9,670
Views
2
CrossRef citations to date
0
Altmetric
Technical Notes

A New Proof of the Asymptotic Diffusion Limit of the SN Neutron Transport Equation

ORCID Icon &
Pages 1347-1358 | Received 11 Oct 2020, Accepted 27 Apr 2021, Published online: 22 Jul 2021

Abstract

It has been well known that the analytic neutron transport solution tends to the analytic solution of a diffusion problem for optically thick systems with small absorption and source. The standard technique for proving the asymptotic diffusion limit is constructing an asymptotic power series of the neutron angular flux in small positive parameter ε, which is the ratio of a typical mean free path of a particle to a typical dimension of the problem domain. In this paper, first, we provide an analysis of the asymptotic properties of the SN transport eigenvalues. Then, we show that the analytical SN transport solution satisfies the diffusion equation in the asymptotic diffusion limit based on a recently obtained closed-form analytical solution to the one-dimensional monoenergetic SN neutron transport equation. The boundary conditions for the diffusion equation are discussed.

I. INTRODUCTION

The asymptotic analysis methodology originally developed by Larsen and others has proved to be a powerful technique to study transport problemsCitation1–5 and analyze numerical methods.Citation6–9 The essential idea of the asymptotic analysis method is to assume a solution (the flux) that depends on a small parameter ε by a simple asymptotic expansion: ψ=i=0ψiεi. Larsen et al. used the asymptotic analysis to show the transition from analytic transport to analytic diffusion for the SN transport equation.Citation6 Most recently, to study the asymptotical properties of numerical schemes for the SN transport equation, we proposed a different asymptotic analysis method that is based on the scaled Taylor expansion.Citation10

In this paper, we present a direct proof to show that the analytical solution of the SN transport equation satisfies the diffusion equation in the asymptotic diffusion limit. To facilitate the discussion, we first present a new closed-form analytical expression of the solution to the monoenergetic SN transport equation in slab geometry,Citation11 which is essential for the proof. Then, the proof will follow naturally as the asymptotic properties of the SN transport eigenvalues.

It should be noted that a number of analytical solutions and their numerical implementations have been proposed for the one-dimensional SN transport equation.Citation12–17 Our solution is essentially similar to those matrix exponential formulations.Citation14,Citation15,Citation18 A comprehensive review of the matrix-exponential formalism in radiative transfer can be found in the literature.Citation19 The SN transport equation is decoupled by eigen decomposition. In general, we can consider the eigenvalues and eigenvectors as known since they can be easily obtained using a modern software with arbitrarily high precision.Citation20 The decoupled set of linear first-order ordinary differential equations is solved analytically in matrix form based on the classical method of characteristics. One can find the detailed derivation in CitationRef. 11. Here, we briefly summarize the solution as follows.

The steady monoenergetic SN equation in slab geometry assuming isotropic scattering and constant neutron source is written as

(1) μddxΨ+ΣtΨ=Σs2WΨ+Q21,(1)

where

(2) Ψ=ψ1ψ2ψNT=angular flux vector(2)
(3) μ=μμN×N =\rm{cosine of the neutron direction relative to}xaxis,(3)

in which

(4) μ=diagμn>0,n=1,,N2(4)
(5) W=wwwwN×N\rm{ = quadrature weights},(5)

in which

(6) w=w1w2wN2w1w2wN2wN2w1w2wN2N2×N2(6)
(7) n=1N2wn=1(7)
(8) 1=111T(8)

Σt= total macroscopic cross section

Σs= macroscopic scattering cross section

Q= constant neutron source.

EquationEquation (1) is a system of coupled first-order differential equations. Matrices μ and W contain the nodes and weights of a symmetric quadrature set (e.g., Gauss-Legendre quadrature).

We can write EquationEq. (1) as

(9) ddxΨ+Σtμ1Ic2WΨ=q,(9)

where

(10) c=ΣsΣt=\rm{scattering ratio}(10)
and
(11) q=Q2μ11.(11)

The transport matrix Σtμ1Ic2W is diagonalizable, and we have

(12) Σtμ1Ic2W=RΛR1,(12)

where Λ is the diagonal matrix with the eigenvalues of the matrix Σtμ1Ic2W and R is the matrix with the columns consisting of eigenvectors. We can find R and Λ through eigen decomposition. If c=0, the eigenvalues and eigenvectors are simply Σtμ1 and the identity matrix I, respectively.

We write Λ as

(13) Λ=Λ+Λ,(13)

where

(14a) Λ+=diagλn,n=1,N2(14a)
and
(14b) Λ=diagλn,n=N2,N.(14b)

We will show in Sec. II that all the eigenvalues are real and that they occur in positive-negative pairs as

(15a) Λ+>0(15a)

and

(15b) Λ=Λ+<0.(15b)

The detailed solution process can be found in CitationRef. 11. We just give the final analytical solution for the angular flux as just

Ψ=Q21Σt1c1RexΛ+eLxΛR1Q21Σt1c1RY+0YL.(16)

Then, the scalar flux can be calculated as

ϕ=WTΨ=QΣt1cWTRexΛ+eLxΛR1Q21Σt1c1RY+0YL, (17)

where

(18a) W=w1w2wN2w1w2wN2T(18a)
(18b) Y+0YL=R11R12eLΛR21eLΛ+R221Ψ+0ΨLR12IeLΛR21IeLΛ+Λ+1b+Λ1b(18b)
(18c) b=b+b=R1q(18c)
(18d) Ψ+0\rm{ = incoming angular flux at}x=0(18d)
(18e) ΨL\rm{ = incoming angular flux at}x=L.(18e)

Remark: We derived the general solution in a straightforward way by combining the left and right incoming angular flux vectors into a single vector as shown in EquationEq. (18b). The final solution is an explicit closed-form expression of eigen exponential functions, and it applies for c0,1. When the scattering ratio c is extremely close to 1, there will be a loss of precision in the solution. To achieve high-precision solutions, one needs to use software with multiprecision capability.Citation20 Our solution does not apply for the case of conservative scattering (c=1), in which the smallest eigenvalues vanish and the transport matrix becomes singular. However, in numerical computation we can always calculate the conservative case with our solution by limiting c1. It is worth noting that an analytical approach to overcoming the difficulty with the conservative scattering is described by Efremenko et al.Citation19 In addition, the analytical solution can be extended to solve the anisotropic scattering case and handle heterogeneous geometry without iterating on surface fluxes.Citation21 It is also worth mentioning that a novel analytical nodal discrete ordinates method has recently been developed for solving the SN equation in two-dimensional Cartesian geometry based on the analytical solution derived for slab geometry.Citation22

The remainder of this paper is organized as follows. In Sec. II, we present an analysis of the eigenvalues of the SN equation in the asymptotic diffusion limit. Section III presents the proof of diffusion theory; i.e., the analytical solution of the SN equation asymptotically satisfies the diffusion equation in the diffusion limit. The boundary conditions (BCs) for the diffusion equation are discussed. A numerical example is given in Sec. IV to compare the difference between the SN transport solution and its diffusion counterpart. Finally, conclusions are given in Sec. V.

II. TRANSPORT EIGENVALUES IN THE ASYMPTOTIC LIMIT

Before we present the proof to show that the solution of the SN transport equation tends to that of the diffusion equation in the asymptotic diffusion limit, we need to investigate the asymptotic properties of the eigenvalues of the transport matrix Σtμ1Ic2W.

Let

(19) A=Σtμ1Ic2W.(19)

The matrix A can be rewritten as

(20) A=Σtμ100Σtμ1Ic2wc2wc2wIc2w=Σtμ1Σtc2μ1wΣtc2μ1wΣtc2μ1wΣtμ1+Σtc2μ1w,(20)

where μ, μ, W, and w are defined in EquationEqs. (3) through (Equation6), respectively. I is the identity matrix.

We can obtain the eigenvalues λ through the following determinant of the matrix:

(21) detAλI=\openup4ptΣtμ1Σtc2μ1wλIΣtc2μ1wΣtc2μ1wΣtμ1+Σtc2μ1wλI=0.(21)

The value of the determinant does not change by adding the second row to the first row:

(22) detAλI=\openup4ptΣtμ1λIΣtμ1λIΣtc2μ1wΣtμ1+Σtc2μ1wλI.(22)

The value of the determinant does not change by adding the first column to the second column:

(23) detAλI=\openup4ptΣtμ1λI2λIΣtc2μ1wΣtμ1+Σtcμ1wλI.(23)

The value of the determinant does not change by multiplying the second column with 12 and then adding it to the first column:

(24) detAλI=\openup4ptΣtμ12λIΣt2μ1+λ2IΣtμ1+Σtcμ1wλI.(24)

The value of the determinant does not change by adding the (12) multiple of the first row to the second row, and then we obtain

(25) detAλI=\openup4ptΣtμ12λIλ2IΣtμ1+Σtcμ1w.(25)

Since Σtμ1 is a nonsingular matrix, the determinant can be expressed as

(26) detAλI=detΣtμ1detΣtμ1+Σtcμ1w+λ2Σt1μ=detΣtμ12detI+cw+λ2Σt2μ2=0.(26)

Since detΣtμ10, we obtain

(27) detI+cw+λ2Σt2μ2=0.(27)

Using the Sherman-Morrison formula,Citation23 we obtain

(28) detI+cw+λ2Σt2μ2=detλ2Σt2μ2I+c1N2×1w1×N2=detλ2Σt2μ2I1+cw1×N2λ2Σt2μ2I11N2×1=0,(28)

where

(29) w1×N2=w1w2wN/2(29)

and

(30) 1N2×1=111.(30)

If c=0, which is the purely absorbing case, then we have

(31) detλ2Σt2μ2I=0,(31)

and thus,

(32) λi=Σtμi,i=1,,N2.(32)

Here, we are interested in the case of c0. Now that detλ2Σt2μ2I0, we have

(33) 1+cw1×N2λ2Σt2μ2I11N2×1=0.(33)

Then, we have

(34) w1×N2λ2Σt2μ2I11N2×1=i=1N/2wi1λ2μi2Σt2=1c.(34)

The characteristic equation, EquationEq. (34), implies that all eigenvalues λ are real and occur in positive-negative pairs for the case of c<1. If the roots of λ were imaginary numbers, then 1λ2μi2Σt2>1, and we would have i=1N/2wi1λ2μi2Σt2<1 since i=1N/2wi=1, which is contradictory with 1c>1. The eigenvalues degenerate to zero in the limit of c=1, and they become imaginary roots for c>1. Here, we assume a symmetric quadrature with weights wi>0 and μi<1 for our argument. It should be noted that Davison deduced the same result for the monoenergetic SN transport equation with isotropic scattering by using a heuristic approach.Citation24 A general proof on the spectrum of the transport operator was given by Kuscer and Vidav.Citation25

If λ2μi2Σt2<1 (it is shown later that the condition is satisfied for the smallest eigenvalue in the asymptotic diffusion limit), the above summation term can be expanded in a series of polynomials of μi:

(35) wi1λ2μi2Σt2=wi1+λ2Σt2μi2+λ2Σt22μi4+λ2Σt23μi6+.(35)

It is known that the weights wi and nodes μi of the Gauss-Legendre quadrature can allow the quadrature rule to integrate degree 2N1 polynomials exactly. Therefore, for N sufficiently large, the summation in EquationEq. (34) is almost exactly equal to the following integral:

(36) i=1N/2wi1λ2μi2Σt20111λ2μ2Σt2dμ.(36)

Then, we have

(37) 0111λ2μ2Σt2dμ=tanh1λΣtλΣt=1c.(37)

It is interesting to note that we arrive at the same formula as derived for the integrodifferential transport equation based on the singular eigenfunction method (separation of variables).Citation26

Now, we introduce the following scaling of cross sections:

(38) Σt=Σt0ε(38)

and

(39) Σa=εΣa0,(39)

which implies

(40) Σs=ΣtΣa=Σt0εεΣa0(40)

and

(41) c=ΣsΣt=1ε2Σa0Σt0,(41)

where ε is a small positive parameter and Σt0 and Σa0 are constants.

Substituting the scaling EquationEqs. (38) and (Equation41) into EquationEq. (37) and letting ε0, we can find the two roots for λ, which are denoted by λ1:

(42) λ1±3Σt0Σa0=±3ΣtΣa.(42)

It shows that in the asymptotic diffusion limit, the absolute value of λ1 tends to a constant, which corresponds to the inverse of the diffusion length in reactor physics. Now, we verify λ12μi2Σt2<1. Since μi1, we only need to have λ12Σt2<1. We have λ12Σt2=3Σt0Σa0Σt0/ε2=3ε2Σa0Σt01, which goes to zero quickly as ε0.

Next, we investigate the properties of other eigenvalues in the asymptotic diffusion limit. Since λ1 are the two eigenvalues of the transport matrix, they should satisfy EquationEq. (34) as

(43) i=1N/2wi1λ12μi2Σt2=1c.(43)

Subtracting EquationEq. (43) from EquationEq. (34), we obtain

(44) i=1N/2wi1λ2μi2Σt2wi1λ12μi2Σt2=i=1N/2wiμi2λ2λ121λ2μi2Σt21λ12μi2Σt2=0.(44)

Since λ2λ120, we have

(45) i=1N/2wiμi21λ2μi2Σt21λ12μi2Σt2=0.(45)

Applying the cross-section scaling by substituting EquationEq. (38) into EquationEq. (45), we obtain

(46) i=1N/2wiμi21ε2λ2μi2Σt021ε2λ12μi2Σt02=0.(46)

As ε0, 1ε2λ12μi2Σt021. Then, EquationEq. (46) can be simplified as

(47) i=1N/2wiμi21ε2λ2μi2Σt02=0.(47)

From EquationEq. (47), we can see that if ε2λ2μi2Σt020, then i=1N/2wiμi20, which is contradictory to i=1N/2wiμi2=01μ2dμ=13. Thus, we must have

(48) ελiΣt0=O1(48)

or

(49) λi=OΣt0ε,(49)

where i=2,,N/2.

Take the S12 Gauss Legendre quadrature as an example. Six positive eigenvalues are given in , for ε=0.01, 0.001 and 0.0001, respectively. It shows that λϱ when ε0. The eigenvalue λ1 is almost the same as the theoretical one calculated by EquationEq. (37), and it tends to the inverse of the diffusion length as well. It can also be seen that all the other eigenvalues increase inversely with ε on the order of O1ε, which confirms our analysis in Sec. II.

Remark: The eigenvalues of the SN transport equation in the diffusion limit are separated into two groups. The first group contains only two eigenvalues with the same absolute value, which is limited to a constant (i.e., the inverse of the diffusion length), while the second group contains all the rest of the eigenvalues, which also appear in positive/negative pairs and are proportional to 1/ε. The properties of these discrete eigenvalues in the diffusion limit are indeed consistent with those for the continuous integrodifferential transport equation.

TABLE I Eigenvalues of the S12 Gauss Legendre Quadrature*

III. PROOF OF DIFFUSION THEORY

Given the asymptotic properties of the transport eigenvalues, now we complete the final proof for which the analytical SN solution satisfies the diffusion equation in the diffusion limit. The analytical scalar flux given by EquationEq. (17) allows us to show that the asymptotic flux satisfies a diffusion solution. We differentiate the scalar flux as

(50) dϕdx=WTRΛ+exΛ+ΛeLxΛR1Q21Σt1c1RY+0YL(50)

and

(51) d2ϕdx2=WTRΛ+2exΛ+Λ2eLxΛR1Q21Σt1c1RY+0YL=WTRΛ+2Λ2exΛ+eLxΛR1Q21Σt1c1RY+0YL.(51)(51)

Multiplying EquationEq. (17) with ϱ2, we have

(52) ϱ2ϕ=ϱ2QΣt1cϱ2WTRexΛ+eLxΛR1Q21Σt1c1RY+0YL,(52)(52)

where

(53) ϱ=3ΣtΣa=Σt31c(53)

with

(54) Σa=Σt1c.(54)

It can be seen that the parameter ϱ is actually the inverse of the diffusion length.

Subtracting EquationEq. (52) from EquationEq. (51), we obtain

(55) d2ϕdx2ϱ2ϕ=WTRΛ+2Λ2exΛ+eLxΛR1Q21Σt1c1RY+0YLϱ2QΣt1c+ϱ2WTRexΛ+eLxΛR1Q21Σt1c1RY+0YL.(55)

Then, we can rewrite EquationEq. (55) as

(56) d2ϕdx2ϱ2ϕ=Aϱ2QΣt1c,(56)

where

(57) A=BR1Q21Σt1c1RY+0YL(57)

with

(58) B=ϱ2WTRWTRΛ+2Λ2exΛ+eLxΛ.(58)

The eigenvalues are given in diagonal matrix (EquationEq. 59):

(59) Λ=Λ+Λ=λ1λN/2λ1λN/2,(59)

where λ1<λ2<<λN/2. Note that we use a symmetric quadrature set for our analysis (e.g., Gauss-Legendre). It is shown in Sec. II that the eigenvalues in the diffusion limit are given as follows:

(60) λ1ϱ=3Σt0Σa0(60)

and

(61) λiOΣt0ε,i=2,,N2.(61)

Thus, we have the following for x away from the boundary (i.e., x0,L):

(62) eλix0, eλiLx0,λi2eλix0,andλi2eλiLx0(62)

for i=2,,N2.

Then, EquationEq. (58) can be simplified as follows:

(63) B=ϱ2WTRexΛ+eLxΛWTRΛ+2exΛ+Λ2eLxΛϱ2WTR\openup1pteλ1x0eλ1Lx0WTR\openup1ptλ12eλ1x0λ12eλ1Lx0=ϱ2WTRλ12WTR\openup1pteλ1x0eλ1Lx0(63)

Since λ1ϱ, we obtain B0, and therefore, A0. Then, EquationEq. (56) can be simplified as

(64) d2ϕdx2ϱ2ϕ=Qϱ2Σt1c.(64)

Substituting ϱ=3Σt0Σa0, Q=Q0ε (source scaling), c=1ε2Σa0Σt0, and Σt=Σt0ε into EquationEq. (64), we obtain inhomogeneous diffusion equation (65):

(65) 13Σt0d2ϕdx2+Σa0ϕ=Q0.(65)

It is concluded that the analytical solution of the SN transport equation satisfies the diffusion solution in the asymptotic diffusion limit.

Now, we neglect the terms containing eλixandeλiLx, i=2,,N2 in EquationEq. (17), to find the BCs for the diffusion equation as

(66a) ϕ0=QΣt1cWTRCDR1Q21Σt1c1RY+0YL(66a)

and

(66b) ϕL=QΣt1cWTRDCR1Q21Σt1c1RY+0YL,(66b)

where

(66c) Y+0YL=R11R12DR21DR221Ψ+0ΨLR12IDR21IDΛ+1b+Λ1b,(66c)
(66d) C=10N/2×N/2,(66d)

and

(66e) D=eλ1L0N/2×N/2.(66e)

It can be seen that the diffusion BCs depend on the transport BCs on both sides of the slab. Only when the slab thickness becomes relatively large will the influence from the opposite side be negligible.

By employing the boundary layer analysis, Habetler and Matkowsky derived the following diffusion BCs for slab geometryCitation2:

(67a) ϕ0=201Wμfμdμ(67a)

and

(67b) ϕL=201Wμgμdμ,(67b)

where

fμ= incoming angular flux at x=0

gμ= incoming angular flux at x=L

(67c) Wμ=μXμ01sXsds10.956μ+1.565μ2.(67c)

Note that Xμ is tabulated for 0μ1 in CitationRef. 27. The function Wμ is smooth and well approximated by the polynomialCitation7

(68) W˜μ=0.956μ+1.565μ2,(68)

where

sup0μ1WμW˜μ=0.0035.

It should be pointed out that the diffusion BCs of Habetler and Matkowsky are valid only for half-space problems with the external source Q=0 and that they would be inappropriate for a finite slab since the derived BCs do not account for the effects from the transport BC on the other side of the slab as shown in EquationEqs. (66a) and (Equation66b) for the SN case.

For a half-space problem with the external source Q=0, EquationEq. (66a) yields the diffusion BC at x=0 as

(69) ϕ0=WTRC0R11R221Ψ+00,(69)

where Ψ+0 is the incoming angular flux at x=0 for the transport equation.

Now, we compare the SN diffusion BCs with those given by Habetler and Matkowsky for a half-space test problem with the incoming flux fμ=1 and fμ=μ, respectively. The results are summarized in . It can be seen that when ε tends to zero (i.e., diffusion limit), the SN diffusion BCs tend to those of Habetler and Matkowsky. However, for less diffusive cases (e.g., ε<0.001), the SN diffusion BCs would be better than those of Habetler and Matkowsky because their BCs were derived in the limit of ε0. In addition, it is interesting to notice that there are almost no appreciable differences between S16 and S512.

Remark: Larsen constructed an asymptotic diffusion solution from Case’s analytic solution for a simple source-free (i.e., the homogeneous transport equation with Q=0), half-space transport problem.Citation4 A similar result was presented earlier by Case and Zweifel.Citation27 Most recently, we extended diffusion theory for inhomogeneous transport problems (Q0) in slab geometry based on Case’s analytic solution.Citation28 A direct proof for the stationary transport problem in the whole space is reported in CitationRef. 29, in which an analytical solution is obtained for an isotropic point source by using Fourier transformation.

TABLE II Comparison of Diffusion BCs*

IV. NUMERICAL RESULTS

Here, we give a numerical example to compare the analytical transport solutions with their diffusion solutions. The model problem is a slab problem. The Gauss-Legendre S64 quadrature set is used for angular discretization. The specifications of the problem are given as

L=1, h = 0.1, Σt=1ε,

Σs=1ε0.8ε, Q=ε

BCs: Ψ+0=\openup2pt50000032×1 at x=0, and Ψ+L=\openup2pt00100032×1 at x=1,

where L is the slab thickness and h is the reference mesh size in units of centimeter; Σt and Σs are in units of inverse centimeter. The problem becomes thick and diffusive as ε decreases.

The SN analytical solution is obtained using EquationEq. (17). The analytical solution of the asymptotic diffusion equation is given as

(70) ϕx=QΣt1c1eϱLeϱL\openup2pteϱLx+eϱLxeϱxeϱxTQΣt1c11ϕ0ϕL,(70)

where

ϱ=Σt31c
Σt=1 cm1
Σa=0.8 cm1
Q=1 cm1.

The diffusion BCs ϕ0 and ϕL are determined by EquationEqs. (66a) and Equation(66b).

shows the results of the analytical solution and diffusion solution. plots the scalar flux for ε=0.01 showing an excellent agreement between the two solutions in the interior region of the slab. The boundary layers develop in the transport solution on both sides of the slab due to the anisotropic BCs. shows that the flux L1 error (transport – diffusion) decreases with ε. For the case of ε=0.001, the flux L1 error is on the order of 106 in the interior region. This model problem demonstrates that the SN analytical solution asymptotically tends to the diffusion solution in the diffusion limit, and the diffusion BCs calculated by EquationEqs. (66a) and (Equation66b) give a very accurate diffusion solution. However, the diffusion BCs proposed by Habetler and Matkowsky would give large differences for this case.

Fig. 1. Analytical solution vs. diffusion solution

Fig. 1. Analytical solution vs. diffusion solution

V. CONCLUSIONS

We have presented a new proof of the asymptotic diffusion limit of the SN transport equation by directly showing that the SN analytical solution satisfies the diffusion equation in the asymptotic diffusion limit. The proof is made possible by showing that the smallest eigenvalue of the transport matrix tends to a constant (which is the inverse of the diffusion length), while other eigenvalues increase on the order of O1/ε, and thus, their contribution to the diffusion solution in the interior of the domain can be neglected. It is natural to find the BCs for the diffusion equation when the analytical transport solution is available. It is noted that there is a deficiency in the BCs derived by Habetler and Matkowsky based on the boundary layer analysis technique. In the strict sense, they are valid only for half-space source-free transport problems in the diffusion limit (i.e., ε = 0) and would be inappropriate for inhomogeneous problems in a finite domain like slab geometry. We have demonstrated with a numerical example that the SN analytical solution asymptotically tends to the diffusion solution.

Acknowledgments

We thank the anonymous reviewers whose comments/suggestions helped improve and clarify this manuscript.

References