3,706
Views
0
CrossRef citations to date
0
Altmetric
Research Article

On Solution Regularity of the Two-Dimensional Radiation Transfer Equation and Its Implication on Numerical Convergence

Abstract

In this paper, we deal with the differential properties of the scalar flux ϕ(x) defined over a two-dimensional bounded convex domain, as a solution to the integral radiation transfer equation. Estimates for the derivatives of ϕ(x) near the boundary of the domain are given based on Vainikko’s regularity theorem. The optimal pointwise error estimates in terms of the scalar flux are presented for the two classic finite difference methods: diamond difference (DD) and step difference (SD). Numerical results indicate the implication of the solution smoothness on the numerical convergence behavior.

1. Introduction

The one-group radiation transfer problem in a three-dimensional (3D) convex domain reads as follows: find an angular flux function ψ: G¯×S2R+ such as (1) j=13Ωjψ(x,Ω)xj+σ(x)ψ(x,Ω)=σs(x)4πΩs(x,Ω,Ω)ψ(x,Ω)dΩ+q(x,Ω), xG,(1) (2) ψ(x,Ω)=ψG(x,Ω), xG, Ω·n̂(x)<0,(2) where G¯ is the closure of the domain GR3 with the boundary G, S2 is the unit sphere of R3, Ω denotes the direction of radiation transfer, σ is the extinction coefficient (or macroscopic total cross section in neutron transport), σs is the scattering coefficient (or macroscopic scattering cross section), s is the phase function of scattering with Ωs(x,Ω,Ω)dΩ=4π, q is the external source function, and n̂ is the unit normal vector of the domain surface. Note that σs(x)<σ(x) under the subcritical condition, and s(x,Ω,Ω)=s(x,Ω,Ω).

Assuming the isotropic scattering s(x,Ω,Ω)=1, isotropic source q(x,Ω)=Q(x)4π, and isotropic boundary condition ψG(x,Ω)=ϕG(x)4π, we can obtain the so-called Peierls integral equation of radiation transfer for the scalar flux ϕ(x) as follows: (3) ϕ(x)=14πGσs(y)eτ(x,y)|xy|2ϕ(y)dy+14πGeτ(x,y)|xy|2Q(y)dy+14πGeτ(x,y)|xy|2|xy|xy|·n(x)|ϕG(y)dSy,(3) (4) τ(x,y)=0|xy|σ(rξΩ)dξ,(4) where dS is the differential element of the domain surface, τ(x,y) is the optical path between x and y. One can find detailed derivation in Vainikko (Citation1993).

For simplicity, we assume σ and σs are constant over the domain. Then EquationEquation (3) can be simplified as (5) ϕ(x)=GK(x,y)ϕ(y)dy+1σsGK(x,y)Q(y)dy+1σsGK(x,y)|xy|xy|·n(x)|ϕG(y)dSy(5) where the 3D radiation kernel is given as (6) K(x,y)=σseσ|xy|4π|xy|2.(6)

The boundary integral term in the above equation can produce singularities in the solution. We omit its discussion in this paper. In other words, here we only consider the problem with the vacuum boundary condition, i.e., ϕG=0. Thus, EquationEquation (6) can be treated as the weakly integral equation of the second kind: (7) ϕ(x)=GK(x,y)ϕ(y)dy+f(x), xG(7) where GRn, n1, is an open bounded domain and the kernel K(x,y) is weakly singular, i.e., |K(x,y)|C|xy|ν, 0νn. Note that f(x)=1σsGK(x,y)Q(y)dy for radiation transfer. Weakly singular integral equations arise in many physical applications such as elliptic boundary problems and particle transport.

Since K(x,y) has a singularity at x=y, the solution of a weakly integral equation is generally not a smooth function and its derivatives at the boundary would become unbounded from a certain order. There was extensive research on the smoothness (regularity) properties of the solutions to weakly integral equations (Mikhlin Citation1965; Mikhlin and Prossdorf Citation1986), especially those early work in neutron transport theory done in the former Soviet Union (Vladimirov Citation1963; Germogenova Citation1969). It is believed that Vladimirov first proved that the scalar flux ϕ(x) possesses the property |ϕ(x+h)ϕ(x)|hlogh for the one-group transport problem with isotropic scattering in a bounded domain (Vladimirov Citation1963). Germogenova analyzed the local regularity of the angular flux ψ(x,Ω) in a neighborhood of the discontinuity interface and obtained an estimate of the first derivative, which has the singularity near the interface (Germogenova Citation1969). Pitkäranta (Citation1980) derived a local singular resolution showing explicitly the behavior of ϕ(x) near the smooth portion of the boundary. Vainikko (Citation1993) introduced weighted spaces and obtained sharp estimates of pointwise derivatives near the smooth boundary for multidimensional weakly singular integral equations.

There exists some previous research work on the regularity of the integral radiation transfer solutions (Johnson and Pitkaranta Citation1983; Hennebach, Junghanns, and Vainikko Citation1995). However, the 2D kernel treated in those studies bears no physical meaning since it does not model the scattering correctly. A physically relevant 2D kernel can be found in Lewis and Miller (Citation1993). In this paper, we rederive the 2D kernel by directly integrating the 3D kernel with respect to the third dimension. We examine the differential properties of the new 2D kernel and provide estimates of pointwise derivatives of the scalar flux according to Vainikko’s regularity theorem for the weakly integral equation of the second kind.

The remainder of the paper is organized as follows. In Sect. 2, we derive the 2D kernel for the integral radiation transfer equation. We examine the derivatives of the kernel and show that they satisfy the boundedness condition of Vainikko’s regularity theorem in Sect. 3. Then the estimates of local regularity of the scalar flux near the boundary of the domain are given. In Sect. 4, we derive local estimates for the spatial discretization error in terms of the scalar flux for the two classical numerical methods, diamond difference (DD) and step difference (SD). Sect. 5 presents numerical results of DD and SD to demonstrate how the asymptotic convergence behavior of the spatial discretization error is affected by the smoothness of the exact solution. Concluding remarks are given in Sect. 6.

2. Two-dimensional radiation transfer equation

In this section, we derive the 2D integral radiation transfer equation from its 3D form, EquationEqs. (5) and Equation(6). In 3D, dy=dy1dy2dy3 and |xy|=(x1y1)2+(x2y2)2+(x3y3)2. Let ρ=(x1y1)2+(x2y2)2, then |xy|=ρ2+(x3y3)2. In a 2D domain GR2, the solution function ϕ(x) only depends on x1 and x2 in Cartesian coordinates. Therefore, we only need to find the 2D radiation kernel, which can be obtained by integrating out y3 as follows: (8) K(x,y)=-σseσ|xy|4π|xy|2dy3=σs4π-eσρ2+(x3y3)2ρ2+(x3y3)2dy3.(8)

To proceed, we introduce the variables t=σρ2+(x3y3)2 and z=y3x3. Then we substitute dy3=dz=tσt2σ2ρ2dt into the above equation to have (9) K(x,y)=σs4π-ett2σ2dz=σs2π0ett2σ2dz=σs2πσρett2σ2tσt2σ2ρ2dt=σsσ2πσρettt2σ2ρ2dt=σs2πρ01eσρt1t2dt.(9)

The above derivation of the 2D kernel is much simpler than the process given in Lewis and Miller (Citation1993), where the integral form is obtained by the projection of the particle flight path on the 2D plane. Note that the last integral is the first Bickley-Naylor function Ki1(r)=01ert1t2dt (Abramowitz and Stegun Citation1970; Altaç Citation2007).

Now we show that the 2D kernel K(x,y) has a singularity at ρ=0 (i.e., x=y) as follows. As ρ is sufficiently small, we can have σρ<1 and (10) K(x,y)=σs2πρ01eσρt1t2dt>σs2πρ01e1t1t2dt>0.3σs2πρ.(10)

Thus, it can be seen that K(x,y) becomes unbounded at ρ=0.

By replacing the 3D kernel as defined by EquationEquation (6) with the above one, EquationEquation (5) becomes the 2D integral radiation transfer equation. Notice that the surface integral in the last term on the right-hand side of EquationEquation (5) should be replaced with a line integral in the 2D domain.

Remark 2.1.

By simply assuming a 2D scattering, Johnson and Pitkaranta derived a 2D kernel, i.e., K(x,y)=e|xy||xy| (where σ=1), which is however physically incorrect since the scattering is essentially a 3D phenomenon (Johnson and Pitkaranta Citation1983). Hennebach, Junghanns, and Vainikko (Citation1995) also used the same 2D kernel for analyzing the radiation transfer solutions. In addition, the integral equations in other geometries such as slab or sphere can be obtained by following the same approach, and they can be found in Bell and Glasstone (Citation1970).

Applying Banach’s fixed-point theorem, we can prove the existence and uniqueness of the solution in the 2D domain by showing that GK(x,y)dy  is bounded below unity as follows. (11) GK(x,y)dy=G(σsσ2πσρettt2σ2ρ2dt)dy=σsσ2πGdyσρettt2σ2ρ2dt=σsσ2πGρdφdρσρettt2σ2ρ2dt(11) where φ is the azimuthal angle. By extending the above bounded domain to the whole space, we have (12) GK(x,y)dy<σsσ2π02πρdρσρettt2σ2ρ2dt=σsσ0ρdρσρettt2σ2ρ2dt(12)

Denoting ζ=σρ, EquationEquation (12) is simplified as (13) GK(x,y)dy<σsσ0ζdζζettt2ζ2dt=σsσ0ζettζt2ζ2dtdζ =σsσ0ettdt0tζt2ζ2dζ=σsσ0etdt=σsσ<1.(13)

Notice we have changed the order of integration to solve the integral. It is apparent that for there exists a unique solution, the subcritical condition, σs<σ, must be satisfied.

Remark 2.2.

A similar proof can be found for the 2D kernel in terms of the Bickley-Naylor function (de Azevedo et al. Citation2018). Existence of the unique solution to the general neutron transport equation in L1 and L has been long established in Case and Zweifel (Citation1967). Corresponding results in Lp for 1  p  can be found in (Agoshkov Citation1998; Latrach Citation2001; Latrach and Zeghal Citation2012; Egger and Schlottbom Citation2014).

3. Solution regularity

We first introduce Vainikko’s regularity theorem (Vainikko Citation1993), which provides a sharp characterization of singularities for the general weakly integral equation of the second kind. Then we analyze the differential properties of the 2D radiation kernel and show that the derivatives are properly bounded. Finally, Vainikko’s theorem is used to give the estimates of pointwise derivatives of the radiation transfer solution.

3.1. Vainikko’s regularity theorem

Before we state the theorem, we introduce the definition of weighted spaces Cm,ν(G).

Weighted space Cm,ν(G). For a λR, introduce a weight function (14) wλ={1, λ<0(1+|logϱ(x)|)1, λ=0ϱ(x)λ, λ>0, xG(14) where GRn is an open bounded domain and ϱ(x)=infyG|xy| is the distance from x to the boundary G. Let mN, νR and ν<n. Define the space Cm,ν(G) as the set of all m times continuously differentiable functions ϕ:G R such that (15) ϕm,ν=αmsupxG(wα(nν)|Dαϕ(x)|)<.(15)

In other words, a m times continuously differentiable function ϕ on G belongs to Cm,ν(G) if the growth of its derivatives near the boundary can be estimated as follows: (16) |Dαϕ(x)|c{1, α<nν1+|logϱ(x)|, α=nνϱ(x)nνα, α>nν, xG, αm(16) where c is a constant. The space Cm,ν(G), equipped with the norm ·m,ν, is a complete Banach space.

After defining the weighted space, we introduce the smoothness assumption about the kernel in the following form: the kernel K(x,y) is m times continuously differentiable on (G×G)\{x=y} and there exists a real number ν(,n) such that the estimate (17) |DxαDx+yβK(x,y)|c{1, ν+α<01+|log|xy||, ν+α=0|xy|να, ν+α>0, x,yG(17) where (18) Dxα=(x1)α1(xn)αn,(18) (19) Dx+yβ=(x1+y1)β1(xn+yn)βn(19) holds for all multi-indices α=(α1,,αn)Z+n and β=(β1,,βn)Z+n with α+βm. Here the following usual conventions are adopted: α=α1++αn, and |x|=x12++xn2.

Now we present Vainikko’s theorem in characterizing the regularity properties of a solution to the weakly integral equation of the second kind (Vainikko Citation1993).

Theorem 3.1.

Let GRn be an open bounded domain, fCm,ν(G) and let the kernel K(x,y) satisfy the condition Equation(17). If the integral Equationequation (7) has a solution, ϕL(G) then ϕCm,ν(G).

Remark 3.1.

The solution does not improve its properties near the boundary G, remaining only in Cm,ν(G), even if G is of class C, and fC(G). A proof can be found in Vainikko (Citation1993). More precisely, for any n and ν (ν<n) there are kernels K(x,y) satisfying Equation(17) and such that EquationEquation (7) is uniquely solvable and, for a suitable fC(G), the normal derivatives of order k of the solution behave near G as logϱ(x) if k=nν, and as ϱ(x)nνk for k>nν.

3.2. Regularity of the radiation transfer solution

To apply the results of Theorem 3.1 to the 2D integral radiation transfer equation, we need to analyze the kernel K(x,y) and show that it satisfies the condition Equation(17), i.e., |DxαDx+yβK(x,y)|c|xy|1α. We can simply set |β|=0 without loss of generality for our problem. α=0: (20) K(x,y)=σs2πρ01eσρt1t2dt<σs2πρ0111t2dt=σs2πρπ2 =σs4ρc|xy|1.(20) α>1: Let ζ=σρ=σ|xy|=σ(x1y1)2+(x2y2)2, then K(x,y)=σs2πζ01eζt1t2dt. By the chain rule, Dxα can be written as (21) Dxα=(x1)α1(x2)α2=(ζζx1)α1(ζζx2)α2=αζα(ζx1)α1(ζx2)α2.(21)

Then we have (22) |DxαK(x,y)|=|αζα(ζx1)α1(ζx2)α2K(x,y)|=|(ζx1)α1(ζx2)α2||αζαK(x,y)|(22) where (23a) |ζx1|=σ|(x1y1)(x1y1)2+(x2y2)2|σ,(23a) (23b) |ζx2|=σ|(x2y2)(x1y1)2+(x2y2)2|σ.(23b)

Substituting EquationEqs. (23a) and Equation(23b) into Equation(22), we obtain (24) |DxαK(x,y)|σα|αζαK(x,y)|.(24)

Apparently, we only need to find the upper bound of |αζαK(x,y)|cζ(α+1), which is shown as follows. (25) |αζαK(x,y)|=|αζα(σs2πζ01eζt1t2dt)|=σs2π|αζα(1ζ01eζt1t2dt)|=σs2π|(1)αζα+101eζt1t2dt+1ζαζα(01eζt1t2dt)|σs2π1ζα+1|01eζt1t2dt|+σs2π1ζ|αζα(01eζt1t2dt)|<σs2π1ζα+1|0111t2dt|+σs2π1ζ|αζα(01eζt1t2dt)|=σs41ζα+1+σs2π1ζ|αζα(01eζt1t2dt)|.(25)

Now we need to find the bound for |αζα(01eζt1t2dt)| in the above equation, which is analyzed as follows. (26) |αζα(01eζt1t2dt)|=|01αζα(eζt)1t2dt|=|01eζttα1t2dt|=|0ζeζttα1t2dt+ζ1eζttα1t2dt|<|0ζeζttα1t2dt+1ζαζ1eζt1t2dt| <|0ζeζttα1t2dt|+1ζα|0111t2dt|=|0ζeζttα1t2dt|+π2ζα=1ζα|1tαet(tζ)21dt|+π2ζα<cζα . (26)

Note that in the above equation, 0<ζ<1 and |1tαet(tζ)21dt|<, i.e., it is bounded. Substituting EquationEquation (26) into Equation(25) and noting that ζ=σ|xy|, we obtain (27) |DxαK(x,y)|<c|xy|α1.(27)

Finally, we conclude that the 2D radiation kernel satisfies the condition Equation(17). Therefore, by Theorem 3.1, the estimates of derivatives of the scalar flux ϕ(x) for radiation transfer are the same as those for the general weakly integral equation of the second kind: (28) |Dαϕ(x)|c{1, α<11+|logϱ(x)|, α=1ϱ(x)1α, α>1, xG.(28)

It should be noted that in radiation transfer f(x)=1σsGK(x,y)Q(y)dy, where Q is the external source. If the function Q is bounded, then fCm,ν(G) (Vainikko Citation1993).

Remark 3.2.

The first derivative of the solution ϕ(x)  behaves as logϱ(x) and becomes unbounded as approaching the boundary. The derivatives of order k behave as ϱ(x)1k for k>1. These pointwise estimates cannot be improved by adding more strong smoothness on the data and domain boundary. Here, we only deal with the solution regularity in the normal direction near the boundary. It is noted that the tangential derivatives behave essentially better than the normal derivatives (Vainikko Citation1993). In other words, the solution is smoother (or less singular) in the tangential direction than the normal direction near the boundary. The obtained pointwise bounds on the derivatives are strong results, which imply bounds in other norms such as Sobolev norms. Golse et al. (Citation1988) studied the regularity of the mean value (with respect to velocity) of the solution of transport equations in terms of fractional Sobolev space.

Remark 3.3.

These sharp error estimates are needed to analyze numerical methods. The lack of smoothness in the exact solution could adversely affect the convergence rate of spatial discretization schemes for solving the radiation transfer equation (Madsen Citation1972; Larsen Citation1982; Wang and Ragusa Citation2009). The spatial discretization error of a numerical method can be expressed as |ϕjϕjN|Chjpϕ(k), where ϕj is the exact solution, ϕjN is its numerical result, hj is the mesh size, p is the order of accuracy, and ϕ(k) is the k-th derivative of the scalar flux. According to the above regularity results, we have ϕ(k)chj1k, then |ϕjϕjN|Chjp+1k. However, for any numerical methods, we would generally have kp and thus the spatial convergence rate would asymptotically reduce to the first order of accuracy or even worse as the mesh is refined.

4. Error analysis of DD and SD

In this section, we present the optimal error estimates in terms of the scalar flux for the two classic numerical methods: DD and SD. The DD method is a second-order central finite difference method, while the SD method is a first-order upwind method. In the next section, numerical results of a homogeneous unit-square problem are presented for both DD and SD to show how the convergence behavior of numerical solutions are impeded by the underlying regularity of the exact transport solution. The derivation follows our previous analysis (Wang Citation2022) using the discrete maximum principle (Miller, O’Riordan, and Shishkin Citation2012; Stynes and Stynes Citation2018). Although the optimal error estimates are obtained for the 1D slab geometry, it is applicable for 2D problems since we are only concerned with the smoothness in the normal direction to the boundary in this work.

4.1. Error estimate for DD

The monoenergetic SN equation in slab geometry with the assumption of isotropic scattering and constant external neutron source is written as (29) μmdψmdx+σψm=σs2m=1Mψmwm+Q2,(29) where μm, m=1,,M, is the neutron direction cosine with respect to the x-axis, ψm is the angular flux, wm is the quadrature weight, σ is the total macroscopic cross section, σs is the macroscopic scattering cross section, and Q is the external neutron source.

The mesh gird employed for discretizing DD and SD is shown in . The index j denotes the center of cell j, j1/2 is the left edge, and j+1/2 is the right edge.

Figure 1. One-dimensional mesh.

Figure 1. One-dimensional mesh.

The DD discretization for the SN equation on the one-dimensional mesh can be written as (30) μmhj(ψm,j+12Nψm,j12N)+σj(ψm,j12N+ψm,j+12N2)=12σs,jm=1M(ψm,j1/2N+ψm,j+1/2N2)wm+Qj2 for j = 1,,N,(30) where (31a) hj=xj+1/2xj1/2,(31a) (31b) ψm,0N=f(μm),(31b) (31c) ψm,NN=g(μm).(31c) σj σs,j, and Qj are assumed to be piecewise constant in each cell. The superscript "N" denotes the numerical solution.

We introduce the linear discrete operator Lψ,DDN(ψm,jN) for the DD discretization as (32) Lψ,DDN(ψm,jN)=Qj2,(32) where (33) Lψ,DDN(ψm,jN)=μmhj(ψm,j+12Nψm,j12N)+σj(ψm,j12N+ψm,j+12N2)12σs,jm=1M(ψm,j1/2N+ψm,j+1/2N2)wm.(33)

By multiplying Lψ,DDN(ψm,jN) with each quadrature weight wm and summing them from m=1 to M, we can obtain the DD discrete operator for the scalar flux: (34) Lϕ,DDN(ϕjN)=m=1MLψ,DDN(ψm,jN)wm =Qj,(34) where (35) Lϕ,DDN(ϕjN)=m=1M[μmhj(ψm,j+12Nψm,j12N)+σj(ψm,j12N+ψm,j+12N2)]wmm=1M[12σs,jm=1M(ψm,j1/2N+ψm,j+1/2N2)wm]wm=1hj(Jj+1/2NJj1/2N)+σa,j(ϕj1/2N+ϕj+1/2N2),(35) with (36a) ϕj1/2N=m=1Mψm,j1/2Nwm,(36a) (36b) Jj1/2N=m=1Mμmψm,j1/2Nwm,(36b) (36c) σa,j=σjσs,j.(36c)

The local truncation error for the scalar flux is calculated as (37) |Lϕ,DDN(ϕjϕjN)|=|Lϕ,DDN(ϕj)Lϕ,DDN(ϕjN)|=|Lϕ,DDN(ϕj)Qj|=|1hj(Jj+1/2Jj1/2)+σa,j(ϕj1/2+ϕj+1/22)Qj|,(37) where ϕj denotes the exact solution. Note that in deriving EquationEquation (37), we have used EquationEqs. (34) and Equation(35).

Now we introduce the linear operator L for the continuous neutron balance equation as (38) Lϕ=ddxJ+σaϕ=Q.(38)

From EquationEquation (38), we can have (39) 12(Jj1/2+Jj+1/2)+σa,j(ϕj1/2+ϕj+1/22)=Qj.(39)

We substitute EquationEquation (39) into Equation(37) to obtain (40) |Lϕ,DDN(ϕjϕjN)|=|1hj(Jj+1/2Jj1/2)12(Jj1/2+Jj+1/2)|.(40)

By Taylor series expansion at xj, we obtain (41) |Lϕ,DDN(ϕjϕjN)|Chj2max|Jj(ξj)|,(41) where ξj[xj1/2, xj+1/2], and C is a sufficiently large constant.

Differentiating EquationEquation (38) twice with respect to x, we obtain (42) Jj(ξj)=σaϕ(ξj).(42)

Note that we have assumed that Q is piecewise constant in each cell. Plugging EquationEquation (42) into Equation(41), the local truncation error is to satisfy (43) |Lϕ,DDN(ϕjϕjN)|Cσahj2max|ϕ(ξj)|,(43)

Next, we use the discrete maximum principle to find the local bound of (ϕjϕjN). We define (44) M=1σa,jmax1jN|Lϕ,DDN(ϕjϕjN)|,(44) and (45) Φj=M±(ϕjϕjN).(45)

Then for 1jN, we have (46) Lϕ,DDN(Φj)=Lϕ,DDN(M)±Lϕ,DDN(ϕjϕjN).(46)

Since M is a constant, we find from EquationEquation (35) (47) Lϕ,DDN(M)=σa,jM.(47)

Substituting EquationEquation (44) into EquationEquation (47), we obtain (48) Lϕ,DDN(M)=max1jN|Lϕ,DDN(ϕjϕjN)|.(48)

Then using EquationEquation (48) in EquationEquation (46), we can have (49) Lϕ,DDN(Φj)0.(49)

However, the above result does not warrant a positive solution Φj0 since the matrix form of LDDN is not always an M-matrix. It is known that the DD method is strictly positive when the mesh size hj2μnσj, which is for the most limiting pure absorbing case (i.e., the scattering ratio c=0). Nevertheless, this condition can be relaxed as the problem becomes more diffusive (i.e., c>0). In this work, we are only interested in the asymptotic convergence behavior. We can have Φj0 when the mesh is sufficiently fine. Then from EquationEquation (45) we obtain by using EquationEquation (43) (50) |ϕjϕjN|M=1σa,jmax1jN|Lϕ,DDN(ϕjϕjN)|Chj2max1jN|ϕ(ξj)|.(50)

Therefore, the DD has the second order convergence rate in spatial discretization if ϕ is bounded. It is the case for 1D geometry, but for multi-dimensional geometry the derivatives of the exact solution become unbounded near the boundary as shown in the previous section.

4.2. Step difference (SD)

The SD discretization for the SN transport equation can be written as (51a) μmhj(ψm,j+1/2Nψm,j1/2N)+σjψm,j+1/2N=12σs,jm=1Mψm,j+1/2Nwm+Qj2, μm>0,(51a) (51b) μmhj(ψm,j+1/2Nψm,j1/2N)+σjψm,j1/2N=12σs,jm=1Mψm,j1/2Nwm+Qj2, μm<0.(51b)

Unlike DD, the SD method employs an upwind stencil instead of the symmetric one. More care should be taken to carry out the error analysis for the SD method.

We introduce the SD discrete operator (52) Lϕ,SDN(ϕjN)=Qj,(52) where (53) Lϕ,SDN(ϕjN)=m=1M2[μmhj(ψm,j+1/2Nψm,j1/2N)+σjψm,j+1/2N12σs,jϕj+1/2N]wm+m=M/2+1M[μmhj(ψm,j+1/2Nψm,j1/2N)+σjψm,j1/2N12σs,jϕj1/2N]wm=1hj(Jj+12NJj12N)+σjm=1M2ψm,j+12Nwm12σs,jϕj+12N+σjm=M2+1Mψm,j12Nwm12σs,jϕj1/2N(53)

The local truncation error for the scalar flux is calculated as (54) |Lϕ,SDN(ϕjϕjN)|=|Lϕ,SDN(ϕj)Lϕ,SDN(ϕjN)|=|Lϕ,SDN(ϕj)Qj|,(54)

By Taylor series expansion of Lϕ,SDN(ϕj) at xj, we have (55) |Lϕ,SDN(ϕjϕjN)|Chjσj|m=1M2ψm,jwmm=M2+1Mψm,jwm|Chjσj|m=1M2ψm,jwmm=M2+1Mψm,jwm|Chjσjmax|m=1Mψm,j(ξj)wm|=Chjσjmax|ϕ(ξj)|.(55) where ξj[xj1/2, xj+1/2], and C is a sufficiently large constant. In the above derivation, we have used the inequality, |m=1M2ψm,jwmm=M2+1Mψm,jwm|C·max|m=1M2ψ(ξj)wm+m=M2+1Mψ(ξj)wm|, which can be always found to hold for some ξj[xj1/2, xj+1/2] when C is sufficiently large. If ϕ=0, then the solution is constant, and the inequality holds trivially. It should be noted that the bound on |m=1M2ψm,jwmm=M2+1Mψm,jwm| can be further improved for the thick diffusive problem (Wang Citation2022).

Like DD, we use the discrete maximum principle to find the local bound of (ϕjϕjN). We define (56) M=1σa,jmax1jN|Lϕ,SDN(ϕjϕjN)|,(56) and (57) Φj=M±(ϕjϕjN).(57)

Then for 1jN, we have (58) Lϕ,SDN(Φj)=Lϕ,SDN(M)±Lϕ,SDN(ϕjϕjN),(58)

Since M is a constant, we find from EquationEquation (53) (59) Lϕ,SDN(M)=σa,jM.(59)

Substituting EquationEquation (56) into EquationEquation (59), we obtain (60) Lϕ,SDN(M)=max1jN1|Lϕ,SDN(ϕjϕjN)|.(60)

Then using EquationEquation (60) in EquationEquation (58), we obtain (61) Lϕ,SDN(Φj)0.(61)

Since the SD method is strictly positive, we always have Φj0, and then we can obtain |ϕjϕjN|M from EquationEquation (57). Using EquationEquation (55), we obtain the local error estimate as (62) |ϕjϕjN|M=1σa,jmax1jN|Lϕ,SDN(ϕjϕjN)|=Chjσjσa,jmax1jN|ϕ(ξj)|.(62)

Thus, the SD method is only first-order accurate, i.e., O(h).

Remark 4.1.

To derive the error estimate for the DD method, we have assumed the positivity of the numerical solution, which is however only true on sufficiently fine meshes. The SD method does not require such assumption since it is strictly positive. It should be noted that the error estimates obtained are sharp. There exist some previous results on the error estimates of finite difference methods for solving the SN neutron transport equation (Madsen Citation1972; Larsen and Miller Citation1980). Madsen performed truncation error analysis in terms of the angular flux by assuming that the exact solution has bounded third partial derivatives, and obtained the error estimate for the DD method, i.e., |ψm,jψm,jN|Chj2ψ. In the work of Larsen and Miller, they obtained the same error estimate for the 1D slab problem, of which the solution is infinitely differentiable. In our analysis, we are only concerned with the scalar flux. With the help of the discrete maximum principle, we can show that the error estimate of the scalar flux only requires the second derivatives to be bounded.

Remark 4.2.

The error estimate indicates that the accuracy of DD is not affected by the physical properties of the problem, which means that the method is an asymptotic preserving scheme (Larsen, Morel, and Miller Citation1987; Wang Citation2019). However, the SD method is not asymptotic preserving since its accuracy will deteriorate as the problem becomes thick and diffusive (i.e., σ and σa0) due to the factor σ/σa in the error estimate.

5. Numerical results

In this section, we demonstrate how the regularity of the exact solution will impact the numerical convergence rate by solving the SN neutron transport equation in its original integro-differential form, using the classic DD and SD methods. The model problem is a 1 cm × 1 cm square with the vacuum boundary condition. Thus, there will be no complication from the boundary condition. The S12 level-symmetric quadrature set is used for angular discretization.

We analyze the following four cases: Case 1: σ=1, σs=0; Case 2: σ=1, σs=0.8; Case 3: σ=10, σs=0, and Case 4: σ=10, σs=0.9. For all the cases, the external source Q=1. Cases 1 and 3 are pure absorption problems, while Case 3 is optically thicker. It is worth noting that the solutions are only determined by the external source for these two cases. Cases 2 and 4 include the scattering effects, while Case 4 is optically thicker and more diffusive. Both the scattering and external source contribute to the solution. The flux distribution calculated with DD for each case is shown in .

Figure 2. Flux distribution on the mesh 160×160.

Figure 2. Flux distribution on the mesh 160×160.

The flux L1 errors as a function of mesh size and the rates of convergence are summarized in and for DD and SD, respectively. In and are depicted the error distributions of DD and SD on the mesh 160×160. The reference solutions for all the cases are obtained on a very fine mesh (5120×5120) using the DD method.

Figure 3. Flux error distribution on the mesh 160×160 (DD).

Figure 3. Flux error distribution on the mesh 160×160 (DD).

Figure 4. Flux error distribution on the mesh 160×160 (SD).

Figure 4. Flux error distribution on the mesh 160×160 (SD).

Table 1. Scalar flux L1 errors and convergence rates of DD.

Table 2. Scalar flux L1 errors and convergence rates of SD.

For the DD, it is evident that the convergence rate decreases as the mesh is refined, and the errors are much larger at the boundary. The “noisier” distributions in Cases 1 and 2 are due to the ray effects of the discrete ordinates (SN) method, which are more pronounced in the optically thin problem. The convergence behavior is similar between the cases with and without the scattering, indicating that the source term plays a significant role in defining the irregularity of the solution. Cases 3 and 4 show the improved convergence rate as compared to Cases 1 and 2 because the exponential function eσ|xy| in the solution makes the kernel less singular for larger total cross section σ. In addition, Case 4 has a slightly better rate of convergence than Case 3 on fine meshes (e.g., 1.80 vs. 1.71 on 640×640), because the transport problem in Case 4 becomes more like an elliptic diffusion problem (Wang and Byambaakhuu Citation2021), which usually has better regularity. It is noted that that in Case 3 the convergence rate is only 1.51 on the coarse mesh. It is because for the pure absorption case, the DD becomes unstable when the mesh size is larger than 2μjσ. While for the SD method, the convergence rate is the first order for all the cases. By comparing Cases 4 and 3, the SD method has relatively large error for thick diffusive problems because of the factor σ/σa in the error estimate.

Remark 5.1.

The optimal error derived for the DD method is |ϕjϕjN|Chj2ϕ. As given by EquationEquation (28), the second derivative ϕ is bounded in the interior of the domain, while it behaves as ϕO(hj1) near the boundary. Therefore, it is expected that the convergence rate of the DD would decrease with refining the mesh, and asymptotically tend to O(hj). It is interesting to note that if the solution were sufficiently smooth (e.g., a manufactured smooth solution), the DD would maintain its second-order accuracy on any mesh size (Wang et al. Citation2019). While for the SD method, |ϕjϕjN|Chjϕ, where ϕO(1) or at most O(1+|loghj|) based on the theoretical results given by EquationEquation (28). From the numerical results, we observe that the SD has maintained the first-order convergence on all the meshes.

Remark 5.2.

The scattering does not appear to play a role in defining the smoothness of the solution. For the problem without the external source, if there exists a nonsmooth or anisotropic incoming flux on the boundary, the scattering may not be able to regularize the solution either, since the irregularity caused by the incoming flux, which is defined by the surface integral term of EquationEquation (3), has nothing to do with the scattering and the solution flux ϕ.

6. Conclusions

We have derived the integral kernel for the two-dimensional integral radiation transfer equation, and examined the differential properties of the integral kernel for fulfilling the boundedness conditions of Vainikko’s theorem. We applied the theorem to estimate the derivatives of the radiation transfer solution near the boundary of the domain. It is noted that the first derivative of the scalar flux ϕ(x) becomes unbounded when approaching the boundary. The derivatives of order k behave as ϱ(x)1k for k>1, where ϱ(x) is the distance to the boundary. We have derived the optimal error estimates in terms of the scalar flux for the DD and SD methods. The numerical results demonstrate that the irregularity of the exact transport solution will adversely reduce the rate of convergence of numerical methods. We are currently extending the analysis to the boundary integral transport problem in considering nonzero incoming boundary conditions and corner effects. In addition, it would be interesting to study the convergence behavior of weak numerical solutions.

Disclosure statement

No potential conflict of interest was reported by the authors.

References

  • Abramowitz, M., and I. A. Stegun. 1970. Handbook of mathematical functions: With formulas, graphs, and mathematical tables. Dover, New York.
  • Agoshkov, V. 1998. Boundary value problems for transport equations. Boston: Birkhauser.
  • Altaç, Z. 2007. Exact series expansions, recurrence relations, properties and integrals of the generalized exponential integral functions. J. Quant. Spectrosc. Radiat. Transf 104 (3):310–25. doi:10.1016/j.jqsrt.2006.09.002
  • Bell, G. J., and S. Glasstone. 1970. Nuclear reactor theory. New York: Van Nostrand Reinhold Company.
  • Case, K. M., and P. F. Zweifel. 1967. Reading, MA: Linear transport theory. Addison-Wesley Publishing Company, Inc.
  • de Azevedo, F. S., E. Sauter, P. H. A. Konzen, M. Thompson, and L. B. Barichello. 2018. Integral formulation and numerical simulations for the neutron transport equation in X-Y geometry. Ann. Nucl. Energy 112:735–47. doi:10.1016/j.anucene.2017.10.017
  • Egger, H., and M. Schlottbom. 2014. An Lp theory for stationary radiative transfer. Appl. Anal. 93 (6):1283–96. doi:10.1080/00036811.2013.826798
  • Germogenova, T. A. 1969. Local properties of the solution of the transport equation. Dokl. Akad. Nauk SSSR 187 (5):978–81.
  • Golse, F., P.-L. Lions, B. Perthame, and R. Sentis. 1988. Regularity of the Moments of the Solution of a Transport Equation. J. Funct. Anal. 76 (1):110–25. doi:10.1016/0022-1236(88)90051-1
  • Hennebach, E., P. Junghanns, and G. Vainikko. 1995. Weakly singular integral equations with operator-valued Kernels and an application to radiation transfer problems. Integr. Equ. Oper. Theory. 22 (1):37–64. doi:10.1007/BF01195489
  • Johnson, C., and J. Pitkaranta. 1983. Convergence of a fully discrete scheme for two-dimensional neutron transport. SIAM J. Math. Anal. 20 (5):951–66.
  • Larsen, E. W., and W. F. Miller. Jr., 1980. Convergence rates of spatial difference equations for the discrete-ordinates neutron transport equations in slab geometry. Nucl. Sci. Eng. 73 (1):76–83. doi:10.13182/NSE80-3
  • Larsen, E. W. 1982. Spatial Convergence Properties of the Diamond Difference Method in x, y Geometry. Nucl. Sci. Eng. 80 (4):710–713.
  • Larsen, E. W., J. E. Morel, and W. F. Miller. Jr., 1987. Asymptotic solutions of numerical transport problems in optically thick, diffusive regimes. J. Comput. Phys. 69 (2):283–324. doi:10.1016/0021-9991(87)90170-7
  • Latrach, K. 2001. Compactness results for transport equations and applications. Math. Models Methods Appl. Sci. 11 (07):1181–202. doi:10.1142/S021820250100129X
  • Latrach, K., and A. Zeghal. 2012. Existence results for a nonlinear transport equation in bounded geometry on L1-spaces. Appl. Math. Comput. 219 (3):1163–72. doi:10.1016/j.amc.2012.07.026
  • Lewis, E. E., and W. F. Miller. Jr., 1993. Computational methods of neutron transport. La Grange Park, Illinois: American Nuclear Society.
  • Madsen, N. K. 1972. Convergence of singular difference approximations for the discrete ordinate equations in x−y geometry. Math. Comput. 26 (117):45–50.
  • Mikhlin, S. G., and S. Prossdorf. 1986. Singular integral operators. Berlin: Springer-Verlag.
  • Mikhlin, S. G. 1965. Multidimensional singular integrals and integral equations. Oxford: Pergamon Press.
  • Miller, J., J. H. O’Riordan, and G. I. Shishkin. 2012. Fitted numerical methods for singular perturbation problems. Singapore: World Scientific.
  • Pitkäranta, J. 1980. Estimates for the derivatives of solutions to weakly singular fredholm integral equations. SIAM J. Math. Anal. 11 (6):952–68. doi:10.1137/0511085
  • Stynes, M., and D. Stynes. 2018. Convection-diffusion problems: An introduction to their analysis and numerical solution. Providence, RI: American Mathematical Society.
  • Vainikko, G. 1993. Multidimensional weakly singular integral equations. Berlin Heidelberg: Springer-Verlag.
  • Vladimirov, V. S. 1963. Mathematical problems in the one-velocity theory of particle transport. (Translated from, 61, 1961). Chalk River, Ontario: Atomic Energy of Canada Limited.
  • Wang, D. 2019. The asymptotic diffusion limit of numerical schemes for the SN transport equation. Nucl. Sci. Eng. 193 (12):1339–54. doi:10.1080/00295639.2019.1638660
  • Wang, D. 2022. Error analysis of numerical methods for thick diffusive neutron transport problems on Shishkin Mesh. 2022. Proceedings of International Conference on Physics of Reactors 2022 (PHYSOR 2022), Pittsburgh, PA, USA, May 15-20, 2022, 977986.
  • Wang, D., and T. Byambaakhuu. 2021. A new proof of the asymptotic diffusion limit of the SN neutron transport equation. Nucl. Sci. Eng. 195 (12):1347–58. doi:10.1080/00295639.2021.1924048
  • Wang, D., T. Byambaakhuu, S. Schunert, and Z. Wu. 2019. Solving the SN transport equation using high order Lax-Friedrichs WENO fast sweeping methods. Proceedings of International Conference on Mathematics and Computational Methods Applied to Nuclear Science and Engineering 2019 (M&C 2019), Portland, OR, USA, August 25–29, 2019, 61–72.
  • Wang, Y., and J. C. Ragusa. 2009. On the convergence of DGFEM applied to the discrete ordinates transport equation for structured and unstructured triangular meshes. Nucl. Sci. Eng. 163 (1):56–72. doi:10.13182/NSE08-72