Implicit-Explicit schemes for BGK kinetic equations ∗

21
Implicit-Explicit schemes for BGK kinetic equations * Sandra Pieraccini, Gabriella Puppo Abstract In this work a new class of numerical methods for the BGK model of kinetic equations is presented. In principle, schemes of any order of accuracy in both space and time can be constructed with this technique. The methods proposed are based on an explicit-implicit time discretization. In particular the convective terms are treated explicitly, while the source terms are implicit. In this fashion even problems with infinite stiffness can be integrated with relatively large time steps. The conservation properties of the schemes are investigated. Numerical results are shown for schemes of order 1, 2 and 5 in space, and up to third order accurate in time. 1 Introduction Navier Stokes equations describe the motion of a fluid when the continuum hypothesis underlying the model is valid: in particular, the ratio between the mean free path (λ) and the characteristic dimensions of the problem (L) is small: Kn = λ/L 1, where Kn is the Knudsen number. When Kn = O(1), the kinetic theory of rarefied gas dynamics comes into play. Here the fundamental equation is Boltzmann’s equation. Traditionally, an important field of application of kinetic theory has been the motion of objects in the rarefied layers of the atmosphere, such as re-entry problems in aerospace engineering. In these cases in fact, the Knudsen number is large, because the mean free path is of orders of magnitude larger than the characteristic length of the space vehicle. Recently however a huge new field of applications of Boltzmann’s equations has begun to develop in the modelling of fluid flows in nanostructures: in this case, the Knudsen number is large because the scale of interest L is so small that the ratio λ/L is of order one and microscopic effects cannot be neglected. The main tool to integrate Boltzmann’s equation numerically is the Direct Simulation Monte Carlo (DSMC) method, see for instance the classical reference [7]; a more recent review can be found in [21], see also references therein. The DSMC scheme however is very slow due to its low convergence rate and to the complexity of the evaluation of the collision term. Moreover its results are polluted by stochastic noise and therefore lack smoothness. Interest has also focused on deterministic methods, see for instance [20]. From a numerical point of view, the BGK [6, 8] model approximating Boltzmann’s equation for moderate Knudsen numbers is particularly attractive. It has a strong theoretical background, see for instance [23], and extensive numerical computations have tested its ability to approximate Boltzmann solutions for moderate Kn and Euler solutions for Kn 1, see [10] and [27]. More recently, extensive research on the elliptic BGK-ES model [13] has drawn attention to the improved approximation of the Navier Stokes regime for low Knudsen numbers, [27, 2, 16]. The BGK model has also been used to evaluate several flows of physical interest, see for instance [3, 4] and [1, 17] for applications to reacting * This work was supported by INDAM project “Kinetic Innovative Models for the Study of the Behaviour of Fluids in Micro/Nano Electromechanical Systems”, and by PRIN04-Brezzi project Dipartimento di Matematica, Politecnico di Torino, corso Duca degli Abruzzi 24, 10129 Torino, Italia, e-mail: [email protected], [email protected] 1

Transcript of Implicit-Explicit schemes for BGK kinetic equations ∗

Page 1: Implicit-Explicit schemes for BGK kinetic equations ∗

Implicit-Explicit schemes for BGK kinetic equations ∗

Sandra Pieraccini, Gabriella Puppo †

Abstract

In this work a new class of numerical methods for the BGK model of kinetic equations is presented.

In principle, schemes of any order of accuracy in both space and time can be constructed with this

technique. The methods proposed are based on an explicit-implicit time discretization. In particular

the convective terms are treated explicitly, while the source terms are implicit. In this fashion even

problems with infinite stiffness can be integrated with relatively large time steps. The conservation

properties of the schemes are investigated. Numerical results are shown for schemes of order 1, 2 and

5 in space, and up to third order accurate in time.

1 Introduction

Navier Stokes equations describe the motion of a fluid when the continuum hypothesis underlying themodel is valid: in particular, the ratio between the mean free path (λ) and the characteristic dimensionsof the problem (L) is small: Kn = λ/L � 1, where Kn is the Knudsen number.

When Kn = O(1), the kinetic theory of rarefied gas dynamics comes into play. Here the fundamentalequation is Boltzmann’s equation. Traditionally, an important field of application of kinetic theory hasbeen the motion of objects in the rarefied layers of the atmosphere, such as re-entry problems in aerospaceengineering. In these cases in fact, the Knudsen number is large, because the mean free path is of ordersof magnitude larger than the characteristic length of the space vehicle. Recently however a huge newfield of applications of Boltzmann’s equations has begun to develop in the modelling of fluid flows innanostructures: in this case, the Knudsen number is large because the scale of interest L is so small thatthe ratio λ/L is of order one and microscopic effects cannot be neglected.

The main tool to integrate Boltzmann’s equation numerically is the Direct Simulation Monte Carlo(DSMC) method, see for instance the classical reference [7]; a more recent review can be found in [21],see also references therein. The DSMC scheme however is very slow due to its low convergence rate andto the complexity of the evaluation of the collision term. Moreover its results are polluted by stochasticnoise and therefore lack smoothness. Interest has also focused on deterministic methods, see for instance[20].

From a numerical point of view, the BGK [6, 8] model approximating Boltzmann’s equation formoderate Knudsen numbers is particularly attractive. It has a strong theoretical background, see forinstance [23], and extensive numerical computations have tested its ability to approximate Boltzmannsolutions for moderate Kn and Euler solutions for Kn � 1, see [10] and [27]. More recently, extensiveresearch on the elliptic BGK-ES model [13] has drawn attention to the improved approximation of theNavier Stokes regime for low Knudsen numbers, [27, 2, 16]. The BGK model has also been used toevaluate several flows of physical interest, see for instance [3, 4] and [1, 17] for applications to reacting

∗This work was supported by INDAM project “Kinetic Innovative Models for the Study of the Behaviour of Fluids inMicro/Nano Electromechanical Systems”, and by PRIN04-Brezzi project

†Dipartimento di Matematica, Politecnico di Torino, corso Duca degli Abruzzi 24, 10129 Torino, Italia, e-mail:[email protected], [email protected]

1

Page 2: Implicit-Explicit schemes for BGK kinetic equations ∗

gas mixtures. We also mention an application of BGK-like ideas to the development of models for thebehaviour of fluids in nanostructures [11].

The importance of the BGK model in applications has prompted a parallel development of numericalmethods tailored to the particular structure of BGK equations. We start mentioning the first ordernumerical scheme proposed in [10]. A second order scheme is described in [4]. In this case, the schemeis constructed with linear second order upwinding: the lack of limiters results in the possible onset ofspurious oscillations. The third order in space scheme appearing in [27] exploits ideas derived fromhigh order schemes for conservation laws: here the onset of spurious oscillations is prevented using ENO(Essentially Non Oscillatory) reconstructions from [12]. The scheme relies on a first order operatorsplitting, and thus its accuracy in time does not match the high accuracy of the space discretization.

The high order schemes mentioned so far do not satisfy exact conservation of the macroscopic variables(mass, momentum and energy) at the discrete level. This problem is solved in [15] and [16]. Exactconservation is obtained computing equilibrium at the discrete level. This construction requires thesolution of a non linear system of 5 equations for the BGK model and 10 equations for the ES-BGKmodel at each grid point in space, even for explicit integration in time.

All schemes described so far require the solution of large linear systems of equations when the BGKmodel becomes stiff, i.e. for small relaxation times. Moreover the heavy non linearity of the collisionterm usually requires further approximations in the evaluation of the Jacobian matrices needed to solvethe non linear system arising from the discretization of the BGK equations in the implicit case.

In this work we propose a new high order scheme in both space and time to solve the BGK equations.Our method is based on a high order implicit-explicit discretization of the time derivative obtained withRunge-Kutta IMEX schemes, following the work of [14], [5] and [22]. In our scheme, the discretization ofthe time derivative is carried out before the approximation of the space derivatives and the discretizationof velocity space, as in [18]. In particular, the convective term is treated explicitly, while the source term isintegrated implicitly. This approach allows to treat even problems with infinite stiffness easily, see [22] and[24] for applications to relaxation systems. In the implicit step, we first evaluate the macroscopic momentsand next we update the distribution function: in this fashion, conservation is naturally enforced, and thesolution of the implicit step is local, with no need to solve large systems of equations. We think thatthis approach yields a noticeable improvement in the efficiency of the scheme, and it allows to achievevery high order accuracy. Space discretization is carried out with WENO (Weighted Essentially NonOscillatory) reconstructions, see [25]. When the grid in velocity space is introduced, exact conservationno longer holds at the discrete level. However, we find in our tests that the error in the conserved variablesis very small. Exact conservation can be enforced following the technique introduced in [15].

In this work, we show results obtained for schemes of order 1, 2, and 5 in space and 1, 2, and 3 in time,respectively. Higher order schemes can be derived with the same techniques, provided suitable higherorder IMEX Runge-Kutta schemes become available.

We start reviewing the BGK model in §2, and we describe the main characteristics of the new numericalscheme in §3. Next we briefly describe the space and time discretizations in §4: this material is wellestablished, but the application of these ideas in the context of kinetic equations is not standard and itallows some simplifications. We end with a few numerical tests in §5.

2 The BGK model

In this section the BGK model is described. For simplicity, we only consider the classical BGK modelintroduced in [6], using the notation of [10] and [15]. The scheme can be easily extended to more generalBGK models.

2

Page 3: Implicit-Explicit schemes for BGK kinetic equations ∗

We consider the initial value problem:

∂f

∂t(x, v, t) + v · ∇xf(x, v, t) =

1

τ(fM (x, v, t) − f(x, v, t)) t ≥ 0, x ∈ R

d, v ∈ RN (1)

f(x, v, 0) = f0(x, v) ≥ 0 given initial data.

Generally d = 3 and N = 3. In the 1D case, d = 1 and N = 3 and ∇x = (∂x1, 0, 0). In (1) fM is the

Maxwellian obtained from the moments of f , namely:

fM (x, v, t) =ρ(x, t)

(2πRT (x, t))N/2exp

(

−‖v − u(x, t)‖2

2RT (x, t)

)

The quantities ρ, u and T are respectively the macroscopic density, velocity and temperature of thegas, and they are obtained from the moments of f , which are defined as follows. Given any functiong : R

N 7→ R, let us denote, as in [15], by 〈g〉 the quantity∫

RN g(v) dv; if g : RN 7→ R

p, p > 1, we stilldenote by 〈g〉 ∈ R

p the vector whose components are given by 〈gi〉. The moments of f are defined by

ρmE

=

f

1v

12‖v‖2

. (2)

(we dropped the dependence from x and t for simplicity). Here m is momentum, so that the macroscopicvelocity is simply u = m/ρ, while E is the total energy, and the temperature is obtained from theinternal energy e, through the relations: ρe = E− 1

2ρu2, e = NRT/2. In most applications, N = 3 whichcorresponds to a monoatomic gas with three translational degrees of freedom. In our tests, for simplicity,we will choose instead N = 1, as in [10]. This corresponds to a gas with a single degree of freedom, sothat e = RT/2. The only difference with respect to a physical monoatomic gas appears in this rescalingof the temperature. With this choice, all velocity integrals will be evaluated in R instead of R

3. Anotherapproach to reduce the computational complexity of the velocity integrals, while maintaining the physicalproperties of the gas, has been introduced in [9] and used, among others, in [4] and [27].

In [4] the relaxation time τ is defined by

τ−1 = Acρ,

where Ac is a constant, in such a way that Acρ represents the collision frequency of the gas molecules.In [15] the BGK model is written with a relaxation time given by

τ−1 = CρT 1−ω

where ω is the exponent of the viscosity law of the gas (for example, for argon one has ω = 0.81). In theadimensional case one has

τ−1 =C

Kn.

As in [10], we will pick this last choice with C = 1.The macroscopic moments of f are conserved, in the sense that:

∂t 〈f〉 + ∇x · 〈fv〉 = 0, (3a)

∂t 〈fv〉 + ∇x · 〈v ⊗ vf〉 = 0, (3b)

∂t

1

2‖v‖2f

+ ∇x ·⟨

1

2‖v‖2vf

= 0. (3c)

Moreover, an entropy principle holds, namely:

∂t 〈f log f〉 + ∇x 〈vf log f〉 ≤ 0, ∀f ≥ 0, (4)

3

Page 4: Implicit-Explicit schemes for BGK kinetic equations ∗

the equality holding only for f = fM .A numerical scheme for (1) should be able not only to yield an accurate solution to equation (1), but

also to satisfy the conservation equations and the entropy principle in some discretized form.

3 The numerical scheme

Following [18] we start discretizing the BGK problem (1) in time with a Runge-Kutta IMEX scheme(for the notation, see §4). Next the scheme will be discretized in space and, lastly, we will consider thediscretization of velocity space and the evaluation of velocity integrals through quadrature. Finally, wewill review the simplified algorithm that will be tested in the numerical problems.

Time discretization

For the sake of simplicity we consider a unidimensional problem in space, and we introduce a uniformgrid in time, with spacing ∆t. Let fn(x, v) = f(x, v, n∆t). The updated density distribution is given by:

fn+1(x, v) = fn(x, v) − ∆t

ν∑

i=1

wiv ∂xf (i)|(x,v) + ∆t

ν∑

i=1

wi

τ (i)(f

(i)M (x, v) − f (i)(x, v)). (5)

where f(i)M (x, v), i = 1, ..., ν, are the Maxwellian functions obtained from the moments of the intermediate

stages f (i)(x, v), which are defined by the following relations (we drop the dependence on (x, v) forsimplicity):

f (1) = fn + ∆ta11

τ (1)(f

(1)M − f (1)), (6a)

f (i) = fn − ∆t

i−1∑

l=1

ailv ∂xf (l) + ∆t

i∑

l=1

ail

τ (l)(f

(l)M − f (l)). (6b)

The coefficients ail, ail, wi and wi define the IMEX Runge-Kutta scheme and are specified in §4.Let φ(v) denote the vector of the collision invariants of f , φ(v) = (1, v, 1

2‖v‖2)T . First we computethe moments of f . For the first stage value we have:

f (1)φ⟩

= 〈fnφ〉 + ∆ta11

τ (1)

⟨(

f(1)M − f (1)

)

φ⟩

.

Since⟨(

f(1)M − f (1)

)

φ⟩

= 0,

we immediately find the macroscopic variables ρ(1), u(1) and T (1) corresponding to f (1). With thesequantities, we compute the corresponding Maxwellian:

f(1)M =

ρ(1)

(2πRT (1))N/2exp

(

−||v − u(1)||22RT (1)

)

(7)

and once the Maxwellian is known ∀x, the stage value f (1) can simply be found solving the linear equation(6a), for each x, v. For the following stage values we note that, iterating the construction, for all i we

have⟨(

f(i)M − f (i)

)

φ⟩

= 0. Thus the moments at the i-th stage are given by:

f (i)φ⟩

= 〈fnφ〉 − ∆ti−1∑

l=1

ail

v ∂xf (l)φ⟩

4

Page 5: Implicit-Explicit schemes for BGK kinetic equations ∗

and can be explicitly computed. Then the Maxwellian f(i)M can be computed as in (7). Now let:

B(i) = fn − ∆ti−1∑

l=1

ailv∂xf (l) + ∆ti−1∑

l=1

ail

τ (l)(f

(l)M − f (l)),

that is B(i) contains all information coming from the previously computed stages. Finally we obtain

f (i) =τ (i)B(i) + ∆t aiif

(i)M

τ (i) + ∆t aii.

Note that if τ (i) � 1, f (i) immediately relaxes on the local Maxwellian f(i)M .

This completes the description of the time discretization of our scheme. We end this part formallystating the conservation property of the time discretization.

Proposition 1 The time discretization (5)-(6) preserves conservation of density, momentum and total

energy.

Proof. Computing the moments in (5) one has

fn+1φ⟩

= 〈fnφ〉 − ∆t∂x

v

ν∑

i=1

wif(i)φ

.

Since this is an explicit Runge-Kutta discretization of (3), it is straightforward that density, momentumand energy are conserved by the time discretization. �

Space discretization

We introduce a uniform grid in space, with grid points xj , where the distribution function f will beevaluated: fn

j (v) = f(xj , v, n∆t). The grid spacing will be denoted by ∆x = xj+1 − xj . The maintask of the space discretization is to provide an accurate approximation of ∂xf , satisfying the followingrequirements:

• the discretization of ∂xf should be non-oscillatory to prevent the onset of spurious oscillations,characteristic of high order schemes in the presence of discontinuities or sharp gradients in thesolution;

• the discretization should be conservative.

There are two main strategies to approximate the space derivative appearing in (5) in a conservativeform. The most straightforward (finite volume formulation) is to integrate equation (5) on each cellIj = (xj − ∆x/2, xj + ∆x/2), and to update in time the cell averages of f :

fn

j (v) =1

∆x

Ij

fn(x, v) dx.

The advantage of this approach is that the cell average of the space derivative is simply:

v∂xfn

j =1

∆xv

(

f(xj +∆x

2) − f(xj −

∆x

2)

)

.

The disadvantage of this approach is that it requires to compute cell averages of the Maxwellian fM . Forfirst and second order accurate schemes, this is easily accomplished, because:

fM j(v) =1

∆x

Ij

fM (x, v) dx = fM (xj , v) + O(∆x)2.

5

Page 6: Implicit-Explicit schemes for BGK kinetic equations ∗

For schemes of higher order accuracy however, a more accurate quadrature rule in space is needed, whichrequires to evaluate fM at quadrature nodes beyond the grid points xj . This need couples the space cellstogether, and the scheme becomes much more complex.

The second approach is to evaluate (5) at the grid points, and approximate the derivative of f witha conservative finite difference formula, as in [26]. This is the approach we will follow in this work. Notethat in this case, to preserve high accuracy the grid spacing ∆x must be uniform or at least smoothlyvarying in x.

Suppose that the values of a flux function F (f) are given at the grid points xj . In the present case,

clearly the flux function is linear, with F (f) = v f . The first step is to look for a function F thatinterpolates the data Fj = F (f(., xj , .)) in the sense of cell averages:

Fj =1

∆x

Ij

F dx,

thus the derivative of F will be written as:

∂xF |xj=

1

∆x

(

F (xj+1/2) − F (xj−1/2))

,

where xj±1/2 = xj±∆x/2. We outline the main steps of the construction: more details will be found in §4and [25]. The approximation to F is typically a piecewise polynomial function, with jump discontinuitiesat the cell borders xj±1/2. To ensure stability, it is necessary to pick information coming from the correctdirection. In other words, it is necessary to introduce upwinding. In this work, we will use flux splitting,which is particularly straightforward for the convective term of the BGK model.

Thus we write the flux F as the sum of its positive and negative parts: F = F + + F− where F+

and F− have only non negative (respectively: non positive) eigenvalues. Clearly, in our case, the fluxsplitting will depend on the sign of v, namely:

F+ =

{

v f if v > 00 otherwise

F− =

{

0 if v ≤ 0v f otherwise

Next, two reconstructions are computed, one for F + and one for F−, which will be called respectivelyF+ and F−. Each of these has a jump discontinuity at xj+1/2: to use upwinding, we pick the value from

the left for the positive flux, F+j+1/2 = F+(x−

j+1/2), and we pick the value from the right for the negative

flux: F−

j+1/2 = F−(x+j+1/2). Thus the numerical flux at each cell border will be given by:

Fj+1/2 = F+(x−

j+1/2) + F−(x+j+1/2), (8)

and the conservative approximation to the space derivative will be given by:

∂x(v f)|j =1

∆x

(

Fj+1/2 − Fj−1/2

)

. (9)

In the present case, the structure of F is particularly simple, and one has:

Fj+1/2(v) = max(v, 0)f(x−

j+1/2) + min(v, 0)f(x+j+1/2), (10)

where f is a piecewise polynomial function such that fj = 1∆x

Ijf dx.

For example, for the first order scheme f is piecewise constant, namely f(x)|Ij≡ fj . Thus f(x−

j+1/2) =

fj while f(x+j+1/2) = fj+1 and the numerical flux in this case will be given by:

Fj+1/2(v) = max(v, 0)fj + min(v, 0)fj+1. (11)

6

Page 7: Implicit-Explicit schemes for BGK kinetic equations ∗

Substituting the space discretization in (5), we find:

fn+1j (v) = fn

j (v) − ∆t

∆x

ν∑

i=1

wi

(

F(i)j+1/2(v) − F

(i)j−1/2(v)

)

+ ∆t

ν∑

i=1

wi

τ(i)j

(f(i)M,j(v) − f

(i)j (v)), (12)

with a similar modification of (6) for the computation of the stage values f (i) at the grid points xj :

f(1)j (v) = fn

j (v) + ∆ta11

τ(1)j

(f(1)M,j(v) − f

(1)j (v)), (13a)

f(i)j (v) = fn

j (v) − ∆t

∆x

i−1∑

l=1

ail

(

F(l)j+1/2(v) − F

(l)j−1/2(v)

)

+ ∆ti∑

l=1

ail

τ(l)j

(f(l)M,j(v) − f

(l)j (v)). (13b)

We end stating the conservation properties of the semi-discrete scheme just described.

Proposition 2 The semi-discrete scheme (12)-(13) preserves conservation of density, momentum and

total energy.

Proof. Computing the moments of (12) and (13) in v space, we obtain a conservative and consistentdiscretization of the conservation equations (3) for ρ, m and E. In fact, the space discretization is carriedout with an algorithm which is conservative by construction. Further, the numerical flux functions givenin (11) and in the next §4 ensure consistency. By the Lax-Wendroff theorem, if the numerical solutionconverges, we conclude that the macroscopic quantities are conserved. �

Velocity discretization

The discretization of the velocity arises several problems. The main difficulty is that the moments of fare computed approximately through a quadrature formula. As a consequence, it is no longer true thatf and fM have the same moments.

A second problem is due to the difficulty of selecting a suitable grid in velocity space. On one hand,a fine grid ensures a good accuracy of the velocity integrals which yield the macroscopic variables. Onthe other hand, f must be updated on each velocity grid point, which means that the scheme increasesrapidly in computational cost as the velocity grid is refined. Moreover, the solution is quite sensitive tothe choice of velocity grid points, which seems to be highly problem dependent, see for instance [4], wherethe velocity grid changes at almost each test problem.

Choose a grid in velocity space, and let {vk}, k ∈ K, be the set of the grid points, where k is amulti-index if N > 1. Given any function g : R

N 7→ R, let:

〈g〉K = Q(g, {vk}) '∫

RN

g(v) dv,

i.e. 〈g〉K denotes the approximation of 〈g〉 obtained by means of a suitable quadrature rule Q( · , {vk})built on the nodes {vk}; for a vector function g : R

N 7→ Rp we use a convention similar to the one already

introduced for 〈g〉, i.e. 〈g〉K ∈ Rp and the application of 〈 · 〉K is meant componentwise.

The macroscopic variables now depend on the quadrature rule and on the grid used in velocity space.Let:

ρK

mK

EK

= 〈fφ〉K

7

Page 8: Implicit-Explicit schemes for BGK kinetic equations ∗

be the moments computed from f with the quadrature Q( · , {vk}). Now construct an approximateMaxwellian with the formula:

MK(f)(x, v, t) =ρK(x, t)

(2πRTK(x, t))N/2exp

(

−||v − uK(x, t)||22RTK(x, t)

)

(14)

The problem is that there is no reason why f and MK(f) should have the same discrete moments, thatis in general:

〈(f − MK(f)) φ〉K 6= 0,

as pointed out in [15]. However, again in [15], it is proven that it is possible to find a discrete MaxwellianM such that:

〈(f −M(f)) φ〉K = 0, M(f)(x, v, t) = exp (α(x, t) · φ(v)) , (15)

where α(x, t) is an unknown vector that depends on the macroscopic quantities. Note that α is computedprecisely solving the non linear system defined by (15).

We can now introduce the discretization in velocity space in the semidiscrete system defined by (12)and (13). Let fn

jk = f(xj , vk, tn). The updated values of f will be given by:

fn+1jk = fn

jk − ∆t

∆x

ν∑

i=1

wi

(

F(i)j+1/2(vk) − F

(i)j−1/2(vk)

)

+ ∆t

ν∑

i=1

wi

τ(i)K,j

(M(f(i)j )(vk) − f

(i)jk ). (16)

To compute the stage values, at each level i, for each space grid point xj , first compute the discrete

moments (ρK , mK , EK)j . Then solve equation (15) to find the local components of the vector α(i)j , and

thus compute the discrete Maxwellian M(f(i)j ). Then the distribution function can easily be evaluated

at each space and velocity grid point:

f(1)jk = fn

jk + ∆ta11

τ(1)K,j

(M(f(1)j )(vk) − f

(1)jk ) (17a)

f(i)jk = fn

jk − ∆t

∆x

i−1∑

l=1

ail

(

F(l)j+1/2(vk) − F

(l)j−1/2(vk)

)

+ ∆t

i∑

l=1

ail

τ(l)K,j

(M(f(l)j )(vk) − f

(l)jk ). (17b)

Computing the discrete moments of (16), we find a consistent and conservative discretization of theconservation equations (3). Thus again density, momentum and total energy are conserved.

It is worthwhile noting that the discrete Maxwellian M(f) is always computed explicitly, even if themethod is implicit in the collision term. This is simpler than in the conservative scheme in [15], wherethe implicit version of the scheme requires an implicit evaluation of the discrete Maxwellian.

Now we describe a simpler version of the scheme defined by (16) and (17), in which the velocitydiscretization is only approximately conservative.

We start from the semidiscrete equations (12) and (13). The updated values of f are now given by:

fn+1jk = fn

jk − ∆t

∆x

ν∑

i=1

wi

(

F(i)j+1/2(vk) − F

(i)j−1/2(vk)

)

+ ∆t

ν∑

i=1

wi

τ(i)K,j

(f(i)M,jk − f

(i)jk ). (18)

where f(i)M,jk is the usual Maxwellian computed from the moments evaluated at the i-th stage in the

following way:

f(i)j φ

K=⟨

fnj φ⟩

K− ∆t

∆x

i−1∑

l=1

ail

⟨(

F(l)j+1/2 − F

(l)j−1/2

)

φ⟩

K+ ∆t

i∑

l=1

ail

τ(l)K,j

⟨(

M(f(l)j ) − f

(l)j

)

φ⟩

K. (19)

8

Page 9: Implicit-Explicit schemes for BGK kinetic equations ∗

Note that since in the equation above the discrete Maxwellian appears, the last term drops out. From

the discrete moments, we compute the corresponding Maxwellian f(i)M,jk. Then the stage values are given

by:

f(1)jk = fn

jk + ∆ta11

τ(1)K,j

(f(1)M,jk − f

(1)jk ), (20a)

f(i)jk = fn

jk − ∆t

∆x

i−1∑

l=1

ail

(

F(l)j+1/2(vk) − F

(l)j−1/2(vk)

)

+ ∆t

i∑

l=1

ail

τ(l)K,j

(f(l)M,jk − f

(l)jk ). (20b)

This scheme requires no solution of the non linear system (15), and is therefore much faster than theconservative scheme described above. In fact, the discrete Maxwellian M(f), although used, is nevercomputed. We wish to stress the fact that this simplified scheme is not exactly conservative. However, inthe numerical tests reported in §5 we will see that the errors on the conserved variables are indeed verysmall.

We end this section with a summary of the simplified scheme defined by (18) and (20). We use a νstages IMEX scheme among those proposed in [19, 22]. Let Nx be such that Nx + 1 is the number ofpoints in the space grid and let Nv be the number of nodes in velocity space. The overall scheme readsas follows:

• Given f0 := f(xj , vk, 0) ∈ RNv×Nx+1

• For n = 0, 1, ...

– For i = 1, ..., ν

∗ Compute the moments m(i) and the approximate Maxwellian f(i)M corresponding to the

point values (xj , vk), j = 0, ..., Nx, k = 1, ..., Nv, with i = 1 corresponding to the startingdata at tn.

∗ If i > 1 construct the numerical fluxes F(i−1)j+1/2(vk), k = 1, ..., Nv, with the reconstruction

procedure described in section 4, and according to equation (8).

∗ Assemble the Right Hand Side of (20) and solve for f (i).

– Compute fn+1 via (18).

4 Time and Space discretizations

In this section, we give a few details on IMEX Runge-Kutta schemes and on high order conservative finitedifference formulas for the evaluation of the space derivatives appearing in (5) and (6). The algorithmswe will overview in this section are drawn from [22] for the IMEX part and from [25] for the WENOdifferentiation formula.

4.1 IMEX Runge-Kutta schemes

To set the notation for the Runge-Kutta IMEX schemes, we consider the autonomous ODE problem:

y′(t) = f(y) + 1τ g(y)

y(t0) = y0.(21)

We suppose that 0 ≤ τ << 1, i.e. g/τ is stiff, so that we wish to integrate it implicitly, while f is nonstiff, but highly non linear, which means that for f an explicit scheme is more efficient.

9

Page 10: Implicit-Explicit schemes for BGK kinetic equations ∗

Let A = (ais) and A = (ais) be two ν × ν matrices, with ais = 0 for s ≥ i and ais = 0 for s > i,and let w, w be two coefficient vectors with ν elements. An IMEX Runge-Kutta scheme for autonomousproblems is represented by the following double Butcher’s tableau:

A

wT

A

wT

The corresponding numerical scheme for (21) is:

yn+1 = yn + ∆tν∑

i=1

wif(y(i)) +∆t

τ

ν∑

i=1

wig(y(i)), (22)

where the stage values y(i) are given by:

y(1) = yn +∆t

τa11 g(y(1)) (23)

y(i) = yn + ∆ti−1∑

l=1

ailf(y(l)) +∆t

τ

i∑

l=1

ail g(y(l)). (24)

The coefficients of the Butcher’s tableaux are computed in order to maximize accuracy. Further,the implicit scheme must be L-stable, to ensure that the numerical solution relaxes on the equilibriumsolution, if τ is very small. Moreover, it is desirable that the IMEX scheme becomes a high order explicitnumerical scheme for the conserved variables, when τ → 0: for this reason, the matrix A must beinvertible, see [19].

In the numerical tests, we will apply IMEX schemes of order 1, 2 and 3. Here we add the correspondingButcher’s tableaux.

IMEX1 1st order scheme, [19]

0

1

1

1

IMEX2 2nd order scheme, [22]0 0 00 0 00 1 0

0 12

12

12 0 0

− 12

12 0

0 12

12

0 12

12

IMEX3 3rd order scheme, [22]

0 0 0 00 0 0 00 1 0 00 1

414 0

0 16

16

23

α 0 0 0−α α 0 00 1− α α 0β η 1

2 − β − η − α α

0 16

16

23

being α = 0.24169426078821, β = 0.06042356519705, η = 0.1291528696059.

10

Page 11: Implicit-Explicit schemes for BGK kinetic equations ∗

4.2 Conservative, high order differentiation

The main formula for the computation of the numerical flux is (8). Here we want to report the formulasthat are needed to compute the numerical flux with first, second and fifth order accuracy.

At first order accuracy, the function f (see §3) is reconstructed as a piecewise constant function; for

the second order scheme, f is reconstructed as a piecewise linear function, while fifth order accuracy isachieved reconstructing f as a piecewise parabolic function with WENO coefficients. We have alreadygiven the formula for the numerical flux of the first order scheme in (11).

In the following, we drop the dependence on t and v.

Second order

For a second order flux, the reconstruction for f is piecewise linear:

f(x)|Ij= fj + σj(x − xj),

where σj is a first order accurate limited slope, as for instance σj = MinMod(fj+1 − fj , fj − fj−1)/∆x,or σj = (fj+1 − fj−1)/(2∆x). For the accuracy test in §5 on the smooth case we used this second choice,since the MinMod slope limiter deteriorates accuracy close to extrema. Thus the numerical flux will be:

Fj+1/2(v) = max(v, 0)(fj +1

2σj) + min(v, 0)(fj+1 −

1

2σj+1). (25)

WENO reconstruction

In the WENO reconstruction f(x)|Ijis obtained through the superposition of the three parabolas P l

j (x),with l = −1, 0, 1 interpolating the data fj+l−1, fj+l, fj+l+1 in the sense of cell averages.

We find three estimates of the left and right values of f at xj+1/2, respectively:

f l(x−

j+1/2) = P lj (xj+1/2), f l(x+

j+1/2) = P lj+1(xj+1/2), l = −1, 0, 1.

The estimates are easily obtained as

f−1(x−

j+1/2) =1

3fj−2 −

7

6fj−1 +

11

6fj

f0(x−

j+1/2) = −1

6fj−1 +

5

6fj +

1

3fj+1

f1(x−

j+1/2) =1

3fj +

5

6fj+1 −

1

6fj+2

and

f−1(x+j−1/2) = −1

6fj−2 +

5

6fj−1 +

1

3fj

f0(x+j−1/2) =

1

3fj−1 +

5

6fj −

1

6fj+1

f1(x+j−1/2) =

11

6fj −

7

6fj+1 +

1

3fj+2.

Next these data are combined with suitable weights to yield the desired values:

f(x−

j+1/2) =1∑

l=−1

ωlj f

l(x−

j+1/2), f(x+j+1/2) =

1∑

l=−1

ωlj f

l(x+j+1/2). (26)

11

Page 12: Implicit-Explicit schemes for BGK kinetic equations ∗

The weights are defined by the relations:

ωlj =

αlj

∑1s=−1 αs

j

, αlj =

dl

(ε + βlj)

2; ωl

j =αl

j∑1

s=−1 αsj

, αlj =

dl

(ε + βlj)

2,

where

d−1 =1

10, d0 =

3

5, d1 =

3

10; d−1 =

3

10, d0 =

3

5, d1 =

1

10

are the accuracy constants: they are computed in order to obtain a fifth order scheme; ε prevents dividingby zero, and it is usually chosen as ε = 10−6. The quantities βl’s are the Smoothness Indicators: theyprevent the inclusion in the combination of parabolas with a non smooth stencil:

βlj =

k=1,2

Ij

∆x2k−1

(

dkP lj

dxk

)2

dx.

In our case, the Smoothness Indicators are simply given by:

β−1j =

13

12(fj−2 − 2fj−1 + fj)

2 +1

4(fj−2 − 4fj−1 + 3fj)

2

β0j =

13

12(fj−1 − 2fj + fj+1)

2 +1

4(fj−1 − fj+1)

2

β1j =

13

12(fj − 2fj+1 + fj+2)

2 +1

4(3fj − 4fj+1 + fj+2)

2.

Once the values f(x−

j+1/2) and f(x+j+1/2) are computed as in (26), we form the numerical flux as:

Fj+1/2(v) = max(v, 0)f(x−

j+1/2) + min(v, 0)f(x+j+1/2). (27)

5 Numerical experiments

In this section we report some numerical results obtained with the simplified methods defined in (18) and(20). In particular, we will test the first order scheme, BGK1, obtained using the IMEX1 scheme definedin §4 and the first order flux (11). The second order scheme, BGK2, is defined by the second order IMEXscheme IMEX2 of §4, and the second order numerical flux (25). Finally the high resolution scheme BGK3is obtained coupling the third order IMEX3 scheme with the fifth order accurate numerical flux (27).

Following [10], we consider a problem which is 1D in velocity space. We considered the following testproblems.

Test 1 We consider an initial distribution of the kind

f(x, v, 0) =ρ√

2πRT· exp(− (v − u)2

2RT), x ∈ [−1, 1],

with constant density ρ = 1 and temperature T = 1 and with

u =1

σ

(

exp(

− (σx − 1)2)

− exp(

− (σx + 1)2))

with σ = 10. Thus initially the distribution function f is smooth, with a localized perturbation invelocity, in a gas with a uniform density and temperature.

12

Page 13: Implicit-Explicit schemes for BGK kinetic equations ∗

Test 2 We consider as initial data a distribution which is discontinuous in space.

f(x, v, 0) =

{

ρL(2πRTL)−1/2 · exp(− (uL−v)2

2RTL), 0 ≤ x ≤ 0.5,

ρR(2πRTR)−1/2 · exp(− (uR−v)2

2RTR), 0.5 < x ≤ 1

with (ρL, uL, TL) = (2.25, 0, 1.125) and (ρR, uR, TR) = (3/7, 0, 1/6). This test is derived from [10].

We used free-flow boundary conditions in both test problems. We considered several values of theKnudsen number Kn, choosing τ = Kn; further, we assumed R = 1. The velocity space was approximatedby the finite interval [−V, V ] with V = 10; the velocity grid points are uniformly distributed around v = 0.

The time step is chosen following the CFL condition for the explicit, convective part, namely: ∆t =0.9∆x/V . The velocity integrals are approximated by trapezoidal or Simpson rule, or by the followingrule

∫ vi+1

vi

g(v) dv =∆v

24(−g(vi+2) + 13g(vi+1) + 13g(vi) − g(vi−1)) + O(∆v4). (28)

5.1 Test 1

We integrated the first test problem up to time t = 0.04, when the solution is still smooth, in order toestimate the rate of convergence of the numerical solutions. In Tables 1-3 we report the errors computed,in the case τ = 10−5, for different values of ∆x and ∆v with the three methods with respect to a referencesolution obtained with scheme BGK3 on a finer grid, with Nx = 1280 and Nv = 1281. We computedthe errors for the density, mean velocity and temperature. Specifically, we used the quadrature rule (28)to compute the integrals with Nv = Nx + 1. Note that on all grids tested, except the coarsest one, thehigher order scheme yields smaller errors. As the grid is fine enough, the convergence rate accelerates.For the accuracy test reported in these tables, in BGK2 we didn’t use the MinMod slope limiter, in orderto avoid the clipping phenomenon on local extrema.

Density, velocity and temperature obtained with τ = 10−5, Nx = 160 and Nv = 161 are shownin figures 1-3. Further, in figure 4 we plot the total entropy H(f) =

R〈f log(f)〉 dx as a function of

time with τ = 10−2 (left) and τ = 10−5 (right), still with Nx = 160 and Nv = 161. Integrals in spaceare computed through the quadrature rule (28). We recall that for τ decreasing to 0, the BGK modelapproaches the hydrodynamic limit, for which the total entropy remains constant for smooth solutions.It can be seen that in the case τ = 10−2 entropy decays due to kinetic collisional effects, while in the caseτ = 10−5 the solution is very close to the regular solution of the Euler equations and the total entropycomputed with schemes BGK2 and BGK3 is nearly constant.

Finally, in Table 4 we report the errors in conservation for density, momentum and total energy in thecase τ = 10−5. The errors are obtained integrating in space the moments and computing the absolutevalue of the difference between the values at the final time (t = 0.04) and at the initial time. The resultsare obtained with Nx = 160 and Nv = 161; integrals in space are approximated by formula (28). Asexpected, errors in conservation are very small, indeed close to machine precision, even for the simplifiedscheme (18)-(20) which does not require the evaluation of the discrete Maxwellian.

5.2 Test 2

We present the results obtained for density, velocity and temperature for t = 0.16, as in [10], for differentvalues of the parameter τ . We used Nx = 200. Concerning the velocity space, we used the trapezoidalrule with Nv = 81.

Figures 5-7 report the results obtained for the three moments ρ, u, T at the time t = 0.16 with thethree schemes BGK1, BGK2, and BGK3, for τ = 10−2 (left) and τ = 10−5 (right). In figure 8 we plotthe behaviour of the total entropy H(f) versus time, again for the two already considered values of theKnudsen number.

13

Page 14: Implicit-Explicit schemes for BGK kinetic equations ∗

Scheme BGK1 Scheme BGK2 Scheme BGK3L1-errors

Nx error order error order error order20 6.483861E-03 4.551779E-03 3.609352E-0340 3.905098E-03 0.7315 1.071216E-03 2.0872 3.968098E-04 3.185280 2.101849E-03 0.8937 1.826311E-04 2.5522 3.505630E-05 3.5007160 1.084978E-03 0.9540 3.936843E-05 2.2138 1.189247E-06 4.8816320 5.530284E-04 0.9722 9.228221E-06 2.0929 2.521739E-08 5.5595640 2.789262E-04 0.9875 2.238795E-06 2.0433 1.993130E-09 3.6613

L∞-errorsNx error order error order error order20 2.314989E-02 1.170818E-02 1.150409E-0240 1.241993E-02 0.8983 2.331520E-03 2.3282 1.694517E-03 2.763280 6.527985E-03 0.9279 5.026083E-04 2.2138 1.391502E-04 3.6062160 3.337493E-03 0.9679 1.208923E-04 2.0557 5.542816E-06 4.6499320 1.682644E-03 0.9880 2.891851E-05 2.0637 1.211303E-07 5.5160640 8.427201E-04 0.9976 7.087809E-06 2.0286 8.720581E-09 3.7960

Table 1: Errors on ρ

We note that for τ = 10−2 there is a good agreement with the results shown in [10]. Since τ isrelatively large, the profiles are still smooth. Note however that the higher order schemes yield somewhatsharper transitions. The entropy decays with time for all the three schemes tested. The fastest entropydecay is obtained with the first order scheme, since this is of course the most diffusive.

In the case with τ = 10−5, the Knudsen number is so small that the solution has already converged tothe solution of Euler equations for a monoatomic gas with a single degree of freedom. In the figures wecan clearly detect the presence of a contact discontinuity and a shock wave. The results obtained withBGK3 are much sharper than the analogous results shown in [10] with a double number of grid pointsin space, and even the solution obtained with BGK1, which has an order of accuracy comparable withthe scheme used in [10], has a better resolution and does not exhibit spurious oscillations. Again, theentropy decreases in time for all schemes tested.

Acknowledgments. The authors would like to thank Roberto Monaco and Giovanni Russo for severalfruitful discussions and helpful suggestions.

References

[1] P. Andries, K. Aoki, and B. Perthame, A consistent BGK-type model for gas mixtures, Journalof Statistical Physics, 106 (2002), pp. 993–1018.

[2] P. Andries, J. F. Bourgat, P. Le-Tallec, and B. Perthame, Numerical comparison be-

tween the Boltzmann and ES-BGK models for rarefied gases, Computer and Methods in AppliedMathematics and Engineering, 31 (2002), p. 3369.

[3] K. Aoki, K. Kanba, and S. Takata, Numerical analysis of a supersonic rarefied flow past a flat

plate, Physics of Fluids, 9 (1997), pp. 1144–1161.

[4] K. Aoki, Y. Sone, and T. Yamada, Numerical analysis of gas flows condensing on its plane

condensed phase on the basis of kinetic theory, Physics of Fluids A, 2 (1990), pp. 1867–1878.

14

Page 15: Implicit-Explicit schemes for BGK kinetic equations ∗

Scheme BGK1 Scheme BGK2 Scheme BGK3L1-errors

Nx error order error order error order20 4.706440E-03 3.860496E-03 5.308953E-0340 2.908094E-03 0.6946 1.517495E-03 1.3471 9.648877E-04 2.460080 1.724832E-03 0.7536 4.298498E-04 1.8198 6.859511E-05 3.8142160 9.576551E-04 0.8489 1.117550E-04 1.9435 2.646778E-06 4.6958320 5.019068E-04 0.9321 2.820160E-05 1.9865 7.314238E-08 5.1774640 2.571487E-04 0.9648 7.089511E-06 1.9920 6.933381E-09 3.3991

L∞-errorsNx error order error order error order20 1.288922E-02 9.087437E-03 1.188032E-0240 7.820597E-03 0.7208 3.166105E-03 1.5212 2.904380E-03 2.032380 4.572828E-03 0.7742 1.131411E-03 1.4846 2.544219E-04 3.5129160 2.613475E-03 0.8071 2.909468E-04 1.9593 1.051056E-05 4.5973320 1.379249E-03 0.9221 7.584702E-05 1.9396 2.972873E-07 5.1438640 7.111019E-04 0.9558 1.979247E-05 1.9381 2.580405E-08 3.5262

Table 2: Errors on u

[5] U. Asher, S. Ruuth, and R. J. Spiteri, Implicit-explicit Runge-Kutta methods for time dependent

partial differential equations, Appl. Numer. Math., 25 (1997), pp. 151–167.

[6] P. L. Bhatnagar, E. P. Gross, and M. Krook, A model for collision processes in gases. Small

amplitude processes in charged and neutral one-component systems, Physical Reviews, 94 (1954),pp. 511–525.

[7] G. A. Bird, Molecular Gas Dynamics, Oxford University Press, London, 1976.

[8] C. Cercignani, Rarefied Gas Dynamics, from Basic Concepts to Actual Calculations, CambridgeUniversity Press, Cambridge, 2000.

[9] C. K. Chu, Kinetic-theoretic description of the formation of a shock wave, Physics of Fluids, 4(1965), pp. 12–22.

[10] F. Coron and B. Perthame, Numerical passage from kinetic to fluid equations, SIAM J. Numer.Anal., 28 (1991), pp. 26–42.

[11] Z. Guo, T. S. Zhao, and Y. Shi, Simple kinetic model for fluid flows in the nanometer scale,Physical Review E, 71 (2005), pp. 035301–1, 035301–4.

[12] A. Harten, B. Engquist, S. Osher, and S. Chakravarthy, Uniformly high order accurate

essentially non-oscillatory schemes III, Journal of Computational Physics, 71 (1987), pp. 231–303.

[13] L. H. Holway, Kinetic Theory of Shock Structure using an Ellipsoidal Distribution Function, Aca-demic Press, 1966, pp. 193–215.

[14] C. A. Kennedy and M. H. Carpenter, Runge-Kutta schemes for convection-diffusion-reaction

equations, Applied Mathematics and Comp., (2002).

[15] L. Mieussens, Discrete velocity model and implicit scheme for the BGK equation of rarefied gas

dynamics, Mathematical Models and Methods in Applied Sciences, 10 (2000), pp. 1121–1149.

15

Page 16: Implicit-Explicit schemes for BGK kinetic equations ∗

Scheme BGK1 Scheme BGK2 Scheme BGK3L1-errors

Nx error order error order error order20 1.348105E-02 1.012190E-02 8.722357E-0340 9.346559E-03 0.5284 2.834595E-03 1.8363 1.083529E-03 3.009080 5.396658E-03 0.7924 5.109318E-04 2.4719 8.679295E-05 3.6420160 2.894390E-03 0.8988 9.085205E-05 2.4915 2.760939E-06 4.9743320 1.500760E-03 0.9476 1.876646E-05 2.2754 8.040301E-08 5.1018640 7.628044E-04 0.9763 4.268115E-06 2.1365 6.982879E-09 3.5254

L∞-errorsNx error order error order error order20 5.079614E-02 2.643118E-02 2.504089E-0240 3.027057E-02 0.7468 6.780829E-03 1.9627 4.651424E-03 2.428580 1.670170E-02 0.8579 1.388356E-03 2.2881 3.441841E-04 3.7564160 8.732693E-03 0.9355 2.773496E-04 2.3236 1.153828E-05 4.8987320 4.450104E-03 0.9726 6.232958E-05 2.1537 2.920835E-07 5.3039640 2.240976E-03 0.9897 1.475972E-05 2.0783 2.259559E-08 3.6923

Table 3: Errors on T

Scheme BGK1 Scheme BGK2 Scheme BGK3ρ 5.684342E-14 6.306067E-14 4.440892E-15m 2.130964E-16 2.501528E-16 4.555868E-16E 2.220446E-14 3.197442E-14 2.664535E-15

Table 4: Errors in conservation

[16] , Schemes for Boltzmann-BGK equation in plane and axisymmetric geometries, Journal ofComputational Physics, 162 (2000), pp. 429–466.

[17] R. Monaco, M. Pandolfi Bianchi, and A. J. Soares, BGK-type models in strong reaction

and kinetic chemical equilibrium regimes, Tech. Rep. 24, Dipartimento di Matematica, Politecnicodi Torino, 2005. Submitted.

[18] L. Pareschi, G. Puppo, and G. Russo, Central Runge-Kutta schemes for conservation laws,SIAM Journal of Scientific Computing, 26 (2005), pp. 979–999.

[19] L. Pareschi and G. Russo, Implicit-Explicit Runge-Kutta schemes for stiff systems of differential

equations, vol. 3 of Recent Trends in Numerical Analysis, L. Brugnano and D. Trigiante eds., NovaScience, New York, 2000, pp. 269–289.

[20] , Numerical solution of the Boltzmann equation I: Spectrally accurate approximation of the

collision operator, SIAM J. Numerical Analysis, 37 (2000), pp. 1217–1245.

[21] , An introduction to Monte Carlo methods for the Boltzmann equation, in ESAIM: Proceedings,vol. 10, 2001, pp. 35–75.

[22] , Implicit-explicit Runge-Kutta methods and applications to hyperbolic systems with relaxation,J. Scientific Computing, 25 (2005), pp. 129–155.

16

Page 17: Implicit-Explicit schemes for BGK kinetic equations ∗

−0.25 −0.2 −0.15 −0.1 −0.05 0 0.05 0.1 0.15 0.2 0.250.95

0.96

0.97

0.98

0.99

1

1.01

1.02

1.03

BGK1BGK2BGK3reference solution

0.15 0.16 0.17 0.18 0.19 0.2 0.21 0.22 0.23 0.241.01

1.012

1.014

1.016

1.018

1.02

1.022

1.024

1.026

1.028

1.03

BGK1BGK2BGK3reference solution

Figure 1: Test 1, density. Left: scheme BGK1 (dashed line), scheme BGK2 (dash-dot line), schemeBGK3 (continuous line), reference solution (dotted line). Right: detail of the right peak.

[23] B. Perthame, Global existence to the BGK model of Boltzmann equation, Journal of DifferentialEquations, 82 (1989), pp. 191–205.

[24] G. Puppo and G. Russo, Staggered finite difference schemes for balance laws. To appear on:Proceedings of HYP 2004, 2005.

[25] C.-W. Shu, Essentially Non-Oscillatory and Weighted Essentially Non-Oscillatory Schemes for

Hyperbolic Conservation Laws, vol. 1697 of Lecture notes in Mathematics, Springer, Berlin, A.Quarteroni ed., 1998, pp. 325–432.

[26] C.-W. Shu and S. Osher, Efficient Implementation of Essentially Non-oscillatory Shock-Capturing

Schemes, Journal of Computational Physics, 77 (1988), pp. 438–471.

[27] J. Y. Yang and J. C. Huang, Rarefied flow computations using nonlinear model Boltzmann

equations, Journal of Computational Physics, 120 (1995), pp. 323–339.

17

Page 18: Implicit-Explicit schemes for BGK kinetic equations ∗

−0.25 −0.2 −0.15 −0.1 −0.05 0 0.05 0.1 0.15 0.2 0.25−0.06

−0.04

−0.02

0

0.02

0.04

0.06

BGK1BGK2BGK3reference solution

0.12 0.13 0.14 0.15 0.16 0.17 0.18 0.19 0.2 0.21 0.220.04

0.042

0.044

0.046

0.048

0.05

0.052

0.054

0.056

0.058

0.06

BGK1BGK2BGK3reference solution

Figure 2: Test 1, velocity. Left: scheme BGK1 (dashed line), scheme BGK2 (dash-dot line), schemeBGK3 (continuous line), reference solution (dotted line). Right: detail of the right peak.

−0.25 −0.2 −0.15 −0.1 −0.05 0 0.05 0.1 0.15 0.2 0.250.9

0.92

0.94

0.96

0.98

1

1.02

1.04

1.06

BGK1BGK2BGK3reference solution

0.15 0.16 0.17 0.18 0.19 0.2 0.21 0.22 0.23 0.241.035

1.04

1.045

1.05

1.055

1.06

BGK1BGK2BGK3reference solution

Figure 3: Test 1, temperature. Left: scheme BGK1 (dashed line), scheme BGK2 (dash-dot line), schemeBGK3 (continuous line), reference solution (dotted line). Right: detail of the right peak.

18

Page 19: Implicit-Explicit schemes for BGK kinetic equations ∗

0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04−2.8381

−2.838

−2.838

−2.8379

−2.8379

−2.8378

−2.8378

BGK1BGK2BGK3

0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04−2.8381

−2.838

−2.838

−2.8379

−2.8379

−2.8378

−2.8378

BGK1BGK2BGK3

Figure 4: Test 1, entropy; τ = 10−2 (left) and τ = 10−5 (right). Scheme BGK1 (dotted line), schemeBGK2 (dash-dot line), scheme BGK3 (continuous line). In both figures the y axis ranges from −2.8381to −2.8378.

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2

2.2

2.4BGK1BGK2BGK3

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2

2.2

2.4BGK1BGK2BGK3

Figure 5: Test 2, density: τ = 10−2 (left) and τ = 10−5 (right). Scheme BGK1 (dashed line), schemeBGK2 (dash-dot line), scheme BGK3 (continuous line)

19

Page 20: Implicit-Explicit schemes for BGK kinetic equations ∗

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9BGK1BGK2BGK3

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

0

0.5

BGK1BGK2BGK3

Figure 6: Test 2, velocity: τ = 10−2 (left) and τ = 10−5 (right). Scheme BGK1 (dashed line), schemeBGK2 (dash-dot line), scheme BGK3 (continuous line)

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.2

0.4

0.6

0.8

1

1.2BGK1BGK2BGK3

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.2

0.4

0.6

0.8

1

1.2BGK1BGK2BGK3

Figure 7: Test 2, temperature: τ = 10−2 (left) and τ = 10−5 (right). Scheme BGK1 (dashed line),scheme BGK2 (dash-dot line), scheme BGK3 (continuous line)

20

Page 21: Implicit-Explicit schemes for BGK kinetic equations ∗

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16−1.13−1.13

−1.12

−1.11

−1.1

−1.09

−1.08

−1.07

−1.06

−1.05

−1.04BGK1BGK2BGK3

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16−1.1

−1BGK1BGK2BGK3

Figure 8: Test 2, total entropy plotted versus time: τ = 10−2 (left) and τ = 10−5 (right). Scheme BGK1(dashed line), scheme BGK2 (dash-dot line), scheme BGK3 (continuous line)

21