd parity fluxes on the interior and boundary surfaces can be expressed as a linear function of the neighbor average even parity fluxes using the coupling coefficient C.
For the interior surface, we have
where
For the vacuum boundary surface, we have
where
For the reflective boundary surface, we have
where
Substituting Eqs. (37)–(39) into Eq. (20), we obtain the following angular flux nonlinear difference linear system:
If the homogenized cross sections and coupling coefficients are obtained from the results of the MOC transport calculation, then the ANFD linear system can be solved. As the even parity fluxes in opposite directions are equal, only the ANFD linear system need to be solved on a half angular space.
3 Solving angular flux nonlinear difference equation
To solve the angular flux nonlinear difference equation, different solving strategies for decoupling the energy group and angle are proposed herein. If the scattering sources are determined simultaneously with the even parity fluxes, then all groups and all angles are coupled together. The terms on the right side of the linear system can be expressed as follows:
If all scattering sources other than the in-group scattering term are provided, then all angles are coupled together, and the linear system can still be solved by group separately, as follows:
If the scattering sources are all treated as specified, then the linear system can be solved by group and angle separately, as follows:
It is noteworthy that the last term in Eq. (41c) appears only when a reflect boundary condition exists on the node surface.
3.1 Decoupling of energy group and angle
To solve the linear system by group and angle separately, we would like to keep
where
(43)
where N, S, W, E, T, and B represent the north, south, west, east, top, and bottom surfaces of the node, respectively. The superscript k denotes the power iteration number. As the symmetric and diagonal dominant matrix is a symmetric positive definite matrix, Eq. (43) will become an expression for a symmetric positive definite linear system. Furthermore, the preconditioned conjugate gradient (CG) method [21] can be used to solve this linear system.
It is noteworthy that decoupling the angle and energy group will increase the iteration number of the ANFD method. However, the ANFD computing costs are low compared with those of MOC calculations, and they can be reduced significantly using the multilevel acceleration method, which has been implemented in several established core simulators [22-24]. The main idea of the multilevel acceleration method is to use a few-group CMFD linear system, such as a one- or two-group CMFD, to reduce the computing burden of the MG CMFD linear system. The few-group CMFD linear system is used to provide an effective estimation of the eigenvalue at the beginning of each power iteration, thereby significantly reducing the outer iteration number of the MG CMFD linear system. Therefore, a new approach similar to the multilevel acceleration method can be used to reduce the iteration number of ANFD required for convergence. This will be investigated further in future studies.
Meanwhile, if we wish to solve the linear system with all angles coupled or all energy groups coupled, it would be meaningless to maintain
3.2 Preconditioned Krylov subspace method
The direct solution of this linear system is unpractical as the matrix can become highly ill-conditioned, causing problems in the numerical solution. The Krylov subspace method has been proven to be extremely effective for solving large, nonsymmetric, sparse linear systems, particularly when a good preconditioner is applied. In this study, the right preconditioned generalized minimal residual (GMRES) method was used to solve the ANFD linear system. The right preconditioned GMRES algorithm can be summarized as follows:
-202012/1001-8042-31-12-011/alternativeImage/1001-8042-31-12-011-F010.jpg)
Meanwhile, if we decoupling the energy group and angle to construct a symmetric positive definite linear system, then the preconditioned conjugate gradient method can be used to solve the ANFD linear system. The left preconditioned conjugate gradient algorithm can be summarized as follows:
-202012/1001-8042-31-12-011/alternativeImage/1001-8042-31-12-011-F011.jpg)
As for precondition technology, a self-developed RSILU preconditioner has been used to improve parallel computing capability and minimize computational burden. The RSILU preconditioner does not require a multicolor ordering strategy for parallel computing and can provide convergence rates comparable to the standard incomplete lower–upper preconditioner. More details are available in the literature [25].
3.3 Iteration scheme of ANFD method
Once the node and surface angular flux obtained from the MOC calculation become available, the ANFD homogenized cross sections as well as the even and odd parity discontinuity factors can be easily obtained using Eq. (16) and (30), respectively. The strategy of decoupling the energy group and angle was used to construct the ANFD linear system, and the coupling coefficients were calculated using Eqs. (37)–(39). Hence, an equivalent ANFD linear system can be established. It is noteworthy that the asymmetric term arising from the construction of a symmetric positive definite matrix should be updated with the even party angular flux when solving the ANFD linear system. The right preconditioned GMRES solver was used to solve the ANFD linear system in this study. The exit from the GMRES solver to the ANFD cycle occurs when the required reduction in the relative L2-norm of the residual vector is achieved.
After solving the ANFD linear system, new even parity flux distributions and eigenvalues were obtained. To update the 2D MOC parameters, the average scalar flux on each node and the average angular flux on each surface were reconstructed as follows, indicated as
The former was used to update the scalar flux in the flat source regions, whereas the latter was used to update the incident angular flux on the boundary surface. The M in Eq. (44a) denotes the total number of discrete directions, which is consistent with the number of MOC azimuthal angles. After updating the average node scalar flux
-202012/1001-8042-31-12-011/alternativeImage/1001-8042-31-12-011-F002.jpg)
4 Numerical results
To investigate the equivalence of the ANFD method proposed herein, a code written in the C language was developed and three different test problems were selected: 2D pin, 2D assembly, and 2D C5G7 benchmark problems. In addition, a self-defined problem was tested to compare the acceleration capabilities of ANFD and CMFD in the void region.
For all these problems, the flux convergence criterion was 10-5 by comparing the infinite norm of the scalar flux in the flat source region, and the eigenvalue convergence criterion was 10-7. For the ray tracing parameter, a 0.01 cm ray spacing with 72 azimuthal angles and a Tabuchi–Yamamoto polar quadrature [26] with three polar angles per half-space were used for all the problems. The computer used to execute the code comprised an Inter(R) Core (TM) i5-7200U CPU @ 2.50 GHz with an 8.0G RAM. The basic computing unit for the ANFD was the pin cell.
For the 2D pin cell and 2D assembly problems, the eigenvalue and flux distribution error in the angle, energy group, and space were compared with those of the MOC calculations. In addition, three parameters were selected to assess the overall relative error of the even parity flux, i.e., the average relative error (AVG), root mean square of relative error (RMS), and maximum relative error (MAX), defined as follows:
where en is the relative error of the even parity flux compared with that of the MOC solution, and N is the total number of even parity fluxes.
4.1 2D pin cell problem
As indicated in Fig. 3(a), the side length of the cell was 1.26 cm, and the radius of the fuel pins was 0.54 cm for the 2D pin cell problem. The cross sections data for the UO2 fuel and moderator were obtained from the 2D C5G7 benchmark report. A vacuum boundary was applied to the east and south boundary surfaces, and the reflected boundary was applied to the north and west boundary surfaces. Three moderator rings and five fuel rings with 16 azimuthal divisions were selected for 128 flat-source regions.
-202012/1001-8042-31-12-011/alternativeImage/1001-8042-31-12-011-F003.jpg)
For the 2D pin cell problem, Fig. 5(a) shows the angular distribution of the even parity flux in the first energy group, and Fig. 5(b) shows the energy group distribution of the even parity flux in the first azimuth angle. It is clear that the even parity flux distribution of the ANFD was consistent with the MOC solution in both the angle and energy groups. Table 1 shows the eigenvalue difference and overall even parity flux relative error. The difference in the eigenvalue was 0 pcm, and the flux maximum relative error was 3.683E-06.
-202012/1001-8042-31-12-011/alternativeImage/1001-8042-31-12-011-F005.jpg)
keff | Flux relative error | ||||
---|---|---|---|---|---|
MOC | ANFD | Error (pcm) | AVG | RMS | MAX |
0.01561 | 0.01561 | 0 | 3.495E-07 | 9.145E-07 | 3.683E-06 |
4.2 2D assembly problem
As indicated in Fig. 3(b), the 2D assembly problem comprised a 17 × 17 lattice of square pin cells containing a UO2 fuel pin cell, guide tube, and fission chamber. The cross sections data for these materials were obtained from the 2D C5G7 benchmark problem. The division of the flat-source region was the same as that of the 2D pin cell problem. As for the boundary conditions, vacuum boundary conditions were applied to the east and south boundary surfaces, whereas the reflected boundary conditions were applied to the north and west boundary surfaces.
Fig. 6 shows the space relative error of the even parity flux in the first energy group and first azimuth angle for the 2D assembly problem. As shown, the even parity flux distribution of the ANFD indicated a relatively good agreement in space compared with the MOC solution. Table 2 shows the eigenvalue difference and overall even parity flux relative error for the 2D assembly problem. The difference in the eigenvalue was 0 pcm, whereas the flux maximum relative error was 1.162E-05.
-202012/1001-8042-31-12-011/alternativeImage/1001-8042-31-12-011-F006.jpg)
keff | Flux relative error | ||||
---|---|---|---|---|---|
MOC | ANFD | Error (pcm) | AVG | RMS | MAX |
0.96848 | 0.96848 | 0 | 1.030E-06 | 1.742E-06 | 1.162E-05 |
4.3 2D C5G7 problem
The 2D C5G7 benchmark problem was originally a pin-cell level heterogeneous benchmark for testing the accuracy of modern deterministic transport methods and codes without spatial homogenization. The 2D C5G7 problem configuration is shown in Fig. 4. The overall geometry of this problem was 64.26 cm × 64.26 cm, and each assembly measured 21.42 cm × 21.42 cm. Each fuel assembly comprised a 17 × 17 lattice of square pin cells, including a moderator, guide tube, fission chamber, MOX fuel pin cell with different enrichment, and UO2 fuel pin cell. For the reflector pin cases, the six additional pin cells of each reflector assembly used 1 × 1 sub mesh, whereas the 11 pin cells near the fuel used a 6 × 6 Cartesian sub mesh. For the other pin cells of the fuel and control rods, 128 flat-source regions comprising five fuel rings and three moderator rings with 16 azimuthal divisions were used. An accurate Monte-Carlo reference solution was provided in the 2D C5G7 benchmark report, which contains a precise eigenvalue solution and normalized pin power (NPP) that can be used as a reference solution. The NPP (
-202012/1001-8042-31-12-011/alternativeImage/1001-8042-31-12-011-F004.jpg)
where
Figure 7 indicates the percent error of the NPP for the 2D C5G7 benchmark problem; as shown, the main error appeared in the fuel pin cell near the reflector.
-202012/1001-8042-31-12-011/alternativeImage/1001-8042-31-12-011-F007.jpg)
Table 3 shows the eigenvalue difference and NPP error from our results and the MCNP reference solutions. Compared with the reference results, the error of the eigenvalue was 3 pcm, which was less than the stochastic uncertainty of the MCNP reference results. In general, the eigenvalue, NPP, etc. agreed well with the reference results.
Parameters | Reference MCNP (Statistic error) | ANFD | Error |
---|---|---|---|
Eigenvalue | |||
keff (error/pcm) | 1.18655 (±9.5 pcm) | 1.18652 | -3 pcm |
Specific NPP data | |||
Maximum NPP (error) | 2.498 (±0.16%) | 2.494 | -0.14% |
Minimum NPP (error) | 0.232 (±0.58%) | 0.234 | +1.04% |
Normalized assembly power data | |||
Inner UO2 power (error) | 492.8 (±0.10%) | 492.3 | -0.10% |
MOX power (error) | 211.7 (±0.18%) | 211.9 | +0.10% |
Outer UO2 power (error) | 139.8 (±0.20%) | 139.9 | +0.10% |
NPP error results | |||
MAX error | -- | 1.25% | -- |
AVG error | 0.32% | 0.21% | -- |
RMS error | 0.34% | 0.30% | -- |
MRE error | 0.27% | 0.16% | -- |
4.4 Void region problem
The ANFD method was developed from transport theory and does not require the calculation of the diffusion coefficient. Therefore, the ANFD method may exhibit a better acceleration than the CMFD method in certain strongly anisotropic cases, such as those involving void regions and anisotropic scattering. In this study, a lattice problem involving void regions was constructed and tested; 5 × 5 square pin cells were set to the void regions, whereas UO2 fuel cells were used in the other regions. The geometry and cross sections of the UO2 fuel cell were the same as those provided in the 2D C5G7 benchmark report. The reflected boundary conditions were applied to all boundary surfaces. The layout of this void region problem is shown in Fig. 8.
-202012/1001-8042-31-12-011/alternativeImage/1001-8042-31-12-011-F008.jpg)
The convergence history of the MOC, ANFD, and CMFD are shown in Table 4 and Fig. 9. As expected, the CMFD failed to converge in this problem, whereas the ANFD converged and remained effective.
-202012/1001-8042-31-12-011/alternativeImage/1001-8042-31-12-011-F009.jpg)
keff | Error | MOC iteration number | Speedup | |
---|---|---|---|---|
No acceleration | 1.32592 | -- | 91 | -- |
CMFD | Divergence | -- | -- | -- |
ANFD | 1.32602 | 10 pcm | 11 | 8.3 |
Void regions are typical in real reactors, such as TREAT [27] and HTR-10 [28]. In a previous study [29], the quasi-diffusion method was used to address the void region problem in TREAT core analysis. In this study, the quasi-diffusion method was implemented using the nodal expansion method. Meanwhile, the nodal wise average cross sections, Eddington factors, and discontinuity factors were generated from 3D whole-core Serpent Monte-Carlo simulations. By comparison, the ANFD is more concise and only requires even and odd parity discontinuity factors generated from MOC calculations.
5 Conclusion
In this study, an innovative equivalent low-order ANFD method for MOC transport calculations was developed and implemented. The even parity discrete ordinates method was used to build a low-order ANFD equation, and different boundary condition treatments were proposed. Even and odd parity discontinuity factors based on the flux information obtained from MOC calculations were strictly defined to ensure the equivalence between this low-order equation and the high-order pin-resolved transport equation. The energy group and angle were decoupled to construct an easier-to-solve symmetric linear system. The self-developed RSILU preconditioned GMRES solver was applied to solve this linear system. The equivalence of the ANFD method was verified for 2D pin, 2D assembly, and 2D C5G7 benchmark problems. The numerical results demonstrated that a low-order ANFD equation that was completely equivalent to the pin-resolved transport equation was established.
This study provides an important foundation for further developing a novel acceleration technique based on the ANFD method. Only an equivalent equation was established, and a converged consistent solution was obtained to achieve acceleration. In the future, the acceleration performance of the ANFD will be tested. Meanwhile, we will compare the pin-homogenized ANFD method with the pin-homogenized MG CMFD method, which is widely used to accelerate MOC calculations. Our preliminary analysis indicated that the ANFD method achieved an excellent and stable acceleration performance compared with the CMFD method when void regions existed. Other anisotropic cases, such as anisotropic scattering, will be tested in our future studies. Subsequently, Fourier analysis will be used for convergence analysis. Additionally, we will investigate the use of the multilevel acceleration method to reduce the computing burden of ANFD and then attempt to extend this method to 3D cases.
Fusion of method of characteristics and nodal method for 3-d whole-core transport calculation
. Trans. Am. Nucl. Soc. 87, 322 (2002).Methods and performance of a three dimensional whole-core transport code DeCART
.Solution of the C5G7MOX benchmark three-dimensional extension problems by the DeCART direct whole core calculation code
. Prog. Nucl. Energy 48, 456 (2006). https://doi.org/10.1016/j.pnucene.2006.01.006Practical numerical reactor employing direct whole core neutron transport and subchannel thermal/hydraulic solvers
. Ann. Nucl. Energy 62, 357 (2013). https://doi.org/10.1016/j.anucene.2013.06.031Stability and accuracy of 3D neutron transport simulations using the 2D/1D method in MPACT
. J. Comput. Phys. 326, 612(2016). https://doi.org/10.1016/j.jcp.2016.08.022Accuracy and analysis of the iteratively coupling 2D/3D method for the three-dimensional transport calculations
. Ann. Nucl. Energy 113, 130 (2018). https://doi.org/10.1016/j.anucene.2017.10.020Verification and validation of STREAM/RASTK for PWR analysis
. Nucl. Eng. Tech. 51 (2), 356 (2019). https://doi.org/10.1016/j.net.2018.10.004Spatial Homogenization Methods for Light Water Reactors. Doctoral thesis
,An optimally diffusive coarse mesh finite difference method to accelerate neutron transport calculations
. Ann. Nucl. Energy 95, 116 (2016). https://doi.org/10.1016/j.anucene.2016.05.004Two-level convergence speedup schemes for p-CMFD acceleration in neutron transport calculation
. Nucl. Sci. Eng. 188, 1(2017). https://doi.org/10.1080/00295639.2017.1332891Convergence Analysis of a CMFD Method Based on Generalized Equivalence Theory
.A linear prolongation approach to stabilizing CMFD
. Nucl. Sci. Eng. 190, 45 (2018). https://doi.org/10.1080/00295639.2017.1417347Generalized coarse-mesh finite difference acceleration for method of characteristics in unstructured meshes
. Chinese Journal of Computational Physics, 27(4), 541 (2010).3D whole-core neutron transport simulation using 2D/1D method via multi-level generalized equivalence theory based CMFD acceleration
. Ann. Nucl. Energy 122, 79 (2018). https://doi.org/10.1016/j.anucene.2018.08.014A new three-level CMFD method based on the loosely coupled parallel strategy
. Ann. Nucl. Energy 145, 107529 (2020). https://doi.org/10.1016/j.anucene.2020.107529Two-level time-dependent GET based CMFD acceleration for 3D whole core transient transport simulation using 2D/1D method
. Ann. Nucl. Energy 142, 107405 (2020). https://doi.org/10.1016/j.anucene.2020.107405Transient methods for pin-resolved whole-core neutron transport. Doctoral thesis
,Analysis of stabilization techniques for CMFD acceleration of neutron transport problems
. Nucl. Sci. Eng. 184, 208 (2016). https://doi.org/10.13182/NSE16-51Stability analysis of the CMFD scheme with linear prolongation
. Ann. Nucl. Energy 129, 298 (2019). https://doi.org/10.1016/j.anucene.2019.02.011Benchmark on deterministic 2-D MOX fuel assembly transport calculations without spatial homogenization
. Prog. Nucl. Energy 45, 107 (2004). https://doi.org/10.1016/j.pnueene.2004.09.003Multi-level Coarse Mesh Finite Difference Acceleration with Local Two-Node Nodal Expansion Method
. Ann. Nucl. Energy 116, 105 (2018). https://doi.org/10.1016/j.anucene.2018.02.002Two-level coarse mesh finite difference formulation with multigroup source expansion nodal kernels
. J. Nucl. Sci. Technol. 45, 668 (2008).Implementation of two-level coarse-mesh finite difference acceleration in an arbitrary geometry, two-dimensional discrete ordinates transport method
. Nucl. Sci. Eng. 158 (3), 289 (2008). https://doi.org/10.13182/NSE06-24TNA novel and efficient hybrid RSILU preconditioner for the parallel GMRES solution of the coarse mesh finite difference equations for practical reactor simulations
. Nucl. Sci. Eng. 194(2), 104 (2019). https://doi.org/10.1080/00295639.2019.1657322Derivation of optimum polar angle quadrature set for the method of characteristics based on approximation error for the Bickley function
. J. Nucl. Sci. Technol. 44 (2), 129 (2007). https://doi.org/10.1080/18811248.2007.9711266“Design Summary Report on the Transient Reactor Test Facility (TREAT)”, ANL-6034
,The design features of the HTR-10
, Nucl. Eng. Des. 218 (3), 25 (2002). https://doi.org/10.1016/S0029-5493(02)00182-6Quasi-diffusion method with 3-D cross sections for TREAT core analysis
, Nucl. Technol. 206(6), 825, (2020). https://doi.org/10.1080/00295450.2019.1672451