This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

An arbitrarily high order and asymptotic preserving kinetic scheme in compressible fluid dynamic

Rémi Abgrall and Fatemeh Nassajian Mojarrad
Institute of Mathematics, University of Zürich
Winterthurerstrasse 190, CH 8057 Zürich
Switzerland
Abstract

We present a class of arbitrarily high order fully explicit kinetic numerical methods in compressible fluid dynamics, both in time and space, which include the relaxation schemes by S. Jin and Z. Xin. These methods can use CFL number larger or equal to unity on regular Cartesian meshes for multi-dimensional case. These kinetic models depend on a small parameter that can be seen as a ”Knudsen” number. The method is asymptotic preserving in this Knudsen number. Also, the computational costs of the method are of the same order of a fully explicit scheme. This work is the extension of Abgrall et al. (2022) [1] to multi-dimensional systems. We have assessed our method on several problems for two dimensional scalar problems and Euler equations and the scheme has proven to be robust and to achieve the theoretically predicted high order of accuracy on smooth solutions.

keywords. Kinetic scheme; Compressible fluid dynamics; High order methods; Explicit schemes; Asymptotic preserving; Defect correction method

1 Introduction

In this paper, we consider a system of hyperbolic conservation laws in multiple spatial dimensions

𝐮t+i=1d𝐀i(𝐮)xi=0,\frac{\partial\mathbf{u}}{\partial t}+\sum\limits_{i=1}^{d}\dfrac{\partial\mathbf{A}_{i}(\mathbf{u})}{\partial x_{i}}=0, (1a)
with the initial condition
𝐮(𝐱,0)=𝐮0(𝐱).\mathbf{u}(\mathbf{x},0)=\mathbf{u}_{0}(\mathbf{x}). (1b)

where 𝐮:d×+K\mathbf{u}:\mathbb{R}^{d}\times\mathbb{R}_{+}\rightarrow\mathbb{R}^{K} and the flux functions 𝐀d\mathbf{A}_{d} are locally Lipschitz continuous on K\mathbb{R}^{K} with values in K\mathbb{R}^{K}. We approximate the solution 𝐮\mathbf{u} by considering a special class of discrete kinetic systems [2, 3].

Consider a solution 𝐟:d×+L\mathbf{f}:\mathbb{R}^{d}\times\mathbb{R}_{+}\rightarrow\mathbb{R}^{L} to the Cauchy problem for the following sequence of semilinear systems

𝐟t+i=1dΛi𝐟xi=𝕄(𝐮ε)𝐟ε,\frac{\partial\mathbf{f}}{\partial t}+\sum\limits_{i=1}^{d}\Lambda_{i}\frac{\partial\mathbf{f}}{\partial x_{i}}=\frac{\mathbb{M}(\mathbf{u}^{\varepsilon})-\mathbf{f}}{\varepsilon}, (2a)
with the initial condition
𝐟(𝐱,0)=𝐟0(𝐱).\mathbf{f}(\mathbf{x},0)=\mathbf{f}_{0}(\mathbf{x}). (2b)

Here Λi\Lambda_{i} are real diagonal L×LL\times L matrices, ε\varepsilon is a positive number, 𝕄:KL\mathbb{M}:\mathbb{R}^{K}\rightarrow\mathbb{R}^{L} is a Lipschitz continuous function, and the function 𝐮ε\mathbf{u}^{\varepsilon} is defined by

𝐮ε=i=1N𝐟i=𝐟\mathbf{u}^{\varepsilon}=\sum\limits_{i=1}^{N}\mathbf{f}_{i}=\mathbb{P}\mathbf{f}

where \mathbb{P} is a real constant coefficients K×LK\times L matrix. To connect problem (2a) - (2b) with problem (1a) - (1b), we assume that 𝕄\mathbb{M} is a Maxwellian function for (1a), i.e.

{𝕄(𝐮)=𝐮,Λi𝕄(𝐮)=𝐀i(𝐮),i=1,,d.\begin{cases}\mathbb{P}\mathbb{M}(\mathbf{u})=\mathbf{u},\\ \mathbb{P}\Lambda_{i}\mathbb{M}(\mathbf{u})=\mathbf{A}_{i}(\mathbf{u}),\quad i=1,\cdots,d.\end{cases} (3)

Clearly, if 𝐟\mathbf{f} converges in some strong topology to a limit 𝐠\mathbf{g} and if 𝐟0\mathbb{P}\mathbf{f}_{0} converges to 𝐮0\mathbf{u}_{0}, then 𝐠\mathbb{P}\mathbf{g} is a solution of problem (1a) - (1b). Actually, the system (2a) is only a BGK approximation for (1a), see e.g [4, 5] and the references therein. A general stability theory is developed in [6], and will implicitly be used throughout this paper, in particular to guaranty that the continuous problem (2), equipped with (3) is well posed.

The method of [2, 7], where the first numerical schemes based on (2) are described, are based on splitting techniques. As a result, the order in time is restricted to 2 and can only be improved by nontrivial manipulations [8]. There exists already some ways for higher than second order. For example, one approach is to use a relaxed upwind schemes which are proposed running up to CFL 1 and up to third order in time and space for the finite volume scheme [9]. A class of high-order weighted essentially nonoscillatory (WENO) reconstructions based on relaxation approximation of hyperbolic systems of conservation laws is presented in [10]. In [11] a splitting approach is adopted with a regular CFL stability condition for the overall finite volume scheme. In [12] a discontinuous Galerkin method for solving general hyperbolic systems of conservation laws is constructed which is CFL independant and can be of arbitrary order in time and space.

Designing algorithms that are uniformly stable and acccurate in ε\varepsilon when ε0\varepsilon\rightarrow 0 has been an active area of research in recent years. Such schemes are called asymptotic preserving in the sense of Jin [13]. Several asymptotic-preserving methods based on IMEX techniques have been recently proposed. In [14, 15] IMEX Runge–Kutta methods is presented for general hyperbolic systems with relaxation. Specific methods for the Boltzmann equation in the hyperbolic and diffusive regimes with a computational cost that is independent of ε\varepsilon is proposed in [16, 17].

Many of the these methods have inherent limitations with respect to the order that can be achieved with the time discretization, for example, due to the time splitting. In [1] an arbitrarily high order class of kinetic numerical methods that can run at least at CFL 1 in one dimension is constructed. This work is an extension of [1] to the multi-dimensional case.

We are interested in a computationally explicit scheme that solves (2) with uniform accuracy of order r>0r>0 for all ε>0\varepsilon>0 and with a CFL condition, based on the matrices Λi,i=1,,d\Lambda_{i},~{}i=1,\cdots,d , that is larger than one for a given system (1). We consider the two dimensional case. The idea is to start from (2a), we describe first the discretisation of Λx𝐟x\Lambda_{x}\frac{\partial\mathbf{f}}{\partial x} and Λy𝐟y\Lambda_{y}\frac{\partial\mathbf{f}}{\partial y}. The second step is to discretise in time. We take into account the source term. The resulting scheme is fully implicit. The next step is to show that, thanks to the operator \mathbb{P}, and using a particular time discretisation, we can make it computationally explicit, and high order accurate. It is independent of ε\varepsilon.

The paper is organised as follows. In Section 2, we describe the time-stepping algorithm. In Section 3, we describe the space discretisation. In Section 4, We study the stability of the discretisation of the homogeneous problem. In Section 5, we illustrate the robustness and accuracy of the proposed method by means of numerical tests. Finally, Section 6 provide some conclusions and future perspectives.

2 Time discretisation

Here, we consider the two dimensional case, i.e.

𝐟t+Λx𝐟x+Λy𝐟y=𝕄(𝐟)𝐟ε,\frac{\partial\mathbf{f}}{\partial t}+\Lambda_{x}\frac{\partial\mathbf{f}}{\partial x}+\Lambda_{y}\frac{\partial\mathbf{f}}{\partial y}=\frac{\mathbb{M}(\mathbb{P}\mathbf{f})-\mathbf{f}}{\varepsilon}, (4)

Knowing the solution 𝐟n\mathbf{f}^{n} at time tnt_{n}, we are looking for the solution at time tn+1t_{n+1}. First we discretise (2a) in space, we get

𝐟t+1ΔxΛxδx𝐟+1ΔyΛyδy𝐟=𝕄(𝐟)𝐟ε,\frac{\partial\mathbf{f}}{\partial t}+\frac{1}{\Delta x}\Lambda_{x}\delta^{x}\mathbf{f}+\frac{1}{\Delta y}\Lambda_{y}\delta^{y}\mathbf{f}=\frac{\mathbb{M}(\mathbb{P}\mathbf{f})-\mathbf{f}}{\varepsilon}, (5)

and notice that

𝐟t+1Δx(Λxδx𝐟)+1Δy(Λyδy𝐟)=0.\frac{\partial\mathbb{P}\mathbf{f}}{\partial t}+\frac{1}{\Delta x}\mathbb{P}(\Lambda_{x}\delta^{x}\mathbf{f})+\frac{1}{\Delta y}\mathbb{P}(\Lambda_{y}\delta^{y}\mathbf{f})=0. (6)

Since we want to have a running CFL number of at least one, we use an IMEX defect correction method. Using this, we follow [1] where a defect correction technique can be used and it is made explicit because the nonlinear term 𝕄(𝐟)\mathbb{M}(\mathbb{P}\mathbf{f}) is explicit.

2.1 An explicit high order timestepping approach

The next step is to discretise in time, we subdivide the interval [tn,tn+1][t_{n},t_{n+1}] into sub-intervals obtained from the partition

tn=t(0)<t(1)<t(p)<<t(M)=tn+1t_{n}=t_{(0)}<t_{(1)}<\cdots t_{(p)}<\cdots<t_{(M)}=t_{n+1}

with t(p)=tn+βpΔtt_{(p)}=t_{n}+\beta_{p}\Delta t. We approximate the integral over time using quadrature formula

tntn+βpΔtϕ(s)𝑑sΔtq=0Mwpqϕ(tn+βqΔt)\int_{t_{n}}^{t_{n}+\beta_{p}\Delta t}\phi(s)\;ds\approx\Delta t\sum_{q=0}^{M}w_{pq}\phi(t_{n}+\beta_{q}\Delta t)

where wpqw_{pq} are the weights. In order to obtain consistent quadrature formula of order q+1q+1, we should have

wpq=0βplq(s)𝑑s,q=0Mwpq=βpw_{pq}=\int_{0}^{\beta_{p}}l_{q}(s)\;ds,\quad\sum_{q=0}^{M}w_{pq}=\beta_{p}

where {lq}q=0M\{l_{q}\}_{q=0}^{M} is the Lagrangian basis polynomial of the q-th node.

Let 𝐱l=(xi,yj)\mathbf{x}_{l}=(x_{i},y_{j}) be a fixed grid point, 𝐟ln,p𝐟(𝐱l,tn+βpΔt)\mathbf{f}_{l}^{n,p}\approx\mathbf{f}(\mathbf{x}_{l},t_{n}+\beta_{p}\Delta t) and 𝐟ln,0=𝐟ln\mathbf{f}_{l}^{n,0}=\mathbf{f}_{l}^{n}. We introduce the corrections r=0,,Rr=0,\cdots,R for each subinterval [tp,tp+1][t_{p},t_{p+1}] and denote the solution at the rr-th correction and the time tpt_{p} by 𝐟n,p,r\mathbf{f}^{n,p,r}. The notation 𝐅\mathbf{F} is the collection of all the approximations for the sub-steps i.e. the vector 𝐅=(𝐟n,1,,𝐟n,M)T\mathbf{F}=(\mathbf{f}^{n,1},\cdots,\mathbf{f}^{n,M})^{T}. The notation 𝐅(r)\mathbf{F}^{(r)} represents the vector 𝐅(r)=(𝐟n,1,r,,𝐟n,M,r)T\mathbf{F}^{(r)}=(\mathbf{f}^{n,1,r},\cdots,\mathbf{f}^{n,M,r})^{T} i.e. the vector of all the approximations for the sub-steps at the rr-th correction. Now we use a defect correction method and proceed within the time interval [tn,tn+1][t_{n},t_{n+1}] as follows:

  1. 1.

    For r=0r=0, set 𝐅(0)=(𝐟n,,𝐟n)T\mathbf{F}^{(0)}=(\mathbf{f}^{n},\cdots,\mathbf{f}^{n})^{T}.

  2. 2.

    For each correction r0r\geq 0, define 𝐅(r+1)\mathbf{F}^{(r+1)} by

    L1(𝐅(r+1))=L1(𝐅(r))L2(𝐅(r))L_{1}(\mathbf{F}^{(r+1)})=L_{1}(\mathbf{F}^{(r)})-L_{2}(\mathbf{F}^{(r)}) (7)
  3. 3.

    Set 𝐅n+1=𝐅(M)\mathbf{F}^{n+1}=\mathbf{F}^{(M)}.

Formulation (7) relies on a Lemma which has been proven in [18].

In the following, we introduce the differential operators L1L_{1} and L2L_{2}. We first define the high order differential operator L2L_{2}. By integrating (5) on [0,t][0,t], we have

𝐟(𝐱,t)𝐟(𝐱,0)+1Δx0tΛxδx𝐟(𝐱,s)𝑑s+1Δy0tΛyδy𝐟(𝐱,s)𝑑s=1ε0t(𝕄(𝐟(𝐱,s))𝐟(𝐱,s))𝑑s\mathbf{f}(\mathbf{x},t)-\mathbf{f}(\mathbf{x},0)+\frac{1}{\Delta x}\int\limits_{0}^{t}\Lambda_{x}\delta^{x}\mathbf{f}(\mathbf{x},s)\;ds+\frac{1}{\Delta y}\int\limits_{0}^{t}\Lambda_{y}\delta^{y}\mathbf{f}(\mathbf{x},s)\;ds=\frac{1}{\varepsilon}\int\limits_{0}^{t}\Big{(}\mathbb{M}\big{(}\mathbb{P}\mathbf{f}(\mathbf{x},s)\big{)}-\mathbf{f}(\mathbf{x},s)\Big{)}\;ds (8)

Hence, we get the following approximation for (4)

𝐟ln,q𝐟ln,0+ΔtΔx(k=0MwqkΛxδlx𝐟n,k)+ΔtΔy(k=0MwqkΛyδly𝐟n,k)Δtεk=0Mwqk(𝕄(𝐟ln,k)𝐟ln,k)=0\mathbf{f}_{l}^{n,q}-\mathbf{f}_{l}^{n,0}+\frac{\Delta t}{\Delta x}\big{(}\sum_{k=0}^{M}w_{qk}\Lambda_{x}\delta_{l}^{x}\mathbf{f}^{n,k}\big{)}+\frac{\Delta t}{\Delta y}\big{(}\sum_{k=0}^{M}w_{qk}\Lambda_{y}\delta_{l}^{y}\mathbf{f}^{n,k}\big{)}-\frac{\Delta t}{\varepsilon}\sum_{k=0}^{M}w_{qk}\big{(}\mathbb{M}(\mathbb{P}\mathbf{f}_{l}^{n,k})-\mathbf{f}_{l}^{n,k}\big{)}=0 (9)

for q=1,,Mq=1,\cdots,M. For any ll, define [L2(𝐅(r))]l[L_{2}(\mathbf{F}^{(r)})]_{l} as

[L2(𝐅(r))]l\displaystyle[L_{2}(\mathbf{F}^{(r)})]_{l} =𝐅l(r)𝐅l(0)+ΔtΔxΛxWδlx𝐅(r)+ΔtΔxΛx𝐰0δlx𝐟n,0+ΔtΔyΛyWδly𝐅(r)+ΔtΔyΛy𝐰0δly𝐟n,0\displaystyle=\mathbf{F}_{l}^{(r)}-\mathbf{F}_{l}^{(0)}+\frac{\Delta t}{\Delta x}\Lambda_{x}W\delta_{l}^{x}\mathbf{F}^{(r)}+\frac{\Delta t}{\Delta x}\Lambda_{x}\mathbf{w}_{0}\otimes\delta_{l}^{x}\mathbf{f}^{n,0}+\frac{\Delta t}{\Delta y}\Lambda_{y}W\delta_{l}^{y}\mathbf{F}^{(r)}+\frac{\Delta t}{\Delta y}\Lambda_{y}\mathbf{w}_{0}\otimes\delta_{l}^{y}\mathbf{f}^{n,0}
ΔtεW(𝕄(𝐅l(r))𝐅l(r))Δtε𝐰0(𝕄(𝐟ln,0)𝐟ln,0)\displaystyle-\frac{\Delta t}{\varepsilon}W\big{(}\mathbb{M}(\mathbb{P}\mathbf{F}_{l}^{(r)})-\mathbf{F}_{l}^{(r)}\big{)}-\frac{\Delta t}{\varepsilon}\mathbf{w}_{0}\otimes\big{(}\mathbb{M}(\mathbb{P}\mathbf{f}_{l}^{n,0})-\mathbf{f}_{l}^{n,0}\big{)}

where

W=(w11w1MwM1wMM),𝐰0=(w10wM0)W=\begin{pmatrix}w_{11}&\cdots&w_{1M}\\ \vdots&\vdots&\vdots\\ w_{M1}&\cdots&w_{MM}\\ \end{pmatrix},\quad\mathbf{w}_{0}=\begin{pmatrix}w_{10}\\ \vdots\\ w_{M0}\end{pmatrix}

and

𝕄(𝐅l(r))=(𝕄(𝐟ln,1,r),,𝕄(𝐟ln,M,r))T\mathbb{M}(\mathbb{P}\mathbf{F}_{l}^{(r)})=\big{(}\mathbb{M}(\mathbb{P}\mathbf{f}_{l}^{n,1,r}),\cdots,\mathbb{M}(\mathbb{P}\mathbf{f}_{l}^{n,M,r})\big{)}^{T}

The resulting scheme derived by L2L_{2} operator is implicit, and it is very difficult to solve. Now we describe the low order differential operator L1L_{1}. We use the forward Euler method on each sub-time step

𝐟ln,q𝐟ln,0+βqΔtΔxΛxδlx𝐟n,0+βqΔtΔyΛyδly𝐟n,0Δtεk=0Mwqk(𝕄(𝐟ln,k)𝐟ln,k)=0\mathbf{f}_{l}^{n,q}-\mathbf{f}_{l}^{n,0}+\beta_{q}\frac{\Delta t}{\Delta x}\Lambda_{x}\delta_{l}^{x}\mathbf{f}^{n,0}+\beta_{q}\frac{\Delta t}{\Delta y}\Lambda_{y}\delta_{l}^{y}\mathbf{f}^{n,0}-\frac{\Delta t}{\varepsilon}\sum_{k=0}^{M}w_{qk}\big{(}\mathbb{M}(\mathbb{P}\mathbf{f}_{l}^{n,k})-\mathbf{f}_{l}^{n,k}\big{)}=0 (10)

for q=1,,Mq=1,\cdots,M. The low order differential operator [L1(𝐅(r))]l[L_{1}(\mathbf{F}^{(r)})]_{l} reads

[L1(𝐅(r))]l\displaystyle[L_{1}(\mathbf{F}^{(r)})]_{l} =𝐅l(r)𝐅l(0)+ΔtΔxBΛxδlx𝐅(0)+ΔtΔyBΛyδly𝐅(0)\displaystyle=\mathbf{F}_{l}^{(r)}-\mathbf{F}_{l}^{(0)}+\frac{\Delta t}{\Delta x}B\Lambda_{x}\delta_{l}^{x}\mathbf{F}^{(0)}+\frac{\Delta t}{\Delta y}B\Lambda_{y}\delta_{l}^{y}\mathbf{F}^{(0)}
ΔtεW(𝕄(𝐅l(r))𝐅l(r))Δtε𝐰0(𝕄(𝐟ln,0)𝐟ln,0)\displaystyle-\frac{\Delta t}{\varepsilon}W\big{(}\mathbb{M}(\mathbb{P}\mathbf{F}^{(r)}_{l})-\mathbf{F}_{l}^{(r)}\big{)}-\frac{\Delta t}{\varepsilon}\mathbf{w}_{0}\otimes\big{(}\mathbb{M}(\mathbb{P}\mathbf{f}_{l}^{n,0})-\mathbf{f}_{l}^{n,0}\big{)}

where B=diag(β1,,βM)B=\text{diag}(\beta_{1},\cdots,\beta_{M}). This is still an implicit approximation in time, but the convection part is now explicit. We also note we have kept the same form of the source term approximation in both cases, for reasons that will be explained bellow.

Now we can write (7) as a multi-step method where each step writes as

𝐅l(r+1)𝐅l(0)+ΔtΔxBΛxδlx𝐅(0)+ΔtΔyBΛyδly𝐅(0)ΔtεW(𝕄(𝐅l(r+1))𝐅l(r+1))Δtε𝐰0(𝕄(𝐟ln,0)𝐟ln,0)=ΔtΔxΛxW(δlx𝐅(0)δlx𝐅(r))+ΔtΔyΛyW(δly𝐅(0)δly𝐅(r))\begin{split}\mathbf{F}_{l}^{(r+1)}-\mathbf{F}_{l}^{(0)}&+\frac{\Delta t}{\Delta x}B\Lambda_{x}\delta_{l}^{x}\mathbf{F}^{(0)}+\frac{\Delta t}{\Delta y}B\Lambda_{y}\delta_{l}^{y}\mathbf{F}^{(0)}-\frac{\Delta t}{\varepsilon}W\big{(}\mathbb{M}(\mathbb{P}\mathbf{F}^{(r+1)}_{l})-\mathbf{F}_{l}^{(r+1)}\big{)}\\ &-\frac{\Delta t}{\varepsilon}\mathbf{w}_{0}\otimes\big{(}\mathbb{M}(\mathbb{P}\mathbf{f}_{l}^{n,0})-\mathbf{f}_{l}^{n,0}\big{)}=\frac{\Delta t}{\Delta x}\Lambda_{x}W(\delta_{l}^{x}\mathbf{F}^{(0)}-\delta_{l}^{x}\mathbf{F}^{(r)})+\frac{\Delta t}{\Delta y}\Lambda_{y}W(\delta_{l}^{y}\mathbf{F}^{(0)}-\delta_{l}^{y}\mathbf{F}^{(r)})\end{split} (11)

for any ll. By applying \mathbb{P} to this equation, we will obtain the following equation for calculating 𝐅l(r+1)\mathbb{P}\mathbf{F}_{l}^{(r+1)}

𝐅l(r+1)=𝐅l(0)ΔtΔxΛxWδlx𝐅(r)ΔtΔx𝐰0Λxδlx𝐟n,0ΔtΔyΛyWδly𝐅(r)ΔtΔy𝐰0Λyδly𝐟n,0\mathbb{P}\mathbf{F}_{l}^{(r+1)}=\mathbb{P}\mathbf{F}_{l}^{(0)}-\frac{\Delta t}{\Delta x}\mathbb{P}\Lambda_{x}W\delta_{l}^{x}\mathbf{F}^{(r)}-\frac{\Delta t}{\Delta x}\mathbf{w}_{0}\otimes\mathbb{P}\Lambda_{x}\delta_{l}^{x}\mathbf{f}^{n,0}-\frac{\Delta t}{\Delta y}\mathbb{P}\Lambda_{y}W\delta_{l}^{y}\mathbf{F}^{(r)}-\frac{\Delta t}{\Delta y}\mathbf{w}_{0}\otimes\mathbb{P}\Lambda_{y}\delta_{l}^{y}\mathbf{f}^{n,0} (12)

and we substitute 𝐅l(r+1)\mathbb{P}\mathbf{F}_{l}^{(r+1)} into the Maxwellian in the (11). Alternatively, one can rewrite (11) as follows

(IdM×M+ΔtεW)𝐅l(r+1)=ΔtεW𝕄(𝐅l(r+1))+𝐅l(0)ΔtΔxΛxWδlx𝐅(r)ΔtΔx𝐰0Λxδlx𝐟n,0ΔtΔyΛyWδly𝐅(r)ΔtΔy𝐰0Λyδly𝐟n,0+Δtε𝐰0(𝕄(𝐟ln,0)𝐟ln,0)\begin{split}(\text{Id}_{M\times M}+\frac{\Delta t}{\varepsilon}W)\mathbf{F}_{l}^{(r+1)}&=\frac{\Delta t}{\varepsilon}W\mathbb{M}(\mathbb{P}\mathbf{F}_{l}^{(r+1)})+\mathbf{F}_{l}^{(0)}-\frac{\Delta t}{\Delta x}\Lambda_{x}W\delta_{l}^{x}\mathbf{F}^{(r)}-\frac{\Delta t}{\Delta x}\mathbf{w}_{0}\otimes\Lambda_{x}\delta_{l}^{x}\mathbf{f}^{n,0}\\ &-\frac{\Delta t}{\Delta y}\Lambda_{y}W\delta_{l}^{y}\mathbf{F}^{(r)}-\frac{\Delta t}{\Delta y}\mathbf{w}_{0}\otimes\Lambda_{y}\delta_{l}^{y}\mathbf{f}^{n,0}+\frac{\Delta t}{\varepsilon}\mathbf{w}_{0}\otimes\big{(}\mathbb{M}(\mathbb{P}\mathbf{f}_{l}^{n,0})-\mathbf{f}_{l}^{n,0}\big{)}\end{split} (13)

If IdM×M+ΔtW/ε\text{Id}_{M\times M}+{\Delta t}W/{\varepsilon} is invertible, the defect correction computes the solution at time tn+1t_{n+1} using MM steps of the form

𝐅l(r+1)=(IdM×M+ΔtεW)1(ΔtεW𝕄(𝐅l(r+1))+𝐅l(0)ΔtΔxΛxWδlx𝐅(r)ΔtΔx𝐰0Λxδlx𝐟n,0ΔtΔyΛyWδly𝐅(r)ΔtΔy𝐰0Λyδly𝐟n,0+Δtε𝐰0(𝕄(𝐟ln,0)𝐟ln,0)).\begin{split}\mathbf{F}_{l}^{(r+1)}&=(\text{Id}_{M\times M}+\frac{\Delta t}{\varepsilon}W)^{-1}\bigg{(}\frac{\Delta t}{\varepsilon}W\mathbb{M}(\mathbb{P}\mathbf{F}_{l}^{(r+1)})+\mathbf{F}_{l}^{(0)}-\frac{\Delta t}{\Delta x}\Lambda_{x}W\delta_{l}^{x}\mathbf{F}^{(r)}-\frac{\Delta t}{\Delta x}\mathbf{w}_{0}\otimes\Lambda_{x}\delta_{l}^{x}\mathbf{f}^{n,0}\\ &-\frac{\Delta t}{\Delta y}\Lambda_{y}W\delta_{l}^{y}\mathbf{F}^{(r)}-\frac{\Delta t}{\Delta y}\mathbf{w}_{0}\otimes\Lambda_{y}\delta_{l}^{y}\mathbf{f}^{n,0}+\frac{\Delta t}{\varepsilon}\mathbf{w}_{0}\otimes\big{(}\mathbb{M}(\mathbb{P}\mathbf{f}_{l}^{n,0})-\mathbf{f}_{l}^{n,0}\big{)}\bigg{)}.\end{split} (14a)
For computing 𝐅l(r+1)\mathbb{P}\mathbf{F}_{l}^{(r+1)}, for any ll, we rewrite (12) as (for simplicity, we drop the superscript nn)
𝐟lq,r+1𝐟l0+ΔtΔxk=0MwqkΛxδlx𝐟k,r+ΔtΔyk=0MwqkΛyδly𝐟k,r=0,\mathbb{P}\mathbf{f}_{l}^{q,r+1}-\mathbb{P}\mathbf{f}_{l}^{0}+\frac{\Delta t}{\Delta x}\sum_{k=0}^{M}w_{qk}\mathbb{P}\Lambda_{x}\delta_{l}^{x}\mathbf{f}^{k,r}+\frac{\Delta t}{\Delta y}\sum_{k=0}^{M}w_{qk}\mathbb{P}\Lambda_{y}\delta_{l}^{y}\mathbf{f}^{k,r}=0, (14b)

for q=1,,Mq=1,\cdots,M.

We write the increment δl𝐟k\delta_{l}\mathbf{f}^{k} as a sum of residuals, as follows

δl𝐟k=ΔyΛx(𝐟^i+12,jk𝐟^i12,jk)+ΔxΛy(𝐟^i,j+12k𝐟^i,j12k)=Φl,(i,j)[i,i+1]×[j,j+1],k+Φl,(i,j)[i1,i]×[j,j+1],k+Φl,(i,j)[i,i+1]×[j1,j],k+Φl,(i,j)[i1,i]×[j1,j],k\begin{split}\delta_{l}\mathbf{f}^{k}&=\Delta y\Lambda_{x}(\hat{\mathbf{f}}_{i+\frac{1}{2},j}^{k}-\hat{\mathbf{f}}_{i-\frac{1}{2},j}^{k})+\Delta x\Lambda_{y}(\hat{\mathbf{f}}_{i,j+\frac{1}{2}}^{k}-\hat{\mathbf{f}}_{i,j-\frac{1}{2}}^{k})\\ &=\Phi_{l,(i,j)}^{[i,i+1]\times[j,j+1],k}+\Phi_{l,(i,j)}^{[i-1,i]\times[j,j+1],k}+\Phi_{l,(i,j)}^{[i,i+1]\times[j-1,j],k}+\Phi_{l,(i,j)}^{[i-1,i]\times[j-1,j],k}\end{split} (15a)
with
Φl,(i,j)[i,i+1]×[j,j+1],k=12(Λx(𝐟^i+1/2,j𝐟ij)Δy+Λy(𝐟^i,j+1/2𝐟ij))Δy),Φl,(i,j)[i1,i]×[j,j+1],k=12(Λx(𝐟^i+1/2,j𝐟ij)Δy+Λy(𝐟ij𝐟^i,j1/2)Δx),Φl,(i,j)[i1,i]×[j1,j],k=12(Λx(𝐟ij𝐟^i1/2,j)Δy+Λy(𝐟ij𝐟^ij1/2)Δx),Φl,(i,j)[i1,i]×[j1,j],k=12(Λx(𝐟ij𝐟^i1/2,j)Δy+Λy(𝐟^i,j+1/2𝐟ij)Δx).\begin{split}\Phi_{l,(i,j)}^{[i,i+1]\times[j,j+1],k}&=\frac{1}{2}\bigg{(}\Lambda_{x}\big{(}\hat{\mathbf{f}}_{i+1/2,j}-\mathbf{f}_{ij})\Delta y+\Lambda_{y}\big{(}\hat{\mathbf{f}}_{i,j+1/2}-\mathbf{f}_{ij})\big{)}\Delta y\bigg{)},\\ \Phi_{l,(i,j)}^{[i-1,i]\times[j,j+1],k}&=\frac{1}{2}\bigg{(}\Lambda_{x}\big{(}\hat{\mathbf{f}}_{i+1/2,j}-\mathbf{f}_{ij}\big{)}\Delta y+\Lambda_{y}(\mathbf{f}_{ij}-\hat{\mathbf{f}}_{i,j-1/2}\big{)}\Delta x\bigg{)},\\ \Phi_{l,(i,j)}^{[i-1,i]\times[j-1,j],k}&=\frac{1}{2}\bigg{(}\Lambda_{x}\big{(}\mathbf{f}_{ij}-\hat{\mathbf{f}}_{i-1/2,j}\big{)}\Delta y+\Lambda_{y}\big{(}\mathbf{f}_{ij}-\hat{\mathbf{f}}_{ij-1/2}\big{)}\Delta x\bigg{)},\\ \Phi_{l,(i,j)}^{[i-1,i]\times[j-1,j],k}&=\frac{1}{2}\bigg{(}\Lambda_{x}\big{(}\mathbf{f}_{ij}-\hat{\mathbf{f}}_{i-1/2,j}\big{)}\Delta y+\Lambda_{y}\big{(}\hat{\mathbf{f}}_{i,j+1/2}-\mathbf{f}_{ij}\big{)}\Delta x\bigg{)}.\\ \end{split} (15b)

We see that

Φl,(i,j)[i,i+1]×[j,j+1],k+Φl,(i+1,j)[i,i+1]×[j,j+1],k+Φl,(i+1,j+1)[i,i+1]×[j,j+1],k+Φl,(i,j+1)[i,i+1]×[j,j+1],k=Δx2Λx(𝐟i+1,j+𝐟i+1,j+1)Δx2Λx(𝐟ij+𝐟i,j+1)+Δy2Λy(𝐟i+1,j+𝐟i+1,j+1)Δy2Λy(𝐟ij+𝐟i,j+1),\begin{split}\Phi_{l,(i,j)}^{[i,i+1]\times[j,j+1],k}&+\Phi_{l,(i+1,j)}^{[i,i+1]\times[j,j+1],k}+\Phi_{l,(i+1,j+1)}^{[i,i+1]\times[j,j+1],k}+\Phi_{l,(i,j+1)}^{[i,i+1]\times[j,j+1],k}\\ &=\frac{\Delta x}{2}\Lambda_{x}\big{(}\mathbf{f}_{i+1,j}+\mathbf{f}_{i+1,j+1}\big{)}-\frac{\Delta x}{2}\Lambda_{x}\big{(}\mathbf{f}_{ij}+\mathbf{f}_{i,j+1}\big{)}\\ &+\frac{\Delta y}{2}\Lambda_{y}\big{(}\mathbf{f}_{i+1,j}+\mathbf{f}_{i+1,j+1}\big{)}-\frac{\Delta y}{2}\Lambda_{y}\big{(}\mathbf{f}_{ij}+\mathbf{f}_{i,j+1}\big{)},\end{split} (16)
Refer to caption
Figure 1: Sample quad element.

i.e. the sum of the residuals assigned to each of the four vertices of the quad [xi,xi+1]×[yj,yj+1][x_{i},x_{i+1}]\times[y_{j},y_{j+1}] is equal to an approximation of the flux across the boundary of this quad, see Fig. 1. This guaranties local conservation, see [19]. The reason of writing the space update as as sum of residual will become clear in section 3.2.

Now we explain the residuals for an example where the velocities are orthogonal. This case is suggested in [7]. In order to construct the system (2a) one must find \mathbb{P}, 𝕄\mathbb{M}, and Λj(j=1,,d)\Lambda_{j}~{}(j=1,\cdots,d) such that the consistency relations (3) are satisfied. We take L=N×KL=N\times K, =(IdK×K,,IdK×K)\mathbb{P}=(\text{Id}_{K\times K},\cdots,\text{Id}_{K\times K}) with NN blocks IdK×K\text{Id}_{K\times K}, the identity matrix in K\mathbb{R}^{K}. Each matrix Λj(j=1,,d)\Lambda_{j}~{}(j=1,\cdots,d) is constituted of NN diagonal blocks of size K×KK\times K, i.e.

Λj=diag(C1(j),,CN(j)),Ci(j)=λi(j)IdK×K,λi(j).\Lambda_{j}=\text{diag}(C_{1}^{(j)},\cdots,C_{N}^{(j)}),\quad C_{i}^{(j)}=\lambda_{i}^{(j)}\text{Id}_{K\times K},\quad\lambda_{i}^{(j)}\in\mathbb{R}.

where j=1,,Nj=1,\cdots,N. Then (2a) can be rewritten as

𝐟jt+i=1dλj(i)𝐟jxi=𝕄j(𝐮ε)𝐟jε,j=1,,N.\frac{\partial\mathbf{f}_{j}}{\partial t}+\sum\limits_{i=1}^{d}\lambda_{j}^{(i)}\frac{\partial\mathbf{f}_{j}}{\partial x_{i}}=\frac{\mathbb{M}_{j}(\mathbf{u}^{\varepsilon})-\mathbf{f}_{j}}{\varepsilon},\,j=1,\cdots,N. (17)

in which 𝐮ε=j=1N𝐟n\mathbf{u}^{\varepsilon}=\sum\limits_{j=1}^{N}\mathbf{f}_{n}. This example works with any number of blocks provided that Nd+1N\geq d+1. We take the velocities such that

Λj=diag(λijIdK×K)1iN,\Lambda_{j}=\text{diag}(\lambda_{ij}\text{Id}_{K\times K})_{1\leq i\leq N},
i=1Nλij=0,j=1,,d,\sum_{i=1}^{N}\lambda_{ij}=0,~{}j=1,\cdots,d,

and

i=1Nλilλij=0,j,l=1,d,jl.\sum_{i=1}^{N}\lambda_{il}\lambda_{ij}=0,~{}j,l=1,\cdots d,~{}j\neq l.

Therefore, the Maxwellian function is given by

𝕄i(𝐮)=𝐮N+j=1d𝐀j(𝐮)aj2λij,i=1,,N.\mathbb{M}_{i}(\mathbf{u})=\frac{\mathbf{u}}{N}+\sum_{j=1}^{d}\frac{\mathbf{A}_{j}(\mathbf{u})}{a_{j}^{2}}\lambda_{ij},~{}i=1,\cdots,N.

where aj2=i=1Nλij2a_{j}^{2}=\sum_{i=1}^{N}\lambda_{ij}^{2}. We fix J,N1J,N^{\prime}\geq 1, λ>0\lambda>0 and set

{λj=nJλ1jJ,J1,Λx=diag(λndiag(cos(iπ2N) I)1i4N)N1,Λy=diag(λndiag(sin(iπ2N) I)1i4N).\begin{cases}\lambda_{j}=\frac{n}{J}\lambda&1\leq j\leq J,~{}J\geq 1,\\ \Lambda_{x}=\text{diag}\Big{(}\lambda_{n}\text{diag}\big{(}\cos(\frac{i\pi}{2N^{{}^{\prime}}})\text{ I}\big{)}_{1\leq i\leq 4N^{{}^{\prime}}}\Big{)}&N^{{}^{\prime}}\geq 1,\\ \Lambda_{y}=\text{diag}\Big{(}\lambda_{n}\text{diag}\big{(}\sin(\frac{i\pi}{2N^{{}^{\prime}}})\text{ I}\big{)}_{1\leq i\leq 4N^{{}^{\prime}}}\Big{)}.\end{cases}

Here we have N=4NJN=4N^{\prime}J. We define the Maxwellian functions by

𝕄j(𝐮)=1N(𝐮+12iJλ(J+1)(2J+1)(𝐀1cosjπ2N+𝐀2sinjπ2N)),1j4N,1iJ.\mathbb{M}_{j}(\mathbf{u})=\frac{1}{N}\Big{(}\mathbf{u}+\frac{12\;i\;J}{\lambda(J+1)(2J+1)}(\mathbf{A}_{1}\cos\frac{j\pi}{2N^{{}^{\prime}}}+\mathbf{A}_{2}\sin\frac{j\pi}{2N^{{}^{\prime}}})\Big{)},~{}1\leq j\leq 4N^{{}^{\prime}},1\leq i\leq J.

Now, we consider the special case where J=N=1J=N^{\prime}=1 and N=4N=4. Set

{Λx=diag(λ diag(cos(iπ2) I)1i4),Λy=diag(λ diag(sin(iπ2) I)1i4).\begin{cases}\Lambda_{x}=\text{diag}\Big{(}\lambda\text{ diag}\big{(}\cos(\frac{i\pi}{2})\text{ I}\big{)}_{1\leq i\leq 4}\Big{)},\\ \Lambda_{y}=\text{diag}\Big{(}\lambda\text{ diag}\big{(}\sin(\frac{i\pi}{2})\text{ I}\big{)}_{1\leq i\leq 4}\Big{)}.\end{cases}

Therefore the Maxwellian functions are

𝕄i(𝐮)=14(𝐮+2λ(𝐀1cosiπ2+𝐀2siniπ2)).\mathbb{M}_{i}(\mathbf{u})=\frac{1}{4}\Big{(}\mathbf{u}+\frac{2}{\lambda}(\mathbf{A}_{1}\cos\frac{i\pi}{2}+\mathbf{A}_{2}\sin\frac{i\pi}{2})\Big{)}.

for 1i41\leq i\leq 4.

2.2 Error estimate

The natural question is: how many iterations should we do in the method (14) to recover the time and space accuracy?

Again, we will consider the two dimensional version of (1a). Let us consider φ:2K\varphi:\mathbb{R}^{2}\rightarrow\mathbb{R}^{K}, which is continuously differentiable on 2\mathbb{R}^{2} with values in K\mathbb{R}^{K} and has a compact support. We will consider the discrete version of L2L^{2} and H1H^{1} norms of φ\varphi as follows

φL22=i,jΔxΔyφi,j2,φH12=φL22+i,jΔxΔy(φi+1,jφijΔx2+φi,j+1φi,jΔy2).\|\varphi\|_{L^{2}}^{2}=\sum\limits_{i,j\in\mathbb{Z}}\Delta x\Delta y\|\varphi_{i,j}\|^{2},\quad\|\varphi\|_{H^{1}}^{2}=\|\varphi\|_{L^{2}}^{2}+\sum\limits_{i,j\in\mathbb{Z}}\Delta x\Delta y\Big{(}\Big{\|}\dfrac{\varphi_{i+1,j}-\varphi_{ij}}{\Delta x}\Big{\|}^{2}+\Big{\|}\dfrac{\varphi_{i,j+1}-\varphi_{i,j}}{\Delta y}\Big{\|}^{2}\Big{)}.

In the following, we will set:

Di+1/2,jφ=φi+1,jφijΔx,Di,j+1/2φ=φi,j+1φi,jΔy and Dφij=(Di+1/2,jφ,Di,j+1/2φ).D_{i+1/2,j}\varphi=\dfrac{\varphi_{i+1,j}-\varphi_{ij}}{\Delta x},D_{i,j+1/2}\varphi=\dfrac{\varphi_{i,j+1}-\varphi_{i,j}}{\Delta y}\text{ and }D\varphi_{ij}=(D_{i+1/2,j}\varphi,D_{i,j+1/2}\varphi).

We have a Poincaré inegality, on any compact.

One can consider the discrete equivalent of Lloc2L^{2}_{loc} and Hloc1H^{-1}_{loc} in an interval I=[a,b]×[c,d]I=[a,b]\times[c,d] as follows

𝐅2,I=supφC01(I)Ki,jΔxΔyφi,j,𝐟i,jφL2,𝐅1,I=supφC01(I)Ki,jΔxΔyφi,j,𝐟i,jφH1.\|\mathbf{F}\|_{2,I}=\sup\limits_{\varphi\in C_{0}^{1}(I)^{K}}\frac{\sum\limits_{i,j\in\mathbb{Z}}\Delta x\Delta y\langle\varphi_{i,j},\mathbf{f}_{i,j}\rangle}{\|\varphi\|_{L^{2}}},\quad\|\mathbf{F}\|_{-1,I}=\sup\limits_{\varphi\in C_{0}^{1}(I)^{K}}\frac{\sum\limits_{i,j\in\mathbb{Z}}\Delta x\Delta y\langle\varphi_{i,j},\mathbf{f}_{i,j}\rangle}{\|\varphi\|_{H^{1}}}.

We have a first result on the Hloc1H^{-1}_{loc} estimate of L2(𝐅)L1(𝐅)L_{2}(\mathbf{F})-L_{1}(\mathbf{F}).

Lemma 2.1.

If 𝐅^i+12,j=l=rsαl𝐅i+l,j\widehat{\mathbf{F}}_{i+\frac{1}{2},j}=\sum\limits_{l=-r}^{s}\alpha_{l}\mathbf{F}_{i+l,j} and 𝐅^i,j+12=l=rsαl𝐅i,j+l\widehat{\mathbf{F}}_{i,j+\frac{1}{2}}=\sum\limits_{l=-r^{\prime}}^{s^{\prime}}\alpha^{\prime}_{l}\mathbf{F}_{i,j+l}, we have

L2(𝐅)L1(𝐅)1,IγΔt𝐅2,I\|L_{2}(\mathbf{F})-L_{1}(\mathbf{F})\|_{-1,I}\leq\gamma\Delta t\|\mathbf{F}\|_{2,I}

where γ=max{maxrls|αl|,maxrls|αl|}×max{maxm|λmx|,maxn|λny|}\gamma=\max\{\max\limits_{-r\leq l\leq s}|\alpha_{l}|,\max\limits_{-r^{\prime}\leq l\leq s^{\prime}}|\alpha^{\prime}_{l}|\}\times\max\{\max\limits_{m}|\lambda_{m}^{x}|,\max\limits_{n}|\lambda_{n}^{y}|\}.

Proof.
|i,jΔxΔy\displaystyle\Big{|}\sum\limits_{i,j}\Delta x\Delta y\langle φi,j,[L2(𝐅)]i,j[L1(𝐅)]i,j|2\displaystyle\varphi_{i,j},[L_{2}(\mathbf{F})]_{i,j}-[L_{1}(\mathbf{F})]_{i,j}\rangle\Big{|}^{2}
=|i,jΔtΔyφi,j,ΛxW(𝐅^i+12,j𝐅^i12,j)+i,jΔtΔxφi,j,ΛyW(𝐅^i,j+12𝐅^i,j12)|2\displaystyle=\Big{|}\sum\limits_{i,j}\Delta t\Delta y\langle\varphi_{i,j},\Lambda_{x}W(\widehat{\mathbf{F}}_{i+\frac{1}{2},j}-\widehat{\mathbf{F}}_{i-\frac{1}{2},j})\rangle+\sum\limits_{i,j}\Delta t\Delta x\langle\varphi_{i,j},\Lambda_{y}W(\widehat{\mathbf{F}}_{i,j+\frac{1}{2}}-\widehat{\mathbf{F}}_{i,j-\frac{1}{2}})\rangle\Big{|}^{2}
=|i,jΔtΔxΔyDi+12,jφ,ΛxW𝐅^i+12,j+i,jΔtΔxΔyDi,j+12φ,ΛyW𝐅^i,j+12|2\displaystyle=\Big{|}\sum\limits_{i,j}\Delta t\Delta x\Delta y\langle D_{i+\frac{1}{2},j}\varphi,\Lambda_{x}W\widehat{\mathbf{F}}_{i+\frac{1}{2},j}\rangle+\sum\limits_{i,j}\Delta t\Delta x\Delta y\langle D_{i,j+\frac{1}{2}}\varphi,\Lambda_{y}W\widehat{\mathbf{F}}_{i,j+\frac{1}{2}}\rangle\Big{|}^{2}
Δt2Λx2W2(i,jΔxΔyDi+12,j2+i,jΔxΔyDi,j+12φ2)\displaystyle\leq\Delta t^{2}\|\Lambda_{x}\|^{2}\|W\|^{2}\Big{(}\sum\limits_{i,j}\Delta x\Delta y\|D_{i+\frac{1}{2},j}\|^{2}+\sum\limits_{i,j}\Delta x\Delta y\|D_{i,j+\frac{1}{2}}\varphi\|^{2}\Big{)}
×(i,jΔxΔy𝐅^i+12,j2+i,jΔxΔy𝐅^i,j+122)\displaystyle\times\Big{(}\sum\limits_{i,j}\Delta x\Delta y\|\widehat{\mathbf{F}}_{i+\frac{1}{2},j}\|^{2}+\sum\limits_{i,j}\Delta x\Delta y\|\widehat{\mathbf{F}}_{i,j+\frac{1}{2}}\|^{2}\Big{)}
γ2Δt2φH12𝐅2,I2.\displaystyle\leq\gamma^{2}\Delta t^{2}\|\varphi\|_{H^{1}}^{2}\|\mathbf{F}\|_{2,I}^{2}.

Then the proof is complete. ∎

Before we proceed to proposition 2.3, we need a further result on the behaviour of L1L_{1}.

Lemma 2.2.

Let us consider 𝐅,𝐅\mathbf{F},\mathbf{F}^{\prime} and 𝐆,𝐆\mathbf{G},\mathbf{G}^{\prime} such that

[L1(𝐅)]l=𝐆l,[L1(𝐅)]l=𝐆l[L_{1}(\mathbf{F})]_{l}=\mathbf{G}_{l},\quad[L_{1}(\mathbf{F}^{\prime})]_{l}=\mathbf{G}^{\prime}_{l}

for some ll. And we assume that there exists γ,γ>0\gamma,\gamma^{\prime}>0 such that

(IdM×M+ΔtεW)1γ,Δtε(IdM×M+ΔtεW)1Wγ,\|(\text{Id}_{M\times M}+\frac{\Delta t}{\varepsilon}W)^{-1}\|\leq\gamma,\quad\frac{\Delta t}{\varepsilon}\|(\text{Id}_{M\times M}+\frac{\Delta t}{\varepsilon}W)^{-1}W\|\leq\gamma^{\prime},

then there exists constant η>0\eta>0, independent of 𝐅,𝐅,I\mathbf{F},\mathbf{F}^{\prime},I and ε\varepsilon such that

𝐅l𝐅l2,Iη𝐆l𝐆l2,I,𝐅l𝐅l1,Iη𝐆l𝐆l1,I,\displaystyle\|\mathbf{F}_{l}-\mathbf{F}^{\prime}_{l}\|_{2,I}\leq\eta\|\mathbf{G}_{l}-\mathbf{G}^{\prime}_{l}\|_{2,I},\quad\|\mathbf{F}_{l}-\mathbf{F}^{\prime}_{l}\|_{-1,I}\leq\eta\|\mathbf{G}_{l}-\mathbf{G}^{\prime}_{l}\|_{-1,I},
Proof.

To prove the lemma, we apply \mathbb{P} to [L1(𝐅)]l=𝐆l[L_{1}(\mathbf{F})]_{l}=\mathbf{G}_{l}, we get

𝐅l=𝐆l+𝐅l(0)ΔtΔxBΛxδlx𝐅(0)ΔtΔyBΛyδly𝐅(0)\mathbb{P}\mathbf{F}_{l}=\mathbb{P}\mathbf{G}_{l}+\mathbb{P}\mathbf{F}_{l}^{(0)}-\frac{\Delta t}{\Delta x}\mathbb{P}B\Lambda_{x}\delta_{l}^{x}\mathbf{F}^{(0)}-\frac{\Delta t}{\Delta y}\mathbb{P}B\Lambda_{y}\delta_{l}^{y}\mathbf{F}^{(0)}

Now, we substitute 𝐅\mathbb{P}\mathbf{F} into the Maxwellian in the equation [L1(𝐅)]l=𝐆l[L_{1}(\mathbf{F})]_{l}=\mathbf{G}_{l}, we have

𝐅l\displaystyle\mathbf{F}_{l} =𝐆l+𝐅l(0)ΔtΔxBΛxδlx𝐅(0)ΔtΔyBΛyδly𝐅(0)\displaystyle=\mathbf{G}_{l}+\mathbf{F}_{l}^{(0)}-\frac{\Delta t}{\Delta x}B\Lambda_{x}\delta_{l}^{x}\mathbf{F}^{(0)}-\frac{\Delta t}{\Delta y}B\Lambda_{y}\delta_{l}^{y}\mathbf{F}^{(0)}
+ΔtεW[𝕄(𝐆l+𝐅l(0)ΔtΔxBΛxδlx𝐅(0)ΔtΔyBΛyδly𝐅(0))𝐅l]+Δtε𝐰0(𝕄(𝐟ln,0)𝐟ln,0)\displaystyle+\frac{\Delta t}{\varepsilon}W\Big{[}\mathbb{M}\big{(}\mathbb{P}\mathbf{G}_{l}+\mathbb{P}\mathbf{F}_{l}^{(0)}-\frac{\Delta t}{\Delta x}\mathbb{P}B\Lambda_{x}\delta_{l}^{x}\mathbf{F}^{(0)}-\frac{\Delta t}{\Delta y}\mathbb{P}B\Lambda_{y}\delta_{l}^{y}\mathbf{F}^{(0)}\big{)}-\mathbf{F}_{l}\Big{]}+\frac{\Delta t}{\varepsilon}\mathbf{w}_{0}\otimes\Big{(}\mathbb{M}(\mathbb{P}\mathbf{f}_{l}^{n,0})-\mathbf{f}_{l}^{n,0}\Big{)}

We can collect all the unknown terms 𝐅l\mathbf{F}_{l} on the left hand side

𝐅l\displaystyle\mathbf{F}_{l} =(IdM×M+ΔtεW)1[𝐆l+𝐅l(0)ΔtΔxBΛxδlx𝐅(0)ΔtΔyBΛyδly𝐅(0)\displaystyle=(\text{Id}_{M\times M}+\frac{\Delta t}{\varepsilon}W)^{-1}\Big{[}\mathbf{G}_{l}+\mathbf{F}_{l}^{(0)}-\frac{\Delta t}{\Delta x}B\Lambda_{x}\delta_{l}^{x}\mathbf{F}^{(0)}-\frac{\Delta t}{\Delta y}B\Lambda_{y}\delta_{l}^{y}\mathbf{F}^{(0)}
+ΔtεW𝕄(𝐆l+𝐅l(0)ΔtΔxBΛxδlx𝐅(0)ΔtΔyBΛyδly𝐅(0))+Δtε𝐰0(𝕄(𝐟ln,0)𝐟ln,0)]\displaystyle+\frac{\Delta t}{\varepsilon}W\mathbb{M}\big{(}\mathbb{P}\mathbf{G}_{l}+\mathbb{P}\mathbf{F}_{l}^{(0)}-\frac{\Delta t}{\Delta x}\mathbb{P}B\Lambda_{x}\delta_{l}^{x}\mathbf{F}^{(0)}-\frac{\Delta t}{\Delta y}\mathbb{P}B\Lambda_{y}\delta_{l}^{y}\mathbf{F}^{(0)}\big{)}+\frac{\Delta t}{\varepsilon}\mathbf{w}_{0}\otimes\big{(}\mathbb{M}(\mathbb{P}\mathbf{f}_{l}^{n,0})-\mathbf{f}_{l}^{n,0}\big{)}\Big{]}

We have

𝐅l𝐅l=(IdM×M+ΔtεW)1(𝐆l𝐆l+ΔtεW𝕄(𝐆l𝐆l)).\mathbf{F}_{l}-\mathbf{F}^{\prime}_{l}=(\text{Id}_{M\times M}+\frac{\Delta t}{\varepsilon}W)^{-1}\Big{(}\mathbf{G}_{l}-\mathbf{G}^{\prime}_{l}+\frac{\Delta t}{\varepsilon}W\mathbb{M}(\mathbb{P}\mathbf{G}_{l}-\mathbb{P}\mathbf{G}^{\prime}_{l})\Big{)}.

This concludes the proof. ∎

Proposition 2.3.

Under the conditions of lemmas 2.1 and 2.2, if 𝐅\mathbf{F}^{*} is the unique solution of L2(𝐅)=0L_{2}(\mathbf{F})=0, there exists θ\theta independent of ε\varepsilon such that we have

𝐅(r+1)𝐅L2(θΔt)r+1𝐅(0)𝐅L2.\|\mathbf{F}^{(r+1)}-\mathbf{F}^{*}\|_{L^{2}}\leq(\theta\Delta t)^{r+1}\|\mathbf{F}^{(0)}-\mathbf{F}^{*}\|_{L^{2}}.
Proof.

We have

L1(𝐅(r+1))L1(𝐅)\displaystyle L_{1}(\mathbf{F}^{(r+1)})-L_{1}(\mathbf{F}^{*}) =L1(𝐅(r))L2(𝐅(r))L1(𝐅)\displaystyle=L_{1}(\mathbf{F}^{(r)})-L_{2}(\mathbf{F}^{(r)})-L_{1}(\mathbf{F}^{*})
=(L1(𝐅(r))L1(𝐅))(L2(𝐅(r))L2(𝐅))\displaystyle=\big{(}L_{1}(\mathbf{F}^{(r)})-L_{1}(\mathbf{F}^{*})\big{)}-\big{(}L_{2}(\mathbf{F}^{(r)})-L_{2}(\mathbf{F}^{*})\big{)}

Also, we can write a Poincaré inequality for φ\varphi as follows

φ2,ICDφ2,I.\|\varphi\|_{2,I}\leq C\|D\varphi\|_{2,I}.

where CC is constant. Using results from Poincaré inequality, lemmas 2.1 and 2.2, the proof is concluded. ∎

Starting from the Chapman-Enskog expansion of the (2a), we can show that our method is asymptotic preserving. We have

𝐟=𝕄(𝐮ε)+𝒪(ε),𝐮εt+𝐀1(𝐮ε)x+𝐀2(𝐮ε)y=𝒪(ε).\begin{split}\mathbf{f}&=\mathbb{M}(\mathbf{u}^{\varepsilon})+\mathcal{O}(\varepsilon),\\ \dfrac{\partial\mathbf{u}^{\varepsilon}}{\partial t}&+\dfrac{\partial\mathbf{A}_{1}(\mathbf{u}^{\varepsilon})}{\partial x}+\dfrac{\partial\mathbf{A}_{2}(\mathbf{u}^{\varepsilon})}{\partial y}=\mathcal{O}(\varepsilon).\end{split} (18)
Proposition 2.4.

(14a) is consistent with the limit model (18) up to an 𝒪(ε)\mathcal{O}(\varepsilon).

Proof.

Let us define 𝐮ε,(r)=𝐅(r)\mathbf{u}^{\varepsilon,(r)}=\mathbb{P}\mathbf{F}^{(r)}, from (12) we get

𝐮ε,(r+1)=𝐅(0)ΔtΔxΛxWδx𝐅(r)ΔtΔx𝐰0Λxδx𝐟n,0ΔtΔyΛyWδy𝐅(r)ΔtΔy𝐰0Λyδy𝐟n,0+𝒪(ε)\mathbf{u}^{\varepsilon,(r+1)}=\mathbb{P}\mathbf{F}^{(0)}-\frac{\Delta t}{\Delta x}\mathbb{P}\Lambda_{x}W\delta^{x}\mathbf{F}^{(r)}-\frac{\Delta t}{\Delta x}\mathbf{w}_{0}\otimes\mathbb{P}\Lambda_{x}\delta^{x}\mathbf{f}^{n,0}-\frac{\Delta t}{\Delta y}\mathbb{P}\Lambda_{y}W\delta^{y}\mathbf{F}^{(r)}-\frac{\Delta t}{\Delta y}\mathbf{w}_{0}\otimes\mathbb{P}\Lambda_{y}\delta^{y}\mathbf{f}^{n,0}+\mathcal{O}(\varepsilon)

Using (3), we have

𝐅(r+1)\displaystyle\mathbf{F}^{(r+1)} =𝕄(𝐅(0)ΔtΔxΛxWδx𝐅(r)ΔtΔx𝐰0Λxδx𝐟n,0\displaystyle=\mathbb{M}\big{(}\mathbb{P}\mathbf{F}^{(0)}-\frac{\Delta t}{\Delta x}\mathbb{P}\Lambda_{x}W\delta^{x}\mathbf{F}^{(r)}-\frac{\Delta t}{\Delta x}\mathbf{w}_{0}\otimes\mathbb{P}\Lambda_{x}\delta^{x}\mathbf{f}^{n,0}
ΔtΔyΛyWδy𝐅(r)ΔtΔy𝐰0Λyδy𝐟n,0)+𝒪(ε)\displaystyle-\frac{\Delta t}{\Delta y}\mathbb{P}\Lambda_{y}W\delta^{y}\mathbf{F}^{(r)}-\frac{\Delta t}{\Delta y}\mathbf{w}_{0}\otimes\mathbb{P}\Lambda_{y}\delta^{y}\mathbf{f}^{n,0}\big{)}+\mathcal{O}(\varepsilon)
=𝕄(𝐮ε,(r+1))+𝒪(ε)\displaystyle=\mathbb{M}(\mathbf{u}^{\varepsilon,(r+1)})+\mathcal{O}(\varepsilon)

Then

𝐮ε,(r+1)\displaystyle\mathbf{u}^{\varepsilon,(r+1)} =𝐮ε,(0)ΔtΔxWδxΛx𝐅(r)ΔtΔx𝐰0δxΛx𝐟n,0ΔtΔyWδyΛy𝐅(r)ΔtΔy𝐰0δyΛy𝐟n,0+𝒪(ε)\displaystyle=\mathbf{u}^{\varepsilon,(0)}-\frac{\Delta t}{\Delta x}W\delta^{x}\mathbb{P}\Lambda_{x}\mathbf{F}^{(r)}-\frac{\Delta t}{\Delta x}\mathbf{w}_{0}\otimes\delta^{x}\mathbb{P}\Lambda_{x}\mathbf{f}^{n,0}-\frac{\Delta t}{\Delta y}W\delta^{y}\mathbb{P}\Lambda_{y}\mathbf{F}^{(r)}-\frac{\Delta t}{\Delta y}\mathbf{w}_{0}\otimes\delta^{y}\mathbb{P}\Lambda_{y}\mathbf{f}^{n,0}+\mathcal{O}(\varepsilon)
=𝐮ε,(0)ΔtΔxWδx𝐀1(𝐮ε,(r))ΔtΔx𝐰0δx𝐀1(𝐮ε,(0))ΔtΔyWδy𝐀2(𝐮ε,(r))\displaystyle=\mathbf{u}^{\varepsilon,(0)}-\frac{\Delta t}{\Delta x}W\delta^{x}\mathbf{A}_{1}(\mathbf{u}^{\varepsilon,(r)})-\frac{\Delta t}{\Delta x}\mathbf{w}_{0}\otimes\delta^{x}\mathbf{A}_{1}(\mathbf{u}^{\varepsilon,(0)})-\frac{\Delta t}{\Delta y}W\delta^{y}\mathbf{A}_{2}(\mathbf{u}^{\varepsilon,(r)})
ΔtΔy𝐰0δy𝐀2(𝐮ε,(0))+𝒪(ε).\displaystyle-\frac{\Delta t}{\Delta y}\mathbf{w}_{0}\otimes\delta^{y}\mathbf{A}_{2}(\mathbf{u}^{\varepsilon,(0)})+\mathcal{O}(\varepsilon).

We observe that if the the spatial discretisation is consistent with the derivative in space, the above formula is the space discretisation and time of the asymptotic model (18), and the result is concluded. ∎

2.3 Time discretisation in the L2L_{2} operator

We consider first, second and fourth order approximation in time in the L2L_{2} operator, when there is no source term.

  1. 1.

    For the first order approximation, we get

    𝐟n,1𝐟n,0+ΔtΔx(Λxδx𝐟n,1)+ΔtΔyΛy(δy𝐟n,1)=0,\mathbb{P}\mathbf{f}^{n,1}-\mathbb{P}\mathbf{f}^{n,0}+\frac{\Delta t}{\Delta x}\mathbb{P}(\Lambda_{x}\delta^{x}\mathbf{f}^{n,1})+\frac{\Delta t}{\Delta y}\Lambda_{y}\mathbb{P}(\delta^{y}\mathbf{f}^{n,1})=0,

    where 𝐟n,0=𝐟n\mathbf{f}^{n,0}=\mathbf{f}^{n} and 𝐟n,1𝐟(tn+1)\mathbf{f}^{n,1}\approx\mathbf{f}(t_{n+1}). The matrix WW becomes W=(1)W=(1). Then, we get

    (Id1×1+ΔtεW)1=εε+Δt,Δtε(Id1×1+ΔtεW)1W=Δtε+Δt,(\text{Id}_{1\times 1}+\frac{\Delta t}{\varepsilon}W)^{-1}=\frac{\varepsilon}{\varepsilon+\Delta t},\quad\frac{\Delta t}{\varepsilon}(\text{Id}_{1\times 1}+\frac{\Delta t}{\varepsilon}W)^{-1}W=\frac{\Delta t}{\varepsilon+\Delta t},

    We observe that these two matrices are uniformly bounded.

  2. 2.

    For the second order approximation, which is Crank-Nicholson method, the scheme becomes

    𝐟n,1𝐟n,0+ΔtΔx(12(Λxδx𝐟n,0)+12(Λxδx𝐟n,1))+ΔtΔy(12(Λyδy𝐟n,0)+12(Λyδy𝐟n,1))=0,\mathbb{P}\mathbf{f}^{n,1}-\mathbb{P}\mathbf{f}^{n,0}+\frac{\Delta t}{\Delta x}\big{(}\frac{1}{2}\mathbb{P}(\Lambda_{x}\delta^{x}\mathbf{f}^{n,0})+\frac{1}{2}\mathbb{P}(\Lambda_{x}\delta^{x}\mathbf{f}^{n,1})\big{)}+\frac{\Delta t}{\Delta y}\big{(}\frac{1}{2}\mathbb{P}(\Lambda_{y}\delta^{y}\mathbf{f}^{n,0})+\frac{1}{2}\mathbb{P}(\Lambda_{y}\delta^{y}\mathbf{f}^{n,1})\big{)}=0,

    also we have W=(12)W=(\frac{1}{2}). Similarly, we see that two matrices (Id1×1+ΔtεW)1(\text{Id}_{1\times 1}+\frac{\Delta t}{\varepsilon}W)^{-1} and Δtε(Id1×1+ΔtεW)1W\frac{\Delta t}{\varepsilon}(\text{Id}_{1\times 1}+\frac{\Delta t}{\varepsilon}W)^{-1}W are uniformly bounded.

  3. 3.

    For the fourth order scheme [20], we get

    𝐟n,1𝐟n,0\displaystyle\mathbb{P}\mathbf{f}^{n,1}-\mathbb{P}\mathbf{f}^{n,0} +ΔtΔx(524(Λxδx𝐟n,0)+13(Λxδx𝐟n,1)124(Λxδx𝐟n,2))\displaystyle+\frac{\Delta t}{\Delta x}\big{(}\frac{5}{24}\mathbb{P}(\Lambda_{x}\delta^{x}\mathbf{f}^{n,0})+\frac{1}{3}\mathbb{P}(\Lambda_{x}\delta^{x}\mathbf{f}^{n,1})-\frac{1}{24}\mathbb{P}(\Lambda_{x}\delta^{x}\mathbf{f}^{n,2})\big{)}
    +ΔtΔy(524(Λyδy𝐟n,0)+13(Λyδy𝐟n,1)124(Λyδy𝐟n,2))=0,\displaystyle+\frac{\Delta t}{\Delta y}\big{(}\frac{5}{24}\mathbb{P}(\Lambda_{y}\delta^{y}\mathbf{f}^{n,0})+\frac{1}{3}\mathbb{P}(\Lambda_{y}\delta^{y}\mathbf{f}^{n,1})-\frac{1}{24}\mathbb{P}(\Lambda_{y}\delta^{y}\mathbf{f}^{n,2})\big{)}=0,
    𝐟n,2𝐟n,0\displaystyle\mathbb{P}\mathbf{f}^{n,2}-\mathbb{P}\mathbf{f}^{n,0} +ΔtΔx(16(Λxδx𝐟n,0)+23(Λxδx𝐟n,1)+16(Λxδx𝐟n,2))\displaystyle+\frac{\Delta t}{\Delta x}\big{(}\frac{1}{6}\mathbb{P}(\Lambda_{x}\delta^{x}\mathbf{f}^{n,0})+\frac{2}{3}\mathbb{P}(\Lambda_{x}\delta^{x}\mathbf{f}^{n,1})+\frac{1}{6}\mathbb{P}(\Lambda_{x}\delta^{x}\mathbf{f}^{n,2})\big{)}
    +ΔtΔy(16(Λyδy𝐟n,0)+23(Λyδy𝐟n,1)+16(Λyδy𝐟n,2))=0,\displaystyle+\frac{\Delta t}{\Delta y}\big{(}\frac{1}{6}\mathbb{P}(\Lambda_{y}\delta^{y}\mathbf{f}^{n,0})+\frac{2}{3}\mathbb{P}(\Lambda_{y}\delta^{y}\mathbf{f}^{n,1})+\frac{1}{6}\mathbb{P}(\Lambda_{y}\delta^{y}\mathbf{f}^{n,2})\big{)}=0,

    where 𝐟n,0=𝐟n\mathbf{f}^{n,0}=\mathbf{f}^{n}, 𝐟n,1𝐟(tn+Δt2)\mathbf{f}^{n,1}\approx\mathbf{f}(t_{n}+\frac{\Delta t}{2}) and 𝐟n,2𝐟(tn+1)\mathbf{f}^{n,2}\approx\mathbf{f}(t_{n+1}). Also, we have

    W=(131242316)W=\begin{pmatrix}\frac{1}{3}&-\frac{1}{24}\\ \frac{2}{3}&\frac{1}{6}\end{pmatrix}

    It is easy to observe that the matrix Id2×2+ΔtεW\text{Id}_{2\times 2}+\frac{\Delta t}{\varepsilon}W is invertible and the matrices (Id2×2+ΔtεW)1(\text{Id}_{2\times 2}+\frac{\Delta t}{\varepsilon}W)^{-1} and Δtε(Id2×2+ΔtεW)1W\frac{\Delta t}{\varepsilon}(\text{Id}_{2\times 2}+\frac{\Delta t}{\varepsilon}W)^{-1}W are uniformly bounded.

3 Space discretisation

3.1 Definition of the δx\delta^{x} and δy\delta^{y} operators

Since Λx\Lambda_{x} and Λy\Lambda_{y} are diagonal matrices, we can consider the scalar transport equation

ft+afx+bfy=0,\dfrac{\partial f}{\partial t}+a\dfrac{\partial f}{\partial x}+b\dfrac{\partial f}{\partial y}=0,

where aa and bb are constants and both of them can not be zero at the same time. Next, we will discuss the approximation of flx\dfrac{\partial f_{l}}{\partial x}, the approximation for fly\dfrac{\partial f_{l}}{\partial y} is obtained in a similar manner. In [21] has been developed a stable numerical finite difference methods for first-order hyperbolics in one dimension, which use ss forward and rr backward steps in the discretisation of the space derivatives that are of order at most 2min{r+1,s}2\min\{r+1,s\}. These methods are based on the approximation of fx(xi,yj)\dfrac{\partial f}{\partial x}(x_{i},y_{j}), i,ji,j\in\mathbb{Z}, by a finite difference

1Δxk=rsαkfi+k,j\frac{1}{\Delta x}\sum\limits_{k=-r}^{s}\alpha_{k}f_{i+k,j}

when a>0a>0 and we say that the method is of the class {r,s}\{r,s\}. If a<0a<0, we set

1Δxk=srαkfi+k,j\frac{1}{\Delta x}\sum\limits_{k=-s}^{r}\alpha_{-k}f_{i+k,j}

Throughout this section we suppose without loss of generality that a>0a>0. Otherwise the roles of rr and ss are reversed. We call an {r,s}\{r,s\} method of the highest order an interpolatory method, and is denoted by [r.s][r.s]. Following the theorem of [21], for all integers r,s0r,s\geq 0 the order of the interpolatory method [r.s][r.s] is q=r+sq=r+s, i.e.

δxfi,jΔxfx(xi,yj)=cr,s(Δx)qq+1fxq+1(xi,yj)+𝒪((Δx)q+1)\frac{\delta^{x}f_{i,j}}{\Delta x}-\dfrac{\partial f}{\partial x}(x_{i},y_{j})=c_{r,s}(\Delta x)^{q}\frac{\partial^{q+1}f}{\partial x^{q+1}}(x_{i},y_{j})+\mathcal{O}((\Delta x)^{q+1})

where the error constant cr,sc_{r,s} is defined by

cr,s=(1)s1r!s!(r+s+1)!c_{r,s}=\frac{(-1)^{s-1}r!s!}{(r+s+1)!}

The coefficients are defined by

αk\displaystyle\alpha_{k} =(1)k+1kr!s!(r+k)!(sk)!,rks,k0,\displaystyle=\frac{(-1)^{k+1}}{k}\frac{r!s!}{(r+k)!(s-k)!},\quad-r\leq k\leq s,\quad k\neq 0,
α0\displaystyle\alpha_{0} =k=r,k0sαk,\displaystyle=-\sum\limits_{k=-r,k\neq 0}^{s}\alpha_{k},

We recall that the [r,r][r,r], [r,r+1][r,r+1] and [r,r+2][r,r+2] schemes are the only stable interpolatory methods, and we will only consider these approximations. Before we proceed to the possible choices for δ\delta, it is important to remark that, we can write

δxfi,j=f^i+12,jf^i12,j\delta^{x}f_{i,j}=\hat{f}_{i+\frac{1}{2},j}-\hat{f}_{i-\frac{1}{2},j}

where

f^i+12,j=k=s+1rβkfi+k,j\hat{f}_{i+\frac{1}{2},j}=\sum\limits_{k=-s+1}^{r}\beta_{k}f_{i+k,j}

with βk=mkαm\beta_{k}=\sum\limits_{m\geq k}\alpha_{m}. In the following, we consider first, second and fourth order approximation in x, that in y is done in a similar manner.

  1. 1.

    First order: Here, we must have r=0r=0, s=1s=1. We get the upwind scheme

    δ1xfi,j=fi,jfi1,j\delta_{1}^{x}f_{i,j}=f_{i,j}-f_{i-1,j}

    Then, we have

    fx(xi,yj)=1Δx(fi,jfi1,j)+c0,1Δx2fx2(xi,yj)+𝒪((Δx)2)\dfrac{\partial f}{\partial x}(x_{i},y_{j})=\frac{1}{\Delta x}(f_{i,j}-f_{i-1,j})+c_{0,1}\Delta x\frac{\partial^{2}f}{\partial x^{2}}(x_{i},y_{j})+\mathcal{O}((\Delta x)^{2})

    and the flux is given by

    f^i+12,j=fi+1,j.\hat{f}_{i+\frac{1}{2},j}=f_{i+1,j}.
  2. 2.

    Second order case. Two cases are possible

    • r=s=1r=s=1: centered case

      δ2xfi,j=12(fi+1,jfi1,j)\delta_{2}^{x}f_{i,j}=\frac{1}{2}\big{(}f_{i+1,j}-f_{i-1,j}\big{)}

      and the flux is

      f^i+1/2,j=fi+1,j+fi,j2.\hat{f}_{i+1/2,j}=\frac{f_{i+1,j}+f_{i,j}}{2}.
    • r=0r=0, s=2s=2. There

      δ2xf=32fi,j2fi1,j+12fi2,j\delta_{2}^{x}f=\frac{3}{2}f_{i,j}-2f_{i-1,j}+\frac{1}{2}f_{i-2,j}

      and the flux is

      f^i+1/2,j=32fi,j12fi1,j.\hat{f}_{i+1/2,j}=\frac{3}{2}f_{i,j}-\frac{1}{2}f_{i-1,j}.

    The centered case will no longer be considered.

  3. 3.

    Third order: Only choice is possible r=1r=1, s=2s=2 and we get

    δ3xfi,j=fi2,j6fi1,j+fi,j2+fi+1,j3\delta_{3}^{x}f_{i,j}=\frac{f_{i-2,j}}{6}-f_{i-1,j}+\frac{f_{i,j}}{2}+\frac{f_{i+1,j}}{3}

    Therefore we have

    fx(xi,yj)=1Δx(fi2,j6fi1,j+fi,j2+fi+1,j3)+c2,1(Δx)34fx4(xi,yj)+𝒪((Δx)4)\dfrac{\partial f}{\partial x}(x_{i},y_{j})=\frac{1}{\Delta x}(\frac{f_{i-2,j}}{6}-f_{i-1,j}+\frac{f_{i,j}}{2}+\frac{f_{i+1,j}}{3})+c_{2,1}(\Delta x)^{3}\frac{\partial^{4}f}{\partial x^{4}}(x_{i},y_{j})+\mathcal{O}((\Delta x)^{4})

    As before, the flux is given by

    f^i+12,j=fi1,j6+56fi,j+fi+1,j3.\hat{f}_{i+\frac{1}{2},j}=-\frac{f_{i-1,j}}{6}+\frac{5}{6}f_{i,j}+\frac{f_{i+1,j}}{3}.
  4. 4.

    Fourth order: we consider two cases

    • If r=s=2r=s=2, we have

      δ4xfi,j=fi2,j1223fi1,j+23fi+1,jfi+2,j12\delta_{4}^{x}f_{i,j}=\frac{f_{i-2,j}}{12}-\frac{2}{3}f_{i-1,j}+\frac{2}{3}f_{i+1,j}-\frac{f_{i+2,j}}{12}

      We can write

      fx(xi,yj)=1Δx(fi2,j1223fi1,j+23fi+1,jfi+2,j12)+c2,2(Δx)45fx5(xi,yj)+𝒪((Δx)5)\dfrac{\partial f}{\partial x}(x_{i},y_{j})=\frac{1}{\Delta x}(\frac{f_{i-2,j}}{12}-\frac{2}{3}f_{i-1,j}+\frac{2}{3}f_{i+1,j}-\frac{f_{i+2,j}}{12})+c_{2,2}(\Delta x)^{4}\frac{\partial^{5}f}{\partial x^{5}}(x_{i},y_{j})+\mathcal{O}((\Delta x)^{5})

      For the flux, we can write as follows

      f^i+12,j=fi1,j12+34fi,j+34fi+1,j+fi+2,j12.\hat{f}_{i+\frac{1}{2},j}=\frac{f_{i-1,j}}{12}+\frac{3}{4}f_{i,j}+\frac{3}{4}f_{i+1,j}+\frac{f_{i+2,j}}{12}.
    • If r=1r=1 and s=3s=3, we have

      δ4xfi,j=fi3,j12+fi2,j232fi1,j+56fi,j+fi+1,j4\delta_{4}^{x}f_{i,j}=-\frac{f_{i-3,j}}{12}+\frac{f_{i-2,j}}{2}-\frac{3}{2}f_{i-1,j}+\frac{5}{6}f_{i,j}+\frac{f_{i+1,j}}{4}

      Hence

      fx(xi,yj)=1Δx(fi3,j12+fi2,j232fi1,j+56fi,j+fi+1,j4)+c1,3(Δx)45fx5(xi,yj)+𝒪((Δx)5)\dfrac{\partial f}{\partial x}(x_{i},y_{j})=\frac{1}{\Delta x}(-\frac{f_{i-3,j}}{12}+\frac{f_{i-2,j}}{2}-\frac{3}{2}f_{i-1,j}+\frac{5}{6}f_{i,j}+\frac{f_{i+1,j}}{4})+c_{1,3}(\Delta x)^{4}\frac{\partial^{5}f}{\partial x^{5}}(x_{i},y_{j})+\mathcal{O}((\Delta x)^{5})

      The flux becomes

      f^i+12,j=fi2,j12512fi1,j+1312fi,j+fi+1,j4.\hat{f}_{i+\frac{1}{2},j}=\frac{f_{i-2,j}}{12}-\frac{5}{12}f_{i-1,j}+\frac{13}{12}f_{i,j}+\frac{f_{i+1,j}}{4}.

      We will only use the case r=1r=1, s=3s=3 because the case r=2r=2, s=2s=2 is centered.

3.2 Limitation

We have explored two ways to introduce a nonlinear stabilisation mechanism. The first one is inspired from [22, 23, 24] and consists in ”limiting” the flux, the second one is inspired by the MOOD paradigm [25, 26].

3.2.1 Limitation of the flux

In this section, we only consider the space discretisation in x, that in y is done in a similar manner. First, we calculate the difference between first and second order approximation in x. We have

δ2xfi,jδ1xfi,j=16(fi2,j3fi,j+2fi+1,j)\delta_{2}^{x}f_{i,j}-\delta_{1}^{x}f_{i,j}=\frac{1}{6}(f_{i-2,j}-3f_{i,j}+2f_{i+1,j})

So for limitation

δ2x~fi,j\displaystyle\widetilde{\delta_{2}^{x}}f_{i,j} =δ1xfi,j+θ6(δ2xfi,jδ1xfi,j)\displaystyle=\delta_{1}^{x}f_{i,j}+\frac{\theta}{6}(\delta_{2}^{x}f_{i,j}-\delta_{1}^{x}f_{i,j})
=δ1xfi,j+θ6(fi2,j3fi,j+2fi+1,j)\displaystyle=\delta_{1}^{x}f_{i,j}+\frac{\theta}{6}(f_{i-2,j}-3f_{i,j}+2f_{i+1,j})
=δ1xfi,j(1+θr6)\displaystyle=\delta_{1}^{x}f_{i,j}(1+\frac{\theta r}{6})

where θ[0,1]\theta\in[0,1] and

r=(fi2,j3fi,j+2fi+1,j)fi,jfi1,j=12fi+1,j+4fi,jfi1,jfi2,jfi,jfi1,jr=\frac{(f_{i-2,j}-3f_{i,j}+2f_{i+1,j})}{f_{i,j}-f_{i-1,j}}=1-\frac{-2f_{i+1,j}+4f_{i,j}-f_{i-1,j}-f_{i-2,j}}{f_{i,j}-f_{i-1,j}}

We can see that if ff is smooth, r0r\approx 0. To have a monotonicity condition, we also need

6+θr06+\theta r\geq 0

We want to find θ(r)\theta(r) such that θ(0)=1\theta(0)=1 and

01+θ(r)r6M0\leq 1+\frac{\theta(r)r}{6}\leq M

where MM is a constant that will dictate the CFL constraint. Since θ(0)=1\theta(0)=1, we observe that M1M\geq 1 is a necessary condition. One solution is to choose a α]0,min{6,6(M1)}[\alpha\in]0,\min\{6,6(M-1)\}[. We take

θ(r)={1if r[6+α,6(M1)α],6+αrif r<6+θ,6(M1)αrif r>6(M1)+α.\theta(r)=\begin{cases}1&\text{if }r\in[-6+\alpha,6(M-1)-\alpha],\\ \frac{-6+\alpha}{r}&\text{if }r<-6+\theta,\\ \frac{6(M-1)-\alpha}{r}&\text{if }r>6(M-1)+\alpha.\end{cases}

After calculations, we get

δ2x~=δ1x(1+θr)=δ1x+ψ(δ1x,δ2x)(δ2xδ1x)\widetilde{\delta_{2}^{x}}=\delta_{1}^{x}(1+\theta r)=\delta_{1}^{x}+\psi(\delta_{1}^{x},\delta_{2}^{x})(\delta_{2}^{x}-\delta_{1}^{x})

where

ψ(r,s)(sr)={srif sr[6+α,6(M1)α],(6+α)rif sr<6+θ,(6(M1)α)rif sr>6(M1)+α.\psi(r,s)(s-r)=\begin{cases}s-r&\text{if }\frac{s}{r}\in[-6+\alpha,6(M-1)-\alpha],\\ (-6+\alpha)r&\text{if }\frac{s}{r}<-6+\theta,\\ (6(M-1)-\alpha)r&\text{if }\frac{s}{r}>6(M-1)+\alpha.\end{cases}

For the first and fourth order approximations, it is possible to follow the same technique.

The main drawback of this approach is that the projection on the conservative variable plays no role, while the variable we are really interested in are these variables …so that we can expect some weaknesses.

3.2.2 Stabilisation by the MOOD technique

This is inspired from [25, 26], the variables on which we test are the physical variables (ρ,𝐯,p)(\rho,\mathbf{v},p). This section also justifies the way that have written the update of the spatial term as in (15), i.e. a sum of residual that are evaluated on the elements Ki+1/2.j+1/2=[xi,xi+1]×[yj,yj+1]K_{i+1/2.j+1/2}=[x_{i},x_{i+1}]\times[y_{j},y_{j+1}].

Starting from 𝐅n\mathbf{F}^{n}, we first compute 𝐅n+1~\widetilde{\mathbf{F}^{n+1}} with the full order method (i.e. full order in time and space). In the numerical examples, we will take the fourth order accurate method, but other choices can be made. The algorithm is as follow: We define a vector of logical Flagp\texttt{Flag}_{p} which is false initially for all the grid points, and a vector of logical Flage\texttt{Flag}_{e} which is also set to false.

  1. 1.

    For each mesh point (xi,yj)(x_{i},y_{j}), we compute V~ij=(ρ,𝐯,p)~n+1\tilde{V}_{ij}=\widetilde{(\rho,\mathbf{v},p)}^{n+1} variables defined by 𝐅n+1~\mathbb{P}\widetilde{\mathbf{F}^{n+1}}. If ρ~ijn+10\tilde{\rho}^{n+1}_{ij}\leq 0 or p~ijn+10\tilde{p}^{n+1}_{ij}\leq 0 or one of the components of V~ij\tilde{V}_{ij} are NaN 111xx is not a number if xxx\neq x., we set Flagp(i,j)=.true.\texttt{Flag}_{p}(i,j)=.true.

  2. 2.

    Then we loop over the quads [xi,xi+1]×[yj,yj+1][x_{i},x_{i+1}]\times[y_{j},y_{j+1}]. If for one of the 4 corners (xl,yq)(x_{l},y_{q}), Flagp(l,q)=.true.\texttt{Flag}_{p}(l,q)=.true., we set Flage(i+1/2,j+1/2)=.true.\texttt{Flag}_{e}(i+1/2,j+1/2)=.true.

  3. 3.

    For each element such that Flage(i+1/2,j+1/2)=.true.\texttt{Flag}_{e}(i+1/2,j+1/2)=.true., we recompute, for each sub-time step the four residuals defined by (15b)

In [25, 26] is also a way to detect local extremas, and in [26] to differenciate the local smooth extremas from the discontinuities. We have not use this here, and there is way of improvement. The first order version of our scheme amounts to global Lax-Friedrichs. To make sure that the first order scheme is domain invariant preserving, we can apply this procedure to each of the cycle of the DeC procedure, we have not done that in the numerical examples.

4 Stability analysis

We will study the stability of the discretisation of the homogeneous problem. As discussed in the previous section, since the matrices Λx\Lambda_{x} and Λy\Lambda_{y} are diagonal, it is enough to consider again the following transport equation

ft+afx+bfy=0,\dfrac{\partial f}{\partial t}+a\dfrac{\partial f}{\partial x}+b\dfrac{\partial f}{\partial y}=0, (19)

Since ab=0ab=0 in the case of four waves model, we have the same results as one dimensional case [1]. In other cases, we perform the Fourier analysis to evaluate the amplification factors of the method, first without defect correction iteration, then with defect correction iteration. We assume, without loss of generality, that a,b>0a,b>0. we denote Fourier symbol of δx\delta^{x} and δy\delta^{y} as g1g_{1} and g2g_{2}, respectively. For δx\delta^{x} and δy\delta^{y} operators, we considered four cases in the previous section, we have

  • First order in both xx and yy: we have g(1)(θ)=1eiθg^{(1)}(\theta)=1-e^{i\theta} and g1=g(1)(θ1)g_{1}=g^{(1)}(\theta_{1}), g2=g(1)(θ2)g_{2}=g^{(1)}(\theta_{2}) We see that (g(1))0\Re(g^{(1)})\geq 0 and maxθ|g(1)|=4\max_{\theta}|g^{(1)}|=4.

  • Second order in xx and yy: g1=g(2)(θ1)g_{1}=g^{(2)}(\theta_{1}) and g2=g(2)(θ2)g_{2}=g^{(2)}(\theta_{2}) where

    g(2)(θ)=322eiθ+e2iθ2.g^{(2)}(\theta)=\frac{3}{2}-2e^{-i\theta}+\frac{e^{-2i\theta}}{2}.

    We notice that (g(2))=(cosθ1)20\Re(g^{(2)})=\big{(}\cos\theta-1\big{)}^{2}\geq 0 and maxθ|g(2)|=2\max_{\theta}|g^{(2)}|=2.

  • Third order in both xx and yy: g1=g(3)(θ1)g_{1}=g^{(3)}(\theta_{1}) and g2=g(3)(θ2)g_{2}=g^{(3)}(\theta_{2}) where

    g(3)=e2iθ6eiθ+12+eiθ3.g^{(3)}=\frac{e^{-2i\theta}}{6}-e^{-i\theta}+\frac{1}{2}+\frac{e^{i\theta}}{3}.

    Again, (g(3))=13(cosθ1)20\Re(g^{(3)})=\frac{1}{3}\big{(}\cos\theta-1\big{)}^{2}\geq 0 and maxθ|g(3)|=32\max_{\theta}|g^{(3)}|=\tfrac{3}{2}

  • Fourth order in both xx and yy: we only consider the case r=1r=1, s=2s=2. We have g1=g(4)(θ)g_{1}=g^{(4)}(\theta) and g2=g(4)(θ2)g_{2}=g^{(4)}(\theta_{2}) where

    g(4)(θ1)=e3iθ12+e2iθ232eiθ+56+eiθ4.g^{(4)}(\theta-1)=-\frac{e^{-3i\theta}}{12}+\frac{e^{-2i\theta}}{2}-\frac{3}{2}e^{-i\theta}+\frac{5}{6}+\frac{e^{i\theta}}{4}.

    and we have (g(4))=13(1cosθ)30\Re(g^{(4)})=\frac{1}{3}\big{(}1-\cos\theta\big{)}^{3}\geq 0 and maxθ|g(4)|=83\max_{\theta}|g^{(4)}|=\frac{8}{3}

Now, we consider first, second and fourth order approximations in time in the L2L_{2} operator. In the sequel, we set g=μg1+νg2g=\mu g_{1}+\nu g_{2} with μ=aΔtΔx\mu=a\frac{\Delta t}{\Delta x} and ν=bΔtΔy\nu=b\frac{\Delta t}{\Delta y}.

  1. 1.

    First order in time: Using Fourier transform, the L2L_{2} operator can be written as follows

    f^n+1f^n+μg1f^n+1+νg2f^n+1=0.\hat{f}^{n+1}-\hat{f}^{n}+\mu g_{1}\hat{f}^{n+1}+\nu g_{2}\hat{f}^{n+1}=0.

    The amplification factor is

    G1=11+gG_{1}=\frac{1}{1+g}

    In order to have a stable scheme for the first order scheme, we should have |G|1\lvert G\rvert\leq 1, and a necessary and sufficient condition is 2(g)+|g|202\Re(g)+|g|^{2}\geq 0.

    The defect correction iteration is written as

    f^(r+1)=f^nμg1f^(r)νg2f^(r)\hat{f}^{(r+1)}=\hat{f}^{n}-\mu g_{1}\hat{f}^{(r)}-\nu g_{2}\hat{f}^{(r)}

    The resulting formula for the amplification factor Gr+1G_{r+1} is given by

    G1,0\displaystyle G_{1,0} =1\displaystyle=1
    G,1r+1(g)\displaystyle G_{,1r+1}(g) =1gG1,r(g)\displaystyle=1-gG_{1,r}(g)

    We can observe that

    G1,r+1(g)G1(g)=(1)r+1gr+1(1G1(g)).G_{1,r+1}(g)-G_{1}(g)=(-1)^{r+1}g^{r+1}\big{(}1-G_{1}(g)\big{)}. (20)

    We note that gr+10g^{r+1}\rightarrow 0 if |g|<1|g|<1.

  2. 2.

    Second order in time: We again use Fourier transform, and write L2L_{2} as

    f^n+1f^n+μ2(g1f^n+g1f^n+1)+ν2(g2f^n+g2f^n+1)=0\hat{f}^{n+1}-\hat{f}^{n}+\frac{\mu}{2}(g_{1}\hat{f}^{n}+g_{1}\hat{f}^{n+1})+\frac{\nu}{2}(g_{2}\hat{f}^{n}+g_{2}\hat{f}^{n+1})=0

    In this case the amplification factor is

    G2=1g21+g2G_{2}=\frac{1-\frac{g}{2}}{1+\frac{g}{2}}

    We conclude that under the following condition stability holds

    Re(g)0\text{Re}(g)\geq 0

    The defect correction iteration reads

    f^(r+1)=f^nμ2(g1f^n+g1f^(r))ν2(g2f^n+g2f^(r))\hat{f}^{(r+1)}=\hat{f}^{n}-\frac{\mu}{2}(g_{1}\hat{f}^{n}+g_{1}\hat{f}^{(r)})-\frac{\nu}{2}(g_{2}\hat{f}^{n}+g_{2}\hat{f}^{(r)})

    we have

    G2,0\displaystyle G_{2,0} =1\displaystyle=1
    G2,r+1(g)\displaystyle G_{2,r+1}(g) =1g2g2G2,r(g)\displaystyle=1-\frac{g}{2}-\frac{g}{2}G_{2,r}(g)

    It is easy to check that

    G2,r+1(g)G3(g)=(1)r+1(g2)r+1(1G2(g)).G_{2,r+1}(g)-G_{3}(g)=(-1)^{r+1}\big{(}\frac{g}{2}\big{)}^{r+1}\big{(}1-G_{2}(g)\big{)}. (21)

    We note that (g2)r+10\big{(}\frac{g}{2}\big{)}^{r+1}\rightarrow 0 if |g|g2|g|g\leq 2.

  3. 3.

    Fourth order in time: Similarly, we have the following formula for L2L_{2} operator

    f^n+12f^n+μ(524g1f^n+13g1f^n+12124g1f^n+1)+ν(524g2f^n+13g2f^n+12124g2f^n+1)=0f^n+1f^n+μ(16g1f^n+23g1f^n+12+16g1f^n+1)+ν(16g2f^n+23g2f^n+12+16g2f^n+1)=0\begin{split}\hat{f}^{n+\frac{1}{2}}&-\hat{f}^{n}+\mu(\frac{5}{24}g_{1}\hat{f}^{n}+\frac{1}{3}g_{1}\hat{f}^{n+\frac{1}{2}}-\frac{1}{24}g_{1}\hat{f}^{n+1})+\nu(\frac{5}{24}g_{2}\hat{f}^{n}+\frac{1}{3}g_{2}\hat{f}^{n+\frac{1}{2}}-\frac{1}{24}g_{2}\hat{f}^{n+1})=0\\ \hat{f}^{n+1}&-\hat{f}^{n}+\mu(\frac{1}{6}g_{1}\hat{f}^{n}+\frac{2}{3}g_{1}\hat{f}^{n+\frac{1}{2}}+\frac{1}{6}g_{1}\hat{f}^{n+1})+\nu(\frac{1}{6}g_{2}\hat{f}^{n}+\frac{2}{3}g_{2}\hat{f}^{n+\frac{1}{2}}+\frac{1}{6}g_{2}\hat{f}^{n+1})=0\end{split} (22)

    We can rewrite (22) in matrix form

    (f^n+12f^n+1)=G4(f^nf^n)\begin{pmatrix}\hat{f}^{n+\frac{1}{2}}\\ \hat{f}^{n+1}\end{pmatrix}=G_{4}\begin{pmatrix}\hat{f}^{n}\\ \hat{f}^{n}\end{pmatrix}

    where, setting θ=24+12g+2g2\theta=24+12g+2g^{2}

    G4(g)=(1+g3g242g31+g6)1(15g241g6)=12θ(24g22412g+2g2)\begin{split}G_{4}(g)&=\begin{pmatrix}1+\frac{g}{3}&-\frac{g}{24}\\ \frac{2g}{3}&1+\frac{g}{6}\end{pmatrix}^{-1}\begin{pmatrix}1-\frac{5g}{24}\\ 1-\frac{g}{6}\end{pmatrix}\\ &\quad=\frac{1}{2\theta}\begin{pmatrix}{24-g^{2}}\\ {24-12g+2g^{2}}\\ \end{pmatrix}\end{split}

    In order to have a stable scheme, one should have max{|G1|,|G2|}1\max\{\lvert G_{1}\rvert,\lvert G_{2}\rvert\}\leq 1. The defect correction iteration reads

    h^1(r+1)=f^nμ(524g1f^n+13g1h^1(r)124g1h^2(r))ν(524g2f^n+13g2h^1(r)124g2h^2(r))h^2(r+1)=f^nμ(16g1f^n+23g1h^1(r)+16g1h^2(r))ν(16g2f^n+23g2h^1(r)+16g2h^2(r))\begin{split}\hat{h}_{1}^{(r+1)}&=\hat{f}^{n}-\mu(\frac{5}{24}g_{1}\hat{f}^{n}+\frac{1}{3}g_{1}\hat{h}_{1}^{(r)}-\frac{1}{24}g_{1}\hat{h}_{2}^{(r)})-\nu(\frac{5}{24}g_{2}\hat{f}^{n}+\frac{1}{3}g_{2}\hat{h}_{1}^{(r)}-\frac{1}{24}g_{2}\hat{h}_{2}^{(r)})\\ \hat{h}_{2}^{(r+1)}&=\hat{f}^{n}-\mu(\frac{1}{6}g_{1}\hat{f}^{n}+\frac{2}{3}g_{1}\hat{h}_{1}^{(r)}+\frac{1}{6}g_{1}\hat{h}_{2}^{(r)})-\nu(\frac{1}{6}g_{2}\hat{f}^{n}+\frac{2}{3}g_{2}\hat{h}_{1}^{(r)}+\frac{1}{6}g_{2}\hat{h}_{2}^{(r)})\end{split} (23)

    Rewriting (23) in matrix form, we obtain

    h^(r+1)=(15g241g6)gMh^(r)\hat{h}^{(r+1)}=\begin{pmatrix}1-\frac{5g}{24}\\ 1-\frac{g}{6}\end{pmatrix}-gM\hat{h}^{(r)}

    where M=(131242316)M=\begin{pmatrix}\frac{1}{3}&\frac{-1}{24}\\ \frac{2}{3}&\frac{1}{6}\end{pmatrix}. We get

    G4,0\displaystyle G_{4,0} =(11)\displaystyle=\begin{pmatrix}1\\ 1\end{pmatrix}
    G4,r+1(g)\displaystyle G_{4,r+1}(g) =(15g241g6)gMG4,r(g)\displaystyle=\begin{pmatrix}1-\frac{5g}{24}\\ 1-\frac{g}{6}\end{pmatrix}-gMG_{4,r}(g)

    Hence

    G4,r+1(g)G4(g)=(1)r+1(g)r+1Mr+1((11)G4(g)).G_{4,r+1}(g)-G_{4}(g)=(-1)^{r+1}(g)^{r+1}M^{r+1}\Bigg{(}\begin{pmatrix}1\\ 1\end{pmatrix}-G_{4}(g)\Bigg{)}. (24)

    We note that (1)r+1(g)r+1Mr+10(-1)^{r+1}(g)^{r+1}M^{r+1}\rightarrow 0 if |g|M2<1|g|\|M\|_{2}<1, i.e. if |g|(337+1043531152)11.745356304|g|\leq\big{(}\tfrac{337+\sqrt{104353}}{1152}\big{)}^{-1}\approx 1.745356304.

Using this relations, we have plotted the zone 𝒜G\mathcal{A}_{G} where |G|1|G|\leq 1 for the second order scheme (Crank Nicholson with DeC iteration) and max(|G(1)|,|G(2)|)<1\max(|G(1)|,|G(2)|)<1. To get a stable scheme, we need that part of 𝒜G\mathcal{A}_{G} has a non empty intersection with {(x,y),x0}\{(x,y),x\geq 0\}. We have plotted on Figure 2 𝒜G\mathcal{A}_{G} for the second order Dec with 3 and 4 iterations, and for the 4th order DeC with 4 and 5 iterations.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 2: Plots of the stability region with the circle of radius ρ\rho for DeC 2nd order (a): ρ22=0.4\rho_{2}^{2}=0.4, 2 iterations, (b): ρ32=3\rho_{3}^{2}=\sqrt{3}, 3 iterations, (c): ρ42=2.2\rho_{4}^{2}=\sqrt{2.2}, 4 iterations and DeC 4th order (d): ρ44=6.8\rho_{4}^{4}=\sqrt{6.8}, 4 iterations, (e): ρ53=3\rho_{5}^{3}=3, 5 iterations, (f): ρ63=22\rho_{6}^{3}=2\sqrt{2}, 6 iterations.

We see that the larger ρ\rho is achieved for r+1r+1 iterations for scheme of order rr.

From (20)-(24), the amplification factor of the two dimensional scheme is Gkr(νg(θ1)+μg(θ2))G_{k}^{r}\big{(}\nu g(\theta_{1})+\mu g(\theta_{2})\big{)}, so we need to check when |νg(θ1)+μg(θ2)|ρkr|\nu g(\theta_{1})+\mu g(\theta_{2})|\leq\rho_{k}^{r}. Since |νg(θ1)+μg(θ2)|2max(|μ|,|ν|)maxθ|g(θ)|||\nu g(\theta_{1})+\mu g(\theta_{2})|\leq 2\max(|\mu|,|\nu|)\max_{\theta}|g(\theta)||, we display max(|μ|,|ν|)\max(|\mu|,|\nu|) in Table 1.

DeC(2,2) DeC(2,3) DeC(2,4)
g(2)g^{(2)} 0.40.4 2/31.152/\sqrt{3}\approx 1.15 2.21.58\sqrt{2.2}\approx 1.58
DeC(4,4) DeC(4,5) DeC(4,6)
g(4)g^{(4)} 0.756.81.950.75\sqrt{6.8}\approx 1.95 2.252.25 1.522.121.5\sqrt{2}\approx 2.12
Table 1: CFL condition for the scheme DeC(r,p): r is the order, p the number of iteraion

Using this amplification factors, we obtain the maximum values of the CFL number for which the various options leads to stability. Even with the BGK source term, the schemes are always stable for CFL=1+δCFL=1+\delta, δ>0\delta>0. For example, the combination 4th order in time/4th ordre in space is stable for CFL1.3CFL\approx 1.3, and this can be checked experimentally on the vortex case bellow, with periodic boundary conditions. For that case, we have not ben able to run higher that CFL 1.3, it may be an effect of the BGK source term. This results are also consistent with [1].

5 Test cases

In this section we will present the numerical results that illustrate the behavior of our scheme on scalar problems and Euler equations. To evaluate the accuracy and robustness of the proposed high order asymptotic preserving kinetic scheme, in the following we perform the convergence analysis for the scalar problems and study several benchmark problems for the Euler equations of gas dynamics.

5.1 Scalar problems

Consider the two-dimensional advection equation:

ut+ux+uy=0,(x,y,t)[2,2]×[2,2]×+,\frac{\partial u}{\partial t}+\dfrac{\partial u}{\partial x}+\dfrac{\partial u}{\partial y}=0,~{}(x,y,t)\in[-2,2]\times[-2,2]\times\mathbb{R}^{+},

and periodic boundary conditions. We consider the following initial condition:

u0(x,y)=sin(πx+πy),(x,y)(2,2)×(2,2).u_{0}(x,y)=\sin(\pi x+\pi y),~{}(x,y)\in(-2,2)\times(-2,2).

The CFL number is set to 1. The convergence for the density is shown in Tables 2 and 3 for final time T=10T=10 for orders 2 and 4, which result in the predicted convergence rates of second and fourth order, respectively.

Table 2: Convergence study for the advection equation for order 2 at T=10T=10.
h L1L^{1}-error slope L2L^{2}-error slope LL^{\infty}-error slope
0.05 1.0698.10+11.0698.10^{+1} - 2.8479.1002.8479.10^{0} - 9.3878.1019.3878.10^{-1} -
0.025 3.5595.1003.5595.10^{0} 1.59 9.6212.1019.6212.10^{-1} 1.57 3.3039.1013.3039.10^{-1} 1.51
0.0125 6.8578.1016.8578.10^{-1} 2.38 1.8812.1011.8812.10^{-1} 2.35 6.5662.1026.5662.10^{-2} 2.33
0.00625 1.4701.1011.4701.10^{-1} 2.22 4.0558.1024.0558.10^{-2} 2.21 1.4243.1021.4243.10^{-2} 2.20
0.003125 3.4890.1023.4890.10^{-2} 2.08 9.6578.1039.6578.10^{-3} 2.07 3.4037.1033.4037.10^{-3} 2.07
Table 3: Convergence study for the advection equation for order 4 at T=10T=10.
h L1L^{1}-error slope L2L^{2}-error slope LL^{\infty}-error slope
0.05 4.7601.1004.7601.10^{0} - 1.2919.1001.2919.10^{0} - 4.1702.1014.1702.10^{-1} -
0.025 3.1678.1013.1678.10^{-1} 3.91 8.5482.1028.5482.10^{-2} 3.92 2.9212.1022.9212.10^{-2} 3.84
0.0125 1.8698.1021.8698.10^{-2} 4.08 5.1232.1035.1232.10^{-3} 4.06 1.7850.1031.7850.10^{-3} 4.03
0.00625 1.1427.1031.1427.10^{-3} 4.03 3.1527.1043.1527.10^{-4} 4.02 1.1072.1041.1072.10^{-4} 4.01
0.003125 7.0804.1057.0804.10^{-5} 4.01 1.9599.1051.9599.10^{-5} 4.01 6.9070.1066.9070.10^{-6} 4.00

5.2 Euler equations

In this section we first test our scheme on the following Euler equations in 2D

𝐮t+𝐀1(𝐮)x+𝐀2(𝐮)y=0.\frac{\partial\mathbf{u}}{\partial t}+\dfrac{\partial\mathbf{A}_{1}(\mathbf{u})}{\partial x}+\dfrac{\partial\mathbf{A}_{2}(\mathbf{u})}{\partial y}=0. (25)

where

𝐮=(ρ,ρ𝐯,E) and 𝐀=(ρ𝐯,ρ𝐯𝐯+pId ,(E+p)𝐯)=(𝐀1,𝐀2),\mathbf{u}=(\rho,\rho\mathbf{v},E)\text{ and }\mathbf{A}=(\rho\mathbf{v},\rho\mathbf{v}\otimes\mathbf{v}+p\text{Id },(E+p)\mathbf{v})=(\mathbf{A}_{1},\mathbf{A}_{2}),

We run some standard cases: the isentropic case, the Sod problem and a strong shock tube.

5.2.1 Isentropic vortex

The first considered test case is the isentropic case. The boundary conditions are periodic. The initial conditions are given by

ρ\displaystyle\rho =[1(γ1)β232γπ2exp(1r2)]1γ1,\displaystyle=\left[1-\frac{(\gamma-1)\beta^{2}}{32\gamma\pi^{2}}\exp\big{(}1-r^{2}\big{)}\right]^{\frac{1}{\gamma-1}},
vx\displaystyle v_{x} =1β4πexp(1r22)(yyc),\displaystyle=1-\frac{\beta}{4\pi}\exp\left(\frac{1-r^{2}}{2}\right)(y-y_{c}),
vy\displaystyle v_{y} =22+β4πexp(1r22)(xxc),\displaystyle=\frac{\sqrt{2}}{2}+\frac{\beta}{4\pi}\exp\left(\frac{1-r^{2}}{2}\right)(x-x_{c}),
p\displaystyle p =ργ,\displaystyle=\rho^{\gamma},

where γ=1.4\gamma=1.4, β=5\beta=5 and r=(xxc)2+(yyc)2r=\sqrt{(x-x_{c})^{2}+(y-y_{c})^{2}}. The computational domain is a square [10,10]×[10,10][-10,10]\times[-10,10]. Also, the free stream conditions are given by:

ρ=1,vx,=1,vy,=32,p=1.\rho_{\infty}=1,\quad v_{x,\infty}=1,\quad v_{y,\infty}=\frac{\sqrt{3}}{2},\quad p_{\infty}=1.

The yy-velocity is chosen such that the particle trajectories never coincide to a mesh point, if one start from a grid point initially. The final time is first set to T=5T=5 for a convergence study (because of the cost mostly). The center of the vortex is set in (xc,yc)=(Tvx,,Tvy,)(x_{c},y_{c})=(Tv_{x,\infty},Tv_{y,\infty}), modulo 2020.

The reference solution is obtained on a regular Cartesian mesh consisting of 100×100100\times 100 elements and 4th order scheme in space and time. The CFL number is set to 1, and we consider the four waves model. In Figs. 3, 4 and 5, we have displayed the pressure, density and norm of the velocity at T=5T=5, respectively. We can observe that the plots show a very good behavior of the numerical scheme. The obtained convergence curves for the three considered error norms are displayed in Fig. 6 and show an excellent correspondence to the theoretically predicted order.

Refer to caption
(a) 4-th order scheme in space and time
Refer to caption
(b) Exact
Figure 3: Plot of the pressure for the vortex problem at T=5T=5.
Refer to caption
(a) 4-th order scheme in space and time
Refer to caption
(b) Exact
Figure 4: Plot of the density for the vortex problem at T=5T=5.
Refer to caption
(a) 4-th order scheme in space and time
Refer to caption
(b) Exact
Figure 5: Plot of the norm of the velocity for the vortex problem at T=5T=5.
Refer to caption
Figure 6: Convergence plot of density for the fourth order scheme in space and time at T=5T=5.

In order to illustrate the long time behavior of the scheme, we show the pressure for T=200T=200 and the error between the computed pressure and the exact one on Fig. 7 and a 200×200200\times 200 grid. Note that the typical time for a vortex to travel across the domain is about 1010.

Refer to caption
(a) pp
Refer to caption
(b) error
Figure 7: Pressure and error between the computed solution and the exact one at T=200T=200 on a 200×200200\times 200 grid. We have pi,jpi,jex[4.2 103,1.6 103]p_{i,j}-p^{ex}_{i,j}\in[-4.2\;10^{-3},1.6\;10^{-3}].
Remark 5.1 (About the stability condition).

In this paper, we have focussed our attention to simulations with CFL=1. However, the stability analysis suggests that higher CFL can be used. In the case of the vortex case, using 5 iteration, we have been able to run this case, up to T=200T=200 with CFL=1.2. This is smaller than what is suggested by table LABEL:stabilityDeC. In this table, only the convection operator is considered, and we are not able to make an analysis where the source term is also included. It seems that the constraints are more severe than those suggested by the linear stability analysis.

5.2.2 Sod test case

Further, we have tested our high order kinetic scheme on a well-known 2D Sod benchmark problem. This test is again solving Euler equation (25). The domain is a square [1,1]×[1,1][-1,1]\times[-1,1]. The initial conditions are given by

(ρ0,vx,0,vy,0,p0)={(1,0,0,1),if r0.5,(0.125,0,0,0.1),otherwise.(\rho_{0},v_{x,0},v_{y,0},p_{0})=\begin{cases}(1,0,0,1),&\text{if }r\leq 0.5,\\ (0.125,0,0,0.1),&\text{otherwise.}\end{cases}

and the boundary conditions are periodic. The final time is T=0.16T=0.16 and the CFL number is set to 1. The two stabilisation methods have been tested and compared. The results for the limitation method are in Fig. 8, while the ones obtained with the MOOD method are displayed on Fig. 9. The two methods provide almost identical results. However, the MOOD method, for this case, never activates the first order scheme, hence the results are obtained with the 4th order scheme. One can observe overshoots and undershoots at the shock, not strong enough to activate the first order scheme. This drawback could be cured if one activates, in the MOOD method, the extrema detection procedure of [25] or [26]. When the limitation method is used, one can observe that the overshoot do not exist any more, while the undershoot are less important but still existing.

Refer to caption
(a) p
Refer to caption
(b) ρ\rho
Refer to caption
(c) 𝐯\|\mathbf{v}\|
Figure 8: Sod problem, T=0.16T=0.16 on a 200×200200\times 200 mesh with the limitation method.
Refer to caption
(a) p
Refer to caption
(b) ρ\rho
Refer to caption
(c) 𝐯\|\mathbf{v}\|
Figure 9: Sod problem, T=0.16T=0.16 on a 200×200200\times 200 mesh with the MOOD method.

5.2.3 Strong shock

The problem is defined on [1.5,1.5]×[1.5,1.5][-1.5,1.5]\times[-1.5,1.5] for T=0.025T=0.025. We had to use the MOOD technique to get the results, the shocks are too strong.

(ρ0,vx,0,vy,0,p0)={(1,0,0,1000) if r0.5(1,0,0,1)else.(\rho_{0},v_{x,0},v_{y,0},p_{0})=\left\{\begin{array}[]{ll}(1,0,0,1000)&\text{ if }r\leq 0.5\\ (1,0,0,1)&\text{else.}\end{array}\right. (26)

The pressure, density and norm of the velocity are displayed in Fig. LABEL:StrongShock, for the final time. The simulation is done with CFL=1=1 on a 200×200200\times 200 grid. On Fig. 10(d), we show the iso-lines of the density (mostly to localize the strong features of the solution) and the elements where the formal accuracy is dropped to first order. These flagged elements are moving in time, and are always localized around the discontinuities of the solution. In most cases, only a very few elements are flagged.

Refer to caption
(a) p
Refer to caption
(b) ρ\rho
Refer to caption
(c) 𝐯\|\mathbf{v}\|
Refer to caption
(d) flag+ρ\rho
Figure 10: Result of case 26 on a 200×200200\times 200 grid, CFL=1CFL=1, space order: 4, time order: 4, MOOD, final time.

6 Conclusion

The purpose of this work was primarily to extend a class of kinetic numerical methods that can run at least at CFL one to the two dimensional case. These methods can handle in a simple manner hyperbolic problems, and in particular compressible fluid mechanics one. Our methodology can be arbitrarily high order and can use CFL number larger or equal to unity on regular Cartesian meshes. We have chosen the minimum number of waves, there are probably better solutions, and this will be the topic of further studies. These methods are not designed only for fluid mechanics, and other type of systems will be explored in the future. One interesting feature of this methods, working for CFL=1, is that the algebra for the streaming part of the algorithm can be made very efficient. This is an interesting feature.

Acknowledgments

F.N.M has been funded by the SNF project 200020_204917 entitled ”Structure preserving and fast methods for hyperbolic systems of conservation laws”.

References

  • [1] R. Abgrall and D. Torlo. Some preliminary results on a high order asymptotic preserving computationally explicit kinetic scheme. Communications in Mathematical Sciences, 20(2):297–326, 2022.
  • [2] S. Jin and Z. Xin. The relaxation schemes for systems of conservation laws in arbitrary space dimensions. Commun. Pure Appl. Math., 48(3):235–276, 1995.
  • [3] R. Natalini. A discrete kinetic approximation of entropy solution to multi-dimensional scalar conservation laws. Journal of differential equations, 148:292–317, 1998.
  • [4] P. Bhatnagar, E. Gross, and M. Krook. A model for collision processes in gases. I. small amplitude processes in charged and neutral one-component systems. Phys. Rev., 94:511–525, 1954.
  • [5] C. Cercignani. The Boltzmann equation and its applications. Springer-Verlag, New York, 1988.
  • [6] F. Bouchut. Construction of BGK models with a family of kinetic entropies for a given system of conservation laws. Journal of Statistical Physics, 95(1/2), 1999.
  • [7] D. Aregba-Driollet and R. Natalini. Discrete kinetic schemes for multidimensional systems of conservation laws. SIAM Journal on Numerical Analysis, 37(6):1973–2004, 2000.
  • [8] P. Csomós and I. Faragó. Error analysis of the numerical solution of split differential equations. Math. Comput. Model., 48(7–8):1090–1106, 2008.
  • [9] H.J. Schroll. High resolution relaxed upwind schemes in gas dynamics. J. Sci. Comput., 17:599–607, 2002.
  • [10] M. Banda and M. Sead. Relaxation weno schemes for multi-dimensional hyperbolic systems of conservation laws. Numer. Methods Partial. Differ. Equ., 23(5):1211–1234, 2007.
  • [11] P. Lafitte, W. Melis, and G. Samaey. A high-order relaxation method with projective integration for solving nonlinear systems of hyperbolic conservation laws. J. Comput. Phys., 340:1–25, 2017.
  • [12] D. Coulette, E. Franck, P. Helluy, M. Mehrenberger, and L. Navoret. High-order implicit palindromic discontinuous galerkin method for kinetic-relaxation approximation. Comput. Fluids, 190:485–502, 2019.
  • [13] S. Jin. Efficient asymptotic-preserving (ap) schemes for some multiscale kinetic equations. SIAM J. Sci. Comput., 21:441–454, 1999.
  • [14] S. Boscarino, L. Pareschi, and G. Russo. Implicit–explicit runge–kutta schemes for hyperbolic systems and kinetic equations in the diffusion limit. SIAM J. Sci. Comput., 35(1):22–51, 2013.
  • [15] S. Boscarino and G. Russo. On a class of uniformly accurate imex runge–kutta schemes and applications to hyperbolic systems with relaxation. SIAM J. Sci. Comput., 31(3):1926–1945, 2009.
  • [16] G. Dimarco and L. Pareschi. Asymptotic-preserving implicit–explicit runge–kutta methods for nonlinear kinetic equations. SIAM J. Numer. Anal., 51(2):1064–1087, 2013.
  • [17] F. Filbet and S. Jin. A class of asymptotic-preserving schemes for kinetic equations and related problems with stiff sources. J. Comput. Phys., 229(20):7625–7648, 2010.
  • [18] R. Abgrall. High order schemes for hyperbolic problems using globally continuous approximation and avoiding mass matrices. J. Sci. Comput., 73(2):461–494, 2017.
  • [19] Rémi Abgrall. Some remarks about conservation for residual distribution schemes. Comput. Methods Appl. Math., 18(3):327–351, 2018.
  • [20] E. Hairer and G. Wanner. Solving ordinary differential equations II. stiff and differential-algebraic problems. Springer Series in Comput. Math., Springer-Verlag, Berlin, 14, 2010.
  • [21] A. Iserles. Order stars and saturation theorem for first-order hyperbolics. IMA J. Numer. Anal., 2:49–61, 1982.
  • [22] P. K. Sweby. High resolution schemes using flux limiters for hyperbolic conservation laws. SIAM J. Numer. Anal., 21:995–1011, 1984.
  • [23] Randall J. LeVeque. Wave propagation algorithms for multidimensional hyperbolic systems. J. Comput. Phys., 131(2):327–353, 1997.
  • [24] H. C. Yee, R. F. Warming, and Ami Harten. On a class of TVD schemes for gas dynamic calculations. Numerical methods for the Euler equations of fluid dynamics, Proc. INRIA Workshop, Rocquencourt/France 1983, 84-107 (1985)., 1985.
  • [25] S. Diot, R. Loubère, and S. Clain. The multidimensional optimal order detection method in the three-dimensional case: very high-order finite volume method for hyperbolic systems. Int. J. Numer. Methods Fluids, 73(4):362–392, 2013.
  • [26] François Vilar. A posteriori correction of high-order discontinuous Galerkin scheme through subcell finite volume formulation and flux reconstruction. J. Comput. Phys., 387:245–279, 2019.