728
Views
0
CrossRef citations to date
0
Altmetric
Research Articles

An Inline Burnup Algorithm

ORCID Icon, ORCID Icon & ORCID Icon
Pages 1681-1699 | Received 25 Apr 2022, Accepted 22 Jul 2022, Published online: 04 Oct 2022

Abstract

Inline algorithms have been proposed for coupling Monte Carlo neutron transport solvers with several other physics, such as xenon and iodine densities and thermal hydraulics. This paper proposes a new inline algorithm that can be applied to burnup calculations. The algorithm is a modification of the predictor-corrector method, where the corrector-step nuclide densities are converged simultaneously with the fission source. This could, in principle, obviate the need for two full neutronics solutions per time-step while still allowing the accuracy of predictor-corrector methods with improved stability. This paper describes the algorithm and demonstrates its stability properties through a Fourier analysis. Although not unconditionally stable, judicious use of batching and relaxation are shown to greatly improve the algorithm’s stability properties in realistic systems.

I. INTRODUCTION

Monte Carlo neutron transport solvers are increasingly common in research and in industry for resolving the neutronic behavior of nuclear reactor cores. This has been enabled by a swell in the availability of computing power over time. Although it is now conceivable to resolve neutronics in isolation, the computational expense of Monte Carlo solvers is such that coupling it with other important physics, such as thermal hydraulics or fuel mechanics, remains challenging due to the need to perform several Monte Carlo calculations to achieve convergence between the solvers. This common strategy for handling multiphysics calculations is known as fixed-point iteration.

Inline algorithms have recently become popular as a means of avoiding the computational expense of fixed-point iteration in converging neutronics with other physics. The logic of such algorithms is that they iteratively pass information from the Monte Carlo solver to the other physics solvers during Monte Carlo’s inactive cycles, i.e., before the neutronics solution is fully converged. These other physics solvers then update the macroscopic cross sections in the neutronics problem (either by modifying densities or temperatures) and the inactive cycles continue with this periodic data exchange. This makes intuitive sense in that it is undesirable to fully converge a neutronics solution when (due to inevitable changes in other physics) the macroscopic cross sections that describe the geometry are known to be erroneous. This strategy has been applied to converging Monte Carlo with xenon,Citation1 thermal hydraulics,Citation2,Citation3 and with both simultaneously.Citation4 Analogous strategies are used and described in coupling deterministic transport with other physics, with the accompanying analysis highlighting the additional numerical stability and convergence acceleration that this provides.Citation5–10

Isotopic depletion, or burnup, is another vital component of reactor physics, governing a reactor’s cycle length, the evolution of its power profile, and its spent fuel inventory, among other quantities of interest. The natural coupling between neutronics and depletion has neutronics generating reaction rates that the depletion solver uses to evolve the geometry’s isotopic composition forward by some chosen time step. Additional sophistication can occur, such as storing and using multiple transport solutions across several time steps to increase the order of accuracy of the time-stepping schemes. This is where much of the work on coupling neutronics and depletion has been to date.Citation11–14 Again, due to the expense of the Monte Carlo solution, it is desirable to minimize the number of neutronics solutions that must be found when considering the depletion of a reactor geometry.

In larger geometries, however, it has been found that numerical instability can also result from this coupling between neutronics and depletion.Citation15–19 This spurious behavior can often be avoided by removing the particularly troublesome depletion chain of 135Xe from the system and converging its nuclide density distribution separately,Citation20 although this is no guarantee of unconditional stability. Other means of ensuring stability introduce the equivalent of an implicit time-stepping scheme, where an iteration takes place between neutronics and depletion, invariably with some form of relaxation.Citation19,Citation21–26 Such schemes unfortunately impose additional undesirable computational expense.

Given the numerical stability and relatively low computational expense of other inline algorithms, it appears natural to propose applying it to depletion to overcome the drawbacks of the current implicit methods. To the authors’ knowledge, this was first proposed in the literature by Cosgrove,Citation27 although a version of the algorithm has actually existed in the Serpent Monte Carlo code as of version 2.1.31, albeit of a slightly different implementation than is considered here.Citation28

This paper will first describe the inline algorithm before performing a stability analysis to demonstrate its beneficial numerical properties.

II. BURNUP ALGORITHMS

Before describing the new algorithm, a brief discussion of the basics of neutronics-depletion coupling is warranted.

Neutronics is governed by the neutron transport equation. In reactor analysis, this tends to be the eigenvalue form of the transport equation:

(1) Ωψ+Σtψ=χ4πk0dE νΣf4πdΩ ψ+0dE 4πdΩ Σs(E E,Ω Ω)ψ,(1)

where

Ω ==

neutron direction of flight

ψ ==

angular neutron flux

Σr,r{t,f,s} ==

macroscopic cross sections for the total, fission, and scattering reactions

χ ==

fission neutron energy spectrum

k ==

criticality eigenvalue

E ==

neutron energy

ν ==

average number of neutrons produced per fission.

As well as the eigenvalue, the quantities of interest from the solution of the transport equation tend to be the reaction rates of neutrons with isotopes present in the system, either as a whole or over some particular volume. These are microscopic reaction rates averaged over a given volume of space of uniform composition where depletion calculations are concerned:

(2) σr,ψ=1V0dEσr4πdΩVdVψ,(2)

where σr is the microscopic cross section for reaction channel r, and V is the volume of space of interest. Instead of obtaining microscopic reaction rates, neutronics codes may obtain one-group microscopic cross sections and flux amplitudes that can be multiplied together to recover Eq. (2):

(3a) σr=0dEσr4πdΩVdVψ0dE4πdΩVdVψ(3a)

and

(3b) ψ=1V0dE4πdΩVdVψ.(3b)

The evolution of the density of a nuclide with index i, Ni, with time t is given as

(4) dNidt=jγijσf,ψ+lijσc,ψ+λijδij(λj+σa,ψ)Nj,(4)

where

γij ==

fission yield of nuclide i on fissioning nuclide j

lij ==

branching ratio producing nuclide i when nuclide j undergoes neutron capture

λij ==

decay constant for nuclide j decaying into nuclide i

δij ==

Kronecker delta.

For a system of nuclides (fission reactor neutronics/depletion tools tend to track at least hundredsCitation29,Citation30), one has the Bateman equation governing their evolutionCitation31:

(5) dNdt=AN,(5)

where N is now a vector of nuclides and A is the burnup matrix with the element Aij describing the production or loss terms of nuclide i from nuclide j. The formal solution of the Bateman equation for a constant burnup matrix after a time step of length Δt from an initial nuclide density N0 is the matrix exponential:

(6) N(Δt)=exp(ΔtA)N0.(6)

Evaluating this matrix exponential is difficult,Citation32 although for burnup matrices this can be accurately done using the Chebyshev rational approximation methodCitation33,Citation34 (CRAM). Other accurate methods of solving the Bateman equation include using more familiar Ordinary Differential Equation (ODE) solvers.Citation29,Citation35 The Bateman equation is valid in a homogeneous region, whereas in reactor problems, one is generally interested in how the many different regions of the problem evolve. Thus burnup problems tend to discretize the reactor geometry into many separate, homogeneous burnable regions in which reaction rates are obtained and where the Bateman equation is solved separately. For large, finely discretized reactor geometries, this may impose a substantial computer memory burden unless care is taken.Citation29 Although relatively unexplored, a spatially continuous approach to depletion has also been proposed and preliminarily investigated.Citation36

The fundamental problem in numerical burnup calculations is that the burnup matrix is not constant in time. It is a function of the neutron flux, which is itself dependent upon the evolving nuclide density distribution in the system through the macroscopic cross sections appearing in EquationEq. (1). That is,

(7) dNdt=A(ψ)N.(7)

Thus, the burnup problem is nonlinear.

The simplest approach to coupling neutronics and depletion and resolving this nonlinearity is the explicit Euler method (sometimes known as the predictor method). Following a neutronics solution and obtaining microscopic reaction rates across the geometry, the Bateman equation is evaluated for each region for a chosen time-step length Δt and the process is repeated until the end of the depletion period of interest As well as possessing a low order of accuracy, necessitating short time steps, the explicit Euler scheme is only conditionally stable, even for simple systems.Citation18

A slightly more sophisticated time-stepping algorithm is the predictor-corrector method, which enjoys a higher-order accuracy in time.Citation14 There are several variants of this scheme with slightly different properties; here we consider what Isotalo refers to as the Constant Extrapolation/Linear Interpolation (CE/LI) method implemented in Serpent,Citation28 MC21 (CitationRef. 37), and OpenMC (CitationRef. 30). This begins with the explicit Euler method, where neutronics is evaluated at some time point with index n, known as the beginning of step or BOS, and the resulting burnup matrix An is used to generate a nuclide density distribution at a time point with index n+1 at Δt in the future, known as the end of step or EOS. This is the predictor step, with its depletion described by the following equation:

(8) Nn+1=expΔtAnNn.(8)

Another neutronics solution is evaluated at this future time and reaction rates are obtained, allowing for the construction of a second burnup matrix An+1. Finally, depletion is performed again to generate the nuclide density distribution to be used at the next time step, but using a time-averaged burnup matrix:

(9) Nn+1=expΔtAn+An+12Nn.(9)

The CE/LI method is shown in Algorithm 1. In the algorithm, ψ(Nn) denotes the evaluation of a transport solution given the nuclide density distribution Nn. Unfortunately, the predictor-corrector method, too, is only conditionally stable for simple systems,Citation16,Citation17,Citation19 necessitating a workaround such as handling xenon separatelyCitation20 or implicit methods.

Algorithm 1: Predictor-Corrector Method

Input: N0,nsteps,Δt

for n=0,,nsteps1 do

 | /* Evaluate BOS transport solution      */

 | An ψ Nn

 | /* Perform predictor depletion        */

 | Nn+1exp[AnΔtn]Nn

 | /* Evaluate EOS transport solution      */

 | An+1 ψ (Nn+1)

 | A12An+An+1

 | /* Perform corrector depletion        */

 | |Nn+1exp[AΔtn]Nn

end

 | |||

All implicit methods work by iterating between neutronics and depletion at the EOS, each of which includes an under relaxation to stabilize the iteration, which is applied either to the reaction rates or the nuclide densities. The first of these was the Stochastic Implicit Euler (SIE) method,Citation21 although it is more akin to a backward Euler method due to iterating with only the EOS burnup matrix. This results in it only being first-order accurate. The stochastic in SIE comes from using the Stochastic Approximation to decide the relaxation factor applied.Citation38 A higher-order method that is equivalent to an implicit CE/LI scheme is the Stochastic Implicit Midpoint method.Citation22 This method begins with a slightly modified version of EquationEq. (8):

(10) Nn+1(0)=expΔtAnNn,(10)

where Nn+1(0) is the initial n+1’th nuclide density vector to begin the fixed-point iteration. On the j+1’th iteration, a transport solution is evaluated from Nn+1(j) to generate the corresponding burnup matrix An+1(j). This is then used in the following corrector depletion equation:

(11) Nn+1(j+1)=αjexpΔtAn+An+1(j)2Nn+(1αj)Nn+1(j),(11)

where αj is the relaxation factor, which may vary with the iteration or remain constant.Citation26 Higher-order implicit methods have also been described, where the predictor and corrector depletion calculations use additional information from previous transport solutions and divide the depletion interval into substeps where the coefficients in front of the burnup matrices are varied.Citation23–25 The implicit mid-point scheme is shown in Algorithm 2.

Algorithm 2: Implicit Midpoint Method

Input: N0,nsteps,Δt,jmax,α

for n=0,,nsteps1do

   /* Evaluate BOS transport solution */

   AnψNn

   /* Perform predictor depletion */

   Nn+1(0)expAnΔtnNn

   /* Iterate on the corrector step */

   for j=0,,jmax1do

     /* Evaluate EOS transport solution */

     An+1jψNn+1j

     A12An+An+1j

     /* Perform corrector depletion with relaxation */

     Nn+1j+1αjexpAΔtnNn+1αjNn+1j

   end

     Nn+1Nn+1jmax

end

Inline algorithms efficiently converge a nonlinearity between neutronics and other physics; in depletion, that nonlinearity is normally handled during the fixed-point iteration on the corrector step, corresponding to EquationEq. (11). The proposed inline algorithm, as with all of the previous depletion algorithms, begins with a standard predictor step to generate the initial EOS nuclide density vector, corresponding to EquationEq. (10). During one or several inactive cycles, denoted with index c, of the EOS neutronics, reaction rates are scored to construct a burnup matrix for each burnable region An+1(c). At the conclusion of this batch of cycles, the nuclide density vectors in each region are updated according to

(12) Nn+1(c)=expΔtAn+An+1(c)2Nn.(12)

Provided the effects of statistical noise are not deleterious, this iteration should hopefully converge the corrector neutronics and densities, as occurs with thermal hydraulics and xenon. As with other implicit methods in burnup calculations, this corrector could be replaced with higher-order versions using substepping and quadratic interpolation. In principle, if convergence is reached, the active cycles of the EOS could be foregone entirely; a single transport solution could act as the BOS solution during the active cycles and as the EOS during the inactive. The only exception to this is during the first transport solution where the fission source must be converged before scoring reaction rates for the predictor. The inline scheme is shown in Algorithm 3. In this algorithm, V has been included as an input to more explicitly account for the geometry and the uniform sampling of the initial fission source across the geometry, shown as F(0)unif(V). In principle, other tallying strategies might be used, other than accumulating tallies across several batches before flushing them and beginning anew. For example, one might employ the moving-window approach suggested by CitationRef. 3 where the corrector update is performed every cycle but using reaction rates accumulated over several previous cycles.

Algorithm 3: Inline Predictor-Corrector Algorithm

Input: N0,nsteps,Δt,ninactive,nactive,V

for n=0,,nsteps1 do

  if n=0 then

    /* The initial BOS neutronics calculation has no feedback */

    Ln,vf,nN0

    κ01

    F0unifV

    for c=0,,ninactive1 do

      ψc+11κcLn1F(c)

      Fc+1vf,nψc+1dΩdE

      κc+1κcFc+1dVFcdV

    end

  else

    /* Converge neutronics with nonlinear feedback during EOS */

    κ01

    F0unifV

    AnΛ

    for c=0,,ninactive1 do

      ψc+11κcLc1Fc

      Fc+1vf,n(c)ψc+1dΩdE

      κc+1κcFc+1dVFcdV

      An+=σ,ψc+1

      A12An1+An

      Nnc+1expAΔtn1Nn1

      Lnc+1,vΣf,nc+1Nnc+1

      AnΛ

    end

    NnNnc+1

    Ln,vΣf,nLnc+1,vΣf,nc+1

  end

  /* Generate BOS burn-up matrix during active cycles */

  AnΛ

  for c=0,,nactive1 do

    ψc+11κcLn1F(c)

    An+=σ,ψc+1

    Fc+1vf,nψc+1dΩdE

    κc+1κcFc+1dVFcdV

  end

  Nn+1expAnΔtnNn

end

III. STABILITY ANALYSIS

The inline depletion scheme is now sufficiently defined to permit a stability analysis. A stability analysis (or Fourier or Von Neumann stability analysis) is used to analyze the behavior of iterative methods or time-stepping schemes across computational physics. Where neutronics and depletion are concerned, several papers on stability analysis have been published to dateCitation18,Citation19,Citation39–41 and the analysis to follow descends directly from them.

A stability analysis consists of, first, applying the equations describing the numerical method to a simple system where their behavior can be determined analytically. The solution of the system on iteration or evolution in time may then be expressed in terms of this known initial solution plus perturbation terms. Provided the system is sufficiently simple, these perturbation terms can be decomposed into Fourier modes. By linearizing the governing equations, one can obtain expressions for how the Fourier modes during one time step are related to those at the next, either by a scalar or matrix. The magnitude of the scalar or the spectral radius of the matrix determines the behavior of a particular Fourier mode across time steps; if this value is greater than one, then that Fourier mode will grow. Growth in a particular Fourier mode implies that the given algorithm is numerically unstable.

The equations used for the inline scheme must be chosen. These are those describing neutronics, predictor depletion, and corrector depletion. For neutronics, rather than EquationEq. (1), the one-dimensional, monoenergetic, neutron diffusion equation in slab geometry will be used in k-eigenvalue form, assuming isotropic scattering:

(13) xDx+Σaϕ=1kνΣfϕ,(13)

where

x[0,L] ==

spatial variable across a domain of length L

D ==

neutron diffusion coefficient

Σa ==

macroscopic absorption cross section

ϕ ==

scalar neutron flux

ν ==

average number of neutrons per fission

Σf ==

macroscopic fission cross section

k ==

criticality eigenvalue.

Provided there is no leakage, the criticality eigenvalue is defined as

(14) k=0LνΣfϕdx0LΣaϕdx.(14)

This expression for k becomes slightly more complicated when accounting for how the flux and composition (in our particular circumstances) evolve during the source iterations. Including the dependence of both on the source iteration index c, this becomes

(15) k(c)=0LνΣf(c1)ϕ(c)dx0LΣa(c1)ϕ(c)dx.(15)

The approximations applied in going from EquationEq. (1) to EquationEq. (13) are quite severe. Some of these are justifiable. Problems in which depletion instabilities occur tend to be large thermal systems dominated by scattering where the diffusion approximation is reasonable, indeed a transport equation can be used instead, but no substantial differences have been observed in problems considered so far.Citation39 This is expected, given that numerical instabilities in burnup arise from the low-frequency Fourier modes and transport and diffusion agree in the low-frequency limit.

As scattering is isotropic, the diffusion coefficient in EquationEq. (13) is

(16) D=13Σt.(16)

The cross sections featured in EquationEq. (13) are the inner product of the nuclide density vector (which is taken to be continuous in x) and the corresponding microscopic cross-section vector, such that

(17a) Σa=σaTN,(17a)
(17b) Σt=σtTN,(17b)

and

(17c) νΣf=(νσf)TN.(17c)

EquationEquation (13) requires a normalization condition to determine the amplitude of the neutron flux; reactor power is a common choice in depletion calculations. In one dimension, this normalization would be to a fixed power per unit area P such that

(18) P=0LκΣfϕdx,(18)

where κ is the average energy release per fission. One must also specify boundary conditions for EquationEq. (13). These may be either reflective or periodic to enable the subsequent Fourier decomposition, with the choice determining the allowable Fourier frequencies.

Two slightly different versions of the diffusion equation feature in the following analysis, accounting for both the change in macroscopic cross sections during burnup and for the equation’s solution method. First, when solving the diffusion equation to full convergence at time point n, it may be written as

(19) xDnx+Σa,nϕn=1knνΣf,nϕn.(19)

This corresponds to the equation solved when the nuclide density vector is Nn. When accounting for a step in the source iteration algorithm (corresponding to the Monte Carlo method of generationsCitation42), prior to convergence of the neutronics solution, a single iteration is instead written as

(20) xDn(c)x+Σa,n(c)ϕn(c+1)=1kn(c)νΣf,n(c1)ϕn(c).(20)

Although not standard to source iteration, the dependence of the cross sections and diffusion coefficients on the iteration index has been included, as it is induced by the change in nuclide densities caused by the inline algorithm. Note that the fission cross section depends on one iteration previous. This is to account for how fission sources are produced in Monte Carlo source iteration: given some initial fission source at iteration c, a new flux distribution ϕ(c+1) is generated following transport. In the case of the inline algorithm, this transport is subject to cross sections generated from depletion using the flux ϕ(c), Σ(c). This transport will also immediately produce a fission source, corresponding to F(c+1)=νΣf(c)ϕ(c+1). This occurs before any additional depletion updates. Hence, when the iteration index is advanced, there will be a dependence on the previous νΣf.

The equations corresponding to EquationEq. (7) during the predictor and corrector step must also be defined. These will be written in terms of Bateman operators B, where the nuclide density vector at one time point and the next are related by

(21) Nn+1=BNn.(21)

The Bateman operators are a function of the time step Δt, the flux at time point n, ϕn, and potentially the projected flux and previous flux, ϕn+1 and ϕn1, respectively. Different Bateman operators may be used on the predictor or corrector step. Corresponding to the notation in CitationRef. 41, for the predictor, this may be the CE operator BCE(ϕn) or the higher-order Linear Extrapolation (LE) operator BLE(ϕn,ϕn1). Likewise, for the corrector, these operators may be the LI operator BLI(ϕn,ϕn+1) or the Quadratic Interpolation (QI) operator BQI(ϕn,ϕn+1,ϕn1). Each of these operators may also use substeps.Citation11,Citation12 Previous work initially used a very simple Bateman operator; for the explicit Euler scheme (which uses the CE operator), this was simply the Taylor expansion of the matrix exponential in EquationEq. (8) truncated to first order:

(22) B(ϕn)=I+ΔtΣϕn+Λ,(22)

where I is the identity matrix, Σ is the matrix containing cross sections and fission yields describing all neutron-induced reactions, and Λ is the matrix describing all decay chains by containing nuclide decay constants. Given these matrices, the burnup matrix can be written as

(23) An=A(ϕn)=Σϕn+Λ.(23)

The simple representation of the Bateman operator in EquationEq. (22) was previously used to facilitate the stability analysis, which calls for the parameter derivative of the Bateman operator with respect to all flux terms.Citation18,Citation41 The matrix exponential does not cleanly permit this analytically, but this derivative can be easily obtained by numerical means where it appears in any expressions derived.Citation41 Another effective approach may be to obtain this derivative expression not from the matrix exponential, but from the numerical method used to calculate it, such as CRAM. The CRAM of order k is given as

(24) Nn+1=α0+2Rei=1k/2αiAnΔtθiI1Nn,(24)

where the values of αi and θi are tabulated for a given k. Written as a Bateman operator, one has

(25) B(ϕn)=α0+2Rei=1k/2αiAnΔtθiI1.(25)

The first derivative of this operator with respect to the scalar flux ϕn given EquationEq. (23) is

(26) ϕnB=2ΔtRei=1k/2αiAnΔtθiI1ΣAnΔtθiI1,(26)

which can be used in the perturbation expressions necessary for stability analyses. In the current work, the Bateman operators will be assumed to be matrix exponentials, and their derivatives with respect to flux terms will be the analytic one given here.

The depleting system requires initial conditions in the form of a nuclide density distribution. This distribution is taken as uniform,

(27) N0(x)=N0.(27)

In practical calculations, even though Monte Carlo neutronics treats space as a continuous variable, it is necessary to discretize the nuclide density distribution across a geometry. The effect of this discretization has previously been accounted for: a continuous nuclide density field imposes a stability penalty as compared to a coarsely discretized field, although this penalty quickly saturates for any realistic discretization of a reactor burnup problem.Citation40

From specifying the initial nuclide density, and given the system of equations, initial and normalizations conditions, and lack of leakage, the initial neutronics solution is

(28a) ϕ0=PκΣf,0L(28a)

and

(28b) k0=νΣf,0Σa,0.(28b)

Given EquationEq. (28) and the initial condition on the nuclide density distribution, the next step in the Fourier analysis is to assert that all solutions at subsequent time points are given by small perturbations to the initial solutions:

(29a) ϕn=ϕ0+δϕn,(29a)

(29b) kn=k0+δkn,(29b)

and

(29c) Nn=N0+δNn.(29c)

Given the simplicity of the domain, these perturbations may be expressed as an exponential Fourier seriesCitation39:

(30a) δϕn=ω=δϕn,ωeiωx,(30a)

(30b) δkn=ω=δkn,ωeiωx,(30b)

and

(30c) δNn=ω=δNn,ωeiωx.(30c)

where ω is the Fourier frequency, allowable values of which are determined by the boundary conditions; the variables with a subscripted ω are the corresponding Fourier amplitudes at time point n and i is the imaginary unit. For reflective boundary conditions, ω=mπLmZ for periodic boundary conditions, this expression is simply multiplied by two to give the allowable frequencies. EquationEquation (30b) simplifies to having only a single, spatially flat Fourier mode, as the eigenvalue is uniform across the geometry. That is,

(31) δkn=δkn,0.(31)

The next step in the Fourier analysis inserts EquationEq. (29) into the equations describing the algorithm under consideration. This will commence with the BOS neutronics, where the neutron flux is assumed to have converged given a nuclide density distribution. Inserting EquationEq. (29) into EquationEq. (19), expanding, linearizing, and neglecting the higher-order terms givesCitation18

(32) D02x2δϕn=Σa,0ϕ0νσfTνΣf,0σaTΣa,0δNnΣa,0ϕ0δknk0.(32)

Expressing the perturbation terms as their Fourier series in EquationEq. (30), each Fourier amplitude can be considered independently due to their orthogonality, giving

(33) δϕn,ω=Σa,0ϕ0ω2D0νσfTνΣf,0σaTΣa,0δNn,ωδωΣa,0ϕ0ω2D0δkn,ωk0.(33)

where δω is the Kronecker delta, present as the eigenvalue only possesses a zeroth Fourier mode. Due to the presence of the unknown eigenvalue perturbation, this equation cannot be solved to find the zeroth mode flux perturbation. This is done using the power normalization condition [EquationEq. (18)] giving

(34) δϕn,0=ϕ0κσfTκΣf,0δNn,0.(34)

As the expression for the flux perturbation resulting from a converged neutronics solution subject to a nuclide density perturbation recurs frequently, it will henceforth be written as

(35) ΦωT=Σa,0ϕ0ω2D0νσfTνΣf,0σaTΣa,0ω0ϕ0κσfTκΣf,0ω=0.(35)

Next, the predictor Bateman operator will undergo the same treatment to propagate a perturbation in the BOS nuclide density vector to the EOS nuclide density vector. Here the CE operator will be used; the LE operator can be used but with a slight additional complication to the analysis.Citation41 For the CE operator, the perturbed version of EquationEq. (21) is

(36) N0+δNn+1(0)=BCE(ϕ0+δϕn)(N0+δNn).(36)

The flux perturbation in the Bateman operator is handled by a truncated Taylor expansion,

(37) BCE(ϕ0+δϕn)B0CE+ϕnBCEδϕn,(37)

where

(38) ϕnBCE=ϕnBCE|ϕn=ϕ0(38)

and

(39) B0CE=BCE(ϕ0).(39)

This gives (neglecting higher-order terms)

(40) δNn+1(0)=B0CEδNn+ϕnBCEN0δϕn+B0CEIN0.(40)

Considering the individual Fourier modes and using the expressions obtained for the flux perturbation gives

(41) δNn+1,ω(0)=B0CE+ϕnBCEN0ΦωTδNn,ω+δωB0CEIN0.(41)

For compactness in future expressions, this will be written as

(42) δNn+1,ω(0)=KωδNn,ω+δωB0CEIN0,(42)

where

(43) Kω=B0CE+ϕnBCEN0ΦωT.(43)

The analysis now turns to the inline portion of the algorithm where, starting from the value of Nn+1 produced by the predictor step and an initial fission source guess, an iteration between the corrector [given by EquationEq. (12) for a LI Bateman operator without substepping] and the source iteration algorithm [given by EquationEq. (20)] occurs.

The predicted nuclide density vector is given a (c) superscript, Nn+1(c), to highlight its dependence on the cycle/source iteration index. The analysis begins by inserting Eq. (29) [with all variables acquiring the (c) superscript] into EquationEq. (20) (updated for being at time point n+1), giving

(44) xD0+δDn+1(c)x+Σa,0+δΣa,n+1(c)ϕ0+δϕn+1(c+1)=1k0+δkn+1(c)νΣf,0+δνΣf,n+1(c1)ϕ0+δϕn+1(c).(44)

From the definition of k in EquationEq. (15), one can show that

(45) 1k0+δkn+1(c)1k0+1k0σaTΣa,0νσfTνΣf,0δNn+1,0(c1).(45)

With this and expanding, linearizing, and neglecting higher-order terms as normal, one obtains the following expression for each Fourier mode:

(46) δϕn+1,ω(c+1)=ρωϕ0δϕn+1,ω(c)ϕ0+νσfTνΣf,0δNn+1,ω(c1)σaTΣa,0δNn+1,ω(c)δωνσfTνΣf,0σaTΣa,0δNn+1,ω(c1)),(46)(46)

where ρω is

(47) ρω=Σa,0Σa,0+ω2D0.(47)

If ω=π/L for reflective boundary conditions (or ω=2π/L for periodic), then ρω is the initial dominance ratio of the system. In the case of ω=0, EquationEq. (46) reduces to

(48) δϕn+1,0(c+1)=ϕ0σaTΣa,0δNn+1,0(c1)δNn+1,0(c)+δϕn+1,0(c).(48)

This approach requires some accounting for the first source iteration given no νΣf,n+1(1) exists. Eigenvalue calculations using source iteration must be initialized with a guessed fission source νΣf,n+1(0)ϕn+1(0); in most eigenvalue neutronics calculations, this fission source guess is spatially flat. Following the predictor calculation, νΣf,n+1(0) is known, allowing ϕn+1(0) to be determined. The fission source normalization follows from the power normalization in EquationEq. (18). The uniform fission source per unit area corresponding to the power normalization is

(49) F=νΣf,0PκΣf,0=νΣf,0ϕ0L.(49)

If one asserts that

(50) FL=νΣf,n+1(0)ϕn+1(0)=νΣf,0+νσfTω=δNn+1,ω(0)eiωxϕ0+ω=δϕn+1,ω(0)eiωx,(50)

then, expanding, subtracting νΣf,0ϕ0, neglecting higher-order terms, and considering orthogonal Fourier modes gives

(51) δϕn+1,ω(0)=ϕ0νσfTνΣf,0δNn+1,ω(0).(51)

Furthermore, either a code or a user must make an initial guess for the eigenvalue, kn+1(0)=kg=k0+δkn+1(0). Inserting both the eigenvalue guess and the expression for the initial uniform fission source into EquationEq. (46), one obtains

(52) δϕn+1,ω(1)=ρωϕ0σaTΣa,0δNn+1,ω(0)δk(0)k0δω.(52)

Alternatively, one could use the same fission source as generated at the conclusion of the predictor step, F=νΣf,nϕn. This would simply give

(53a) δϕn+1,ω(0)=ΦωTδNn,ω(53a)

and

(53b) δNn+1,ω(1)=δNn,ω.(53b)

This will not be considered here, although the analysis would only differ slightly.

To estimate the stability of the inline scheme, it only remains to introduce the corrector update equation. Using an LI Bateman operator, the corrector step can be written as

(54) Nn+1(c+1)=BLI(ϕn,ϕn+1(c+1))Nn.(54)

Performing an identical expansion as for EquationEq. (36), one obtains

(55) δNn+1,ω(c+1)=B0LI+ϕnBLIN0ΦωTδNn,ω+ϕn+1BLIN0δϕn+1,ω(c+1)+δωB0LIIN0.(55)

Given EquationEqs. (46), (Equation52), and (Equation55), one can construct a final expression describing the behavior of the Fourier modes. For ω0, this is

(56) δNn+1,ω(c+1)=Gω(c+1)δNn,ω,(56)

where

(57a) Gω(c+1)=B0LI+ϕnBLIN0ΦωT+ϕn+1BLIN0Hω(c+1)(57a)

and

(57b) Hω(c+1)=ρωϕ01ϕ0Hω(c)+νσfTνΣf,0Gω(c1)σaTΣa,0Gω(c),(57b)

and these are initialized by

(58a) Gω(0)=Kω,(58a)

(58b) Gω(1)=0,(58b)

and

(58c) Hω(0)=[00].(58c)

As the behavior of the ω=0 frequency does not provide an indication of stability (as it is physical and permissible for this frequency to grow), the expression describing its behavior will not be derived or further analyzed.

III.A. Accumulating Reaction Rates over Several Cycles

The preceding analysis can be generalized to account for the effect of accumulating reaction rates over several cycles before performing a depletion update, similar to strategies that can be used for enforcing xenon equilibrium inline or inline thermal-hydraulic coupling.Citation1,Citation2 This accumulation was intended to eliminate deleterious effects from stochastic noise, but may also affect the stability behavior and so should be analyzed.

As the flux/fission source is no longer updated simultaneously with the isotopic field, it is convenient to define a staggered indexing to account for this. Two new iteration variables are defined such that

(59) c=j+lJ.(59)

The index j[0,J1] is the cycle index for a fixed material composition, where a depletion update is performed after J such cycles, this is to say that J is the number of cycles/batches over which the reaction rates are accumulated. After the depletion update is performed, j is reset to 0 and l is incremented. The value of l is 0 initially and indicates the number of nuclide density updates that have been performed. With this indexing, the equations describing neutronics and depletion during the corrector step become

(60a) xDn+1(l)x+Σa,n+1(l)ϕn+1(0,l)=1kn+1(j,l1)νΣf,n+1(l1)ϕn+1(j,l1),ifj=J1,(60a)

(60b) xDn+1(l)x+Σa,n+1(l)ϕn+1(j+1,l)=1kn+1(j,l)νΣf,n+1(l)ϕn+1(j,l),ifj<J1,(60b)

(60c) 1kn+1(j,l)=0LΣa,n+1(l)ϕn+1(j,l)dx0LνΣf,n+1(l)ϕn+1(j,l)dx,(60c)

and

(60d) Nn+1(l+1)=BLIϕn,1Jj=0J1ϕn+1(j,l)Nn.(60d)

The neutronics solution is still initialized by a flat fission source guess as before. These equations have the Fourier mode expressions:

(61a) δϕn+1,ω(0,l)=ρωϕ0δϕn+1,ω(j,l1)ϕ0+νσfTνΣf,0δNn+1,ω(l1)σaTΣa,0δNn+1,ω(l)δωνσfTνΣf,0σaTΣa,0δNn+1,ω(l1)),if j=J1,(61a)(61a)
(61b) δϕn+1,ω(j+1,l)=ρωϕ0δϕn+1,ω(j,l)ϕ0+νσfTνΣf,0σaTΣa,0δNn+1,ω(l)δωνσfTνΣf,0σaTΣa,0δNn+1,ω(l)),ifj<J1,(61b)

(61c) 1k0+δkn+1(j,l)1k0+1k0σaTΣa,0νσfTνΣf,0δNn+1,0(l),(61c)

and

(61d) δNn+1,ω(l+1)=B0LI+ϕnBLIN0ΦωTδNn,ω+ϕn+1BLIN01Jj=0J1δϕn+1,ω(j,l)+δωB0LIIN0.(61d)

For some initial flux perturbation at material update l, δϕn+1,ω(0,l), one can show that after j source iterations, the resulting flux perturbation is

(62) δϕn+1,ω(j,l)=ρωjδϕn+1,ω(0,l)+(1δω)ϕ0i=1jρωiνσfTνΣf,0σaTΣa,0δNn+1,ω(l).(62)

For ω0, given ρω<1, in the limit of infinite source iterations this gives

(63) limjδϕn+1,ω(j,l)=ρω1ρωνσfTνΣf,0σaTΣa,0δNn+1,ω(l).(63)

This expression is identical to the flux perturbation, which is obtained from solving neutronics to convergence, shown in EquationEq. (33) when the Kronecker delta disappears. Inserting EquationEq. (62) into EquationEq. (61d) gives

(64) δNn+1,ω(l+1)=B0LI+ϕnBLIN0ΦωTδNn,ω+δωB0LIIN0+ϕn+1BLIN01Jj=0J1ρωjδϕn+1,ω(0,l)+(1δω)(1δj)ϕ0i=1jρωiνσfTνΣf,0σaTΣa,0δNn+1,ω(l).(64)

This expression reduces to EquationEq. (55) when J=1. It only remains to express δϕn+1,ω(0,l) and δNn+1,ω(l) in terms of δNn,ω. The former for l=0 can be expressed using a modified version of EquationEq. (52) to account for the change in indexing:

(65) δϕn+1,ω(0,0)=ρωϕ0σaTΣa,0δNn+1,ω(0)δk(0)k0δω,(65)

where δNn+1,ω(0) and δk(0) are as defined previously.

Finally, the expression governing the Fourier modes (for ω0) is

(66) δNn+1,ω(l+1)=Gω(l+1)δNn,ω,(66)

where Gω(l+1) is given by

(67) Gω(l+1)=B0LI+ϕnBLIN0ΦωT+ϕn+1BLIN01Jj=0J1ρωjHω(l)+(1δJ1)ϕ0j=1J1i=1jρωiνσfTνΣf,0σaTΣa,0Gω(l),(67)

with

(68) Hω(l)=ρωϕ01ϕ0Hω(l1)+νσfTνΣf,0+(1δJ1)j=1J1ρωjνσfTνΣf,0σaTΣa,0Gω(l1)σaTΣa,0Gω(l).(68)

These are initialized by

(69a) Gω(0)=Kω,(69a)

(69b) Gω(1)=0,(69b)

and

(69c) Hω(0)=[00].(69c)

If J=1, EquationEqs. (66), Equation(67), and Equation(68) reduce to their single batch counterparts.

III.B. Relaxing the Corrector Iteration

Finally, one last generalization of the algorithm is considered, namely, applying a relaxation during the inline corrector iteration. This would replace the corrector update given by EquationEq. (60d) with

(70) Nn+1(l+1)=αBLIϕn,1Jj=0J1ϕn+1(j,l)Nn+(1α)Nn+1(l),(70)

where α is a relaxation factor. The other equations in Eq. (60) remain unchanged. Following the same logic as earlier, one ultimately obtains another equation of the form of EquationEq. (66), the only difference being that Gω(l+1) is now given by

Gω(l+1)=αB0LI+ϕnBLIN0ΦωT+(1α)Gω(l)+αϕn+1BLIN01Jj=0J1ρωjHω(l)+(1δJ1)ϕ0j=1J1i=1jρωiνσfTνΣf,0σaTΣa,0Gω(l).(71)

Meanwhile, the expression for Hω(l) and the initialization conditions remain the same.

IV. NUMERICAL INVESTIGATIONS

To investigate the stability of the inline scheme and the accuracy of the expressions derived, two simple nuclide systems were considered. These are identical to those described in CitationRefs. 18 and Citation41. The first is a two-nuclide system with a fissile isotope and a nondecaying fission product. The second is a semirealistic system with eight nuclides corresponding to 235U, 238U, 239Np, 239Pu, a generic nondecaying fission product, 135I, 135Xe, and 1H. The cross sections for these isotopes were generated by Serpent and their decay constants and fission yields are realistic. Where simulations are shown, they are the burnup of a slab diffusion problem with reflective boundary conditions. The code implementing the diffusion/depletion solver and for obtaining all stability estimates is in MATLAB. It must be highlighted that the MATLAB model used does not include stochastic noise, which would be present in any Monte Carlo implementation of the inline burnup algorithm. This is done to demonstrate the correctness of the stability analysis, which also does not account for stochastic effects. In future work, where the algorithm will be implemented in a Monte Carlo burnup code, the effect of noise must be numerically investigated to convincingly demonstrate the effectiveness of the algorithm.

IV.A. Two-Nuclide System

The first system of nuclide has cross sections given (in barns) as

σa=3000150,σf=30000,σs=10000,

with σt=σa+σs. The fission yield of the product nuclide is one, giving (also in barns)

Σ=300003000150.

As stated before, these nuclides do not undergo decay and so Λ=0. The average neutron production per fission ν is 2.3 and the energy release per fission κ is 200 MeV. The problem has a length L of 3 m and a power per unit area P of 10 kW/cm2.

The results for the two-nuclide system are shown for various settings for the inline depletion scheme. Although this system is not fully realistic, its simplicity serves to highlight some features of the inline scheme before examining its performance on a more realistic problem. For this problem, the number of cycles during the inline portion of the algorithm has a negligible effect on stability so it will not be explored; 100 cycles will be used in all stability estimates and numerical examples to follow.

First, the stability eigenvalues of the first few spurious modes are shown for the inline depletion scheme in , calculated by extracting eigenvalues from EquationEq. (57). Recall that the allowable frequencies are given by ω=mπLmZ, such that the lowest spurious frequency that can appear is given by m=1. It should be highlighted that there are two eigenvalues per mode (given the stability matrix is 2 × 2), but only the most unstable is shown; the other is persistently about 1 for all frequencies, but slowly decreases with increasing time-step length. Notably, the scheme is not unconditionally stable; as is the case for most other burnup schemes, the first mode is the most unstable. That said, the inline scheme is substantially more stable than both the explicit Euler and predictor-corrector schemes applied to the same problem, as seen in (with their stability matrices coming from CitationRefs. 18 and Citation19, respectively).

Fig. 1. Stability properties of the inline scheme applied to the two-nuclide system.

Fig. 1. Stability properties of the inline scheme applied to the two-nuclide system.

A numerical demonstration of the inline scheme’s stability is shown in , with an oscillating instability growing when the stable time-step length of about 11 days is exceeded.

Fig. 2. Flux solutions produced by the inline scheme at subsequent time points.

Fig. 2. Flux solutions produced by the inline scheme at subsequent time points.

IV.B. Eight-Nuclide System

The eight-nuclide system has cross sections (in barns):

σa=31.40.835086.55001280000,σf=31.40.005086.50000,σs=000000013.5.

The ordering of these nuclides is the same as listed at the beginning of Sec. V. The three fissile isotopes are presumed to have identical fission yields for product nuclides, with 90% of fissions producing the generic product, 5% producing iodine, and 1% producing xenon. Hence (in barns),

Σ=31.4000000000.83500000008.300000000086.5000028.260.0045077.85500001.570.0002504.32500000.3140.000504.32500128000000000000.

Only three of the nuclides undergo decay—neptunium, iodine, and xenon—with their standard half-lives used here. Hence (in h–1),

Λ=ln20000000000000000001/57.600000001/57.60000000000000000001/700000001/71/9.1000000000.

The initial nuclide density for the system corresponds to a fresh UO2 pin. This is (in atom/b/cm)

N0=0.0003940.007389000000.31296.

Other quantities are mostly identical to the two-nuclide case: ν is 2.3 across all fissile nuclides, κ is 200 MeV, and the problem is 3 m long. The power per unit area differs, in order to ensure a pressurized water reactor (PWR) power density of 104 W/cm3; hence, the power per unit area P is 104L W/cm2.

As there are eight nuclides, there will be eight eigenvalues governing the growth of each spurious mode. Hence, for simplicity of presentation, only the spectral radius (the maximum absolute eigenvalue) of a given mode will be shown. Furthermore, as discussed in CitationRef. 41, due to the initial xenon transient to which PWR-like problems are subjected, stability estimates evolve rapidly and substantially from the estimates given by fresh fuel without xenon and iodine. Therefore, the value of N0 used to estimate stability uses the saturated values of xenon and iodine corresponding to xenon equilibrium; these are 3.5641×109 and 5.8998×109 atoms/b/cm, respectively.

In this eight-nuclide system, as opposed to the two-nuclide system, the number of cycles/source iterations has a significant effect on stability. This is shown (before batching and relaxation) in . These results are relatively optimistic. For a semirealistic system of nuclides, with sufficient iterations, the inline scheme can be unconditionally stable. It is worth emphasizing that realistic problems of similar dimensions often require ~100 cycles to converge the fission source, at least when source convergence acceleration is not applied.

Fig. 3. Spectral radius of the inline scheme for various source iteration settings applied to the eight-nuclide system.

Fig. 3. Spectral radius of the inline scheme for various source iteration settings applied to the eight-nuclide system.

The stability of the system can be further improved by accumulating fluxes/reaction rates over several cycles and relaxing the nuclide density update, albeit there is an apparently optimal choice of batch size. This is shown in . Both the relaxed 5-batch and 10-batch stability estimates are co-linear and stable below one. The eigenvalues were extracted from EquationEq. (71). From this, one can see that choosing the batch size and relaxation factor well can result in an apparently unconditionally stable burnup scheme. One possible explanation for the apparently optimal batch size is the tradeoff between increasing the number of accumulation batches (which might approximately correspond to a relaxation) and performing additional corrector updates, which can better converge the nuclide densities; as the batch size increases, fewer corrector updates are performed for a fixed number of inactive/corrector cycles.

Fig. 4. Spectral radius of the inline scheme for various batch sizes and relaxation factors over 70 cycles applied to the eight-nuclide system.

Fig. 4. Spectral radius of the inline scheme for various batch sizes and relaxation factors over 70 cycles applied to the eight-nuclide system.

Unfortunately, however, these stability results are overly optimistic. In spite of estimating stability with saturated xenon and iodine, the stability properties of each of the inline schemes evolves substantially with time. This behavior was originally highlighted in CitationRef. 41 for more traditional burnup schemes, although the changes in stability properties shown there were quite mild following some initial burnup. The evolution of the 70-cycle scheme with a batch size of five and a relaxation factor of 0.6 is shown in . This was generated by burning a one-region problem forward and using the resulting nuclide density to evaluate stability at different levels of burnup. In this figure, burnup is given in units of fissions per initial fissile atom (FIFA). Fortunately, performing the same exercise when applying 100 cycles with a batch size of five and a relaxation factor of 0.6 is apparently unconditionally stable (at least up to the same level of burnup). The relative stability of these two different settings is shown in . Perhaps reassuringly, the instability that emerges with 70 cycles takes a (nearly unphysically) long time to develop and even then its relative magnitude is quite small.

Fig. 5. Spectral radius of the inline scheme for different degrees of burnup when using 70 cycles, 5 cycles per batch, and a relaxation factor of 0.6.

Fig. 5. Spectral radius of the inline scheme for different degrees of burnup when using 70 cycles, 5 cycles per batch, and a relaxation factor of 0.6.

Fig. 6. Flux solutions produced by the inline scheme at subsequent time points for different numbers of cycles when applying a batch size of five and a relaxation factor of 0.6.

Fig. 6. Flux solutions produced by the inline scheme at subsequent time points for different numbers of cycles when applying a batch size of five and a relaxation factor of 0.6.

V. CONCLUSIONS

A new algorithm for burnup calculations was proposed in the style of methods applied to coupling neutronics with thermal hydraulics and xenon equilibrium. The stability properties of several variants of the algorithm were analyzed, and theory compared favorably with numerical results. Therefore, the algorithm appears to be an efficient means of stabilizing depletion, as well perhaps, as an efficient means of converging depletion alongside several other elements of reactor physics.

Work is currently under way to implement this proposed depletion algorithm in both SerpentCitation28 and SCONE (CitationRef. 43). This will allow for a fairer appraisal of the algorithm’s computational benefits in a high-fidelity Monte Carlo code, as well as how it performs when subjected to stochastic phenomena. Importantly, this will also allow the accuracy and efficiency of the algorithm to be examined when varying the statistics applied to obtaining a transport solution, i.e., subject to variations in the particle population per cycle. Furthermore, this will allow investigation of more complex chains of nuclides (such as the full chain resulting from ENDF or the CASL chain, as considered in CitationRef. 30) and their effects on stability. In the present paper, xenon and generic, absorbing fission products were examined under the presumption that they are responsible for much of the stability challenge. However, it may well be the case that prominent isotopes like samarium and gadolinium (which were not considered) may also pose unique challenges to the algorithm and stability.

Future work will consider how the presently proposed depletion algorithm interacts with the inline treatment of other physics, such as thermal hydraulics and xenon equilibrium; similar investigations were previously performed by GillCitation4 with promising results. In large, finely divided reactor problems, it is possible that repeated depletion calculations per time point may become comparable to or exceed the cost of the neutron transport calculations.Citation29 This might be partially circumvented either by using a smaller depletion system (at least during the inline portion of the algorithm),Citation29,Citation30 or potentially through the use of a faster, reduced-order Bateman solver.Citation44 Finally, it should also be straightforward to extend the work to include higher-order burnup algorithms, e.g., with substepping, linear extrapolation, and quadratic interpolation.Citation12

Disclosure Statement

No potential conflict of interest was reported by the authors.

Data Availability Statement

To the best of the authors’ knowledge, this paper and references herein contain all the data needed to reproduce and validate the results presented.

References

  • D. P. GRIESHEIMER, “In-line Xenon Convergence Algorithm for Monte Carlo Reactor Calculations,” Proceedings of International Conference on Physics of Reactors (PHYSOR 2010), Pittsburgh, Pennsylvania (2010).
  • D. F. GILL, D. P. GRIESHEIMER, and D. L. AUMILLER, “Numerical Methods in Coupled Monte Carlo and Thermal-Hydraulic Calculations,” Nucl. Sci. Eng., 185, 1, 194 (2017); https://doi.org/10.13182/NSE16-3.
  • M. KREHER, B. FORGET, and K. SMITH, “Single-Batch Monte Carlo Multiphysics Coupling,” Proceedings of International Conference on Mathematics and Computational Methods Applied to Nuclear Science and Engineering, Portland, Oregon. (2019).
  • D. GILL, “Inline Thermal and Xenon Feedback Iterations in Monte Carlo Reactor Calculations,” Proceedings of PHYSOR2020 – International Conference on Physics of Reactors: Transition to a Scalable Nuclear Future, Cambridge, UK (2020).
  • B. KOCHUNAS, A. FITZGERALD, and E. LARSEN, “Fourier Analysis of Iteration Schemes for k-eigenvalue Transport Problems with Flux-Dependent Cross Sections,” J. Comput. Phys., 345, 294 (2017); https://doi.org/10.1016/j.jcp.2017.05.028.
  • J. P. SENECAL and W. JI, “Approaches for Mitigating Over-Solving in Multiphysics Simulations,” Int. J. Numer. Methods Eng., 112, 6, 503 (2017); https://doi.org/10.1002/nme.5516.
  • Q. SHEN, N. ADAMOWICZ, and B. KOCHUNAS, “Relationship Between Relaxation and Partial Convergence of Nonlinear Diffusion Acceleration for Problems with Feedback,” Proceedings of International Conference on Mathematics and Computational Methods Applied to Nuclear Science and Engineering, Portland, Oregon. (2019).
  • Q. SHEN et al., “X-CMFD: A Robust Iteration Scheme for CMFD-based Acceleration of Neutron Transport Problems with Nonlinear Feedback,” Proceedings of International Conference on Mathematics and Computational Methods Applied to Nuclear Science and Engineering, Portland, Oregon. (2019).
  • Q. SHEN and B. KOCHUNAS, “A Robust Relaxation-Free Multiphysics Iteration Scheme for CMFD-Accelerated Neutron Transport k-Eigenvalue Calculations—I: Theory,” Nucl. Sci. Eng., 195, 1176 (2021); https://doi.org/10.1080/00295639.2021.1906585.
  • Q. SHEN, S. CHOI, and B. KOCHUNAS, “A Robust Relaxation-Free Multiphysics Iteration Scheme for CMFD—Accelerated Neutron Transport k-Eigenvalue calculation—II: Numerical Results,” Nucl. Sci. Eng., 195, 1202 (2021); https://doi.org/10.1080/00295639.2021.1906586.
  • A. ISOTALO and P. AARNIO, “Higher Order Methods for Burnup Calculations with Bateman Solutions,” Ann. Nucl. Energy, 38, 1987 (2011); https://doi.org/10.1016/j.anucene.2011.04.022.
  • A. ISOTALO and P. AARNIO, “Substep Methods for Burnup Calculations with Bateman Solutions,” Ann. Nucl. Energy, 38, 11, 2509 (2011); https://doi.org/10.1016/j.anucene.2011.07.012.
  • A. ISOTALO and V. SAHLBERG, “Comparison of Neutronics-Depletion Coupling Schemes for Burnup Calculations,” Nucl. Sci. Eng., 179, 4, 434 (2015); https://doi.org/10.13182/NSE14-35.
  • C. JOSEY, B. FORGET, and K. SMITH, “High Order Methods for the Integration of the Bateman Equations and Other Problems of the Form of Y´ = F(y, T) Y,” J. Comput. Phys., 350, 296 (2017); https://doi.org/10.1016/j.jcp.2017.08.025.
  • J. DUFEK and J. E. HOOGENBOOM, “Numerical Stability of Existing Monte Carlo Burnup Codes in Cycle Calculations of Critical Reactors,” Nucl. Sci. Eng., 162, 3, 307 (2009); https://doi.org/10.13182/NSE08-69TN.
  • J. DUFEK et al., “Numerical Stability of the Predictor-Corrector Method in Monte Carlo Burnup Calculations of Critical Reactors,” Ann. Nucl. Energy, 56, 34 (2013); https://doi.org/10.1016/j.anucene.2013.01.018.
  • D. KOTLYAR and E. SHWAGERAUS, “On the Use of Predictor-Corrector Method for Coupled Monte Carlo Burnup Codes,” Ann. Nucl. Energy, 58, 228 (2013); https://doi.org/10.1016/j.anucene.2013.03.034.
  • J. D. DENSMORE, D. F. GILL, and D. P. GRIESHEIMER, “Stability Analysis of Burnup Calculations,” Trans. Am. Nucl. Soc., 109, 695 (2013); https://doi.org/10.1007/s10967-012-2210-3.2.
  • P. COSGROVE, E. SHWAGERAUS, and G. T. PARKS, “Stability Analysis of Predictor-Corrector Schemes for Coupling Neutronics and Depletion,” Ann. Nucl. Energy, 149, 107781 ( Dec. 2020); https://doi.org/10.1016/j.anucene.2020.107781.
  • A. E. ISOTALO, J. LEPPÄNEN, and J. DUFEK, “Preventing Xenon Oscillations in Monte Carlo Burnup Calculations by Enforcing Equilibrium Xenon Distribution,” Ann. Nucl. Energy, 60, 78 (2013); https://doi.org/10.1016/j.anucene.2013.04.031.
  • J. DUFEK, D. KOTLYAR, and E. SHWAGERAUS, “The Stochastic Implicit Euler Method: A Stable Coupling Scheme for Monte Carlo Burnup Calculations,” Ann. Nucl. Energy, 60, 295 (2013); https://doi.org/10.1016/j.anucene.2013.05.015.
  • D. KOTLYAR and E. SHWAGERAUS, “Numerically Stable Monte Carlo-Burnup-Thermal Hydraulic Coupling Schemes,” Ann. Nucl. Energy, 63, 371 (2014); https://doi.org/10.1016/j.anucene.2013.08.016.
  • D. KOTLYAR and E. SHWAGERAUS, “Stochastic Semi-Implicit Substep Method for Coupled Depletion Monte-Carlo Codes,” Ann. Nucl. Energy, 92, 52 (2016); https://doi.org/10.1016/j.anucene.2016.01.022.
  • C. JOSEY, “Development and Analysis of High Order Neutron Transport-Depletion Coupling Algorithms,” Massachusetts Institute of Technology, Department of Nuclear Science and Engineering, PhD Thesis (2017).
  • V. VALTAVIRTA and J. LEPPÄNEN, “New Stochastic Substep Based Burnup Scheme for Serpent,” PHYSOR 2018: Reactor Physics paving the way towards more efficient systems, Cancun, Mexico, 2, (2018).
  • P. COSGROVE, E. SHWAGERAUS, and G. T. PARKS, “A Simple Implicit Coupling Scheme for Monte Carlo Neutronics and Isotopic Depletion,” Ann. Nucl. Energy, 141, 107374 (June 2020); https://doi.org/10.1016/j.anucene.2020.107374.
  • P. COSGROVE, “Numerical Stability of Monte Carlo Neutron Transport and Isotopic Depletion for Nuclear Reactor Analysis,” University of Cambridge, PhD Thesis (2020).
  • J. LEPPÄNEN et al., “The Serpent Monte Carlo Code: Status, Development and Applications in 2013,” Ann. Nucl. Energy, 82, 142 (2015); https://doi.org/10.1016/j.anucene.2014.08.024.
  • D. P. GRIESHEIMER, D. C. CARPENTER, and M. H. STEDRY, “Practical Techniques for Large-Scale Monte Carlo Reactor Depletion Calculatons,” Prog. Nucl. Energy, 101, 409 (2017); https://doi.org/10.1016/j.pnucene.2017.05.018.
  • P. K. ROMANO et al., “Depletion Capabilities in the OpenMC Monte Carlo Particle Transport Code,” Ann. Nucl. Energy, 152, 107989 (Mar. 2021); https://doi.org/10.1016/j.anucene.2020.107989.
  • G. BELL and S. GLASSTONE, “Nuclear Reactor Theory,” US Atomic Energy Commission, Washington, DC (1970).
  • C. MOLER and C. V. LOAN, “Nineteen Dubious Ways to Compute the Exponential of a Matrix, Twenty-Five Years Later,” SIAM Rev., 45, 1, 3 (2003); https://doi.org/10.1137/S00361445024180.
  • M. PUSA and J. LEPPÄNEN, “Computing the Matrix Exponential in Burnup Calculations,” Nucl. Sci. Eng., 164, 2, 140 (2010); https://doi.org/10.13182/NSE09-14.
  • M. PUSA, “Rational Approximations to the Matrix Exponential in Burnup Calculations,” Nucl. Sci. Eng., 169, 2, 155 (2011); https://doi.org/10.13182/NSE10-81.
  • J. C. SUBLET et al., “FISPACT-II: An Advanced Simulation System for Activation, Transmutation and Material Modelling,” Nucl. Data Sheets, 139, 77 (2017); https://doi.org/10.1016/j.nds.2017.01.002.
  • M. ELLIS et al., “Spatially Continuous Depletion Algorithm for Monte Carlo Simulations,” Trans. Am. Nucl. Soc., 115, 1.
  • D. GRIESHEIMER et al., “MC21 V.6.0—A Continuous-Energy Monte Carlo Particle Transport Code with Integrated Reactor Feedback Capabilities,” Ann. Nucl. Energy, 82, 29 (2015); https://doi.org/10.1016/j.anucene.2014.08.020.
  • H. ROBBINS and S. MONRO, “A Stochastic Approximation Method,” Ann. Math. Statist., 22, 3, 400 (1951); https://doi.org/10.1214/aoms/1177729586.
  • N. ADAMOWICZ and P. COSGROVE, “Recent Advances in the Stability Analysis of Burn-Up Calculations,” Proc. M&C 2021, Raleigh, North Carolina (2021).
  • P. COSGROVE and N. ADAMOWICZ, “Stability Analysis of Spatial Discretisation in Burn-Up Calculations,” Proc. M&C 2021, Raleigh, North Carolina (2021).
  • P. COSGROVE and E. SHWAGERAUS, “Stability Analysis of Higher-Order Neutronics-Depletion Coupling Schemes and Bateman Operators,” J. Comput. Phys., 448, 110702 ( Jan. 2022); https://doi.org/10.1016/j.jcp.2021.110702.
  • J. LIEBEROTH, “A Monte Carlo Technique to Solve the Static Eigenvalue Problem of the Boltzmann Transport Equation,” Nukleonik, 11, 213 (1968).
  • M. A. KOWALSKI et al., “Scone: A Student-Oriented Modifiable Monte Carlo Particle Transport Framework,” J. Nucl. Eng., 2, 1, 57 (2021); https://doi.org/10.3390/jne2010006.
  • C. CASTAGNA et al., “Development of a Reduced Order Model for Fuel Burnup Analysis,” Energies, 13, 4, 890 (2020); https://doi.org/10.3390/en13040890.