8
ORIGINAL Comparison of different kinetic models for the heat transfer problem Irina A. Graur A. P. Polikarpov Received: 26 June 2009 / Accepted: 30 October 2009 / Published online: 20 November 2009 Ó Springer-Verlag 2009 Abstract The steady-state heat transfer problem between two parallel plates is investigated using kinetic models of the Boltzmann equation: BGK, S-model and ES-model. The discrete velocity method is used to determine the values of physical parameters: density, bulk velocity, temperature and heat flux. The obtained results are compared with the analytical expressions and some experimental data. List of symbols t Time f One particle distribution function r = (x,y,z) Position vector v Molecular velocity u Bulk velocity V = v - u Peculiar velocity v m The most probable molecular speed c Dimensionless molecular velocity n Number density T Temperature q Heat flux vector P ij Stress tensor p Gas pressure m Molecular mass of the gas k Boltzmann constant Pr Prandtl number H Distance between the plates Greek symbols a Accommodation coefficient d Rarefaction parameter DT Temperature difference between the plates s Relaxation time U, W Reduced distribution functions l Shear viscosity Subscript 0 Values on the symmetry axis (y = 0) w Wall quantities FM Free molecular quantities Superscript mod Model distribution function M Maxwellian distribution function ± Parameters of reflected/incident molecules 1 Introduction A systematic research effort in micro mechanics devices started in the late 1980s. Microducts, micronozzles, micropumps, microturbines and microvalves are examples of small devices involving liquid and gas flows. The characteristic length scale of these devices is less than 1 mm but more than 1 micron. Therefore, even at atmo- spheric conditions, the mean-free path to characteristic dimension ratio cannot be neglected and so in the flow dynamic associated with microelectromechanical systems (MEMS) rarefied gas phenomena become apparent. The correct description of the rarefied gas flows for the complete Knudsen number range is based on the Boltzmann kinetic equation. The difficulties associated with its resolu- tion come from the large number of the independent vari- ables (seven in the non stationary case), but also from a very complicated structure of the collision integral. Therefore, in the late sixties, several models of kinetic equations were suggested (BGK, ES, S models), where the collision integral I. A. Graur (&) A. P. Polikarpov Universite ´ de Provence, Ecole Polytechnique Universitaire de Marseille, UMR CNRS 6595, 5 rue Enrico Fermi, 13453 Marseille, France e-mail: [email protected] 123 Heat Mass Transfer (2009) 46:237–244 DOI 10.1007/s00231-009-0558-x

Comparison of different kinetic models for the heat transfer problem

Embed Size (px)

Citation preview

ORIGINAL

Comparison of different kinetic models for the heat transferproblem

Irina A. Graur • A. P. Polikarpov

Received: 26 June 2009 / Accepted: 30 October 2009 / Published online: 20 November 2009

� Springer-Verlag 2009

Abstract The steady-state heat transfer problem between

two parallel plates is investigated using kinetic models of

the Boltzmann equation: BGK, S-model and ES-model. The

discrete velocity method is used to determine the values of

physical parameters: density, bulk velocity, temperature

and heat flux. The obtained results are compared with the

analytical expressions and some experimental data.

List of symbols

t Time

f One particle distribution function

r = (x,y,z) Position vector

v Molecular velocity

u Bulk velocity

V = v - u Peculiar velocity

vm The most probable molecular speed

c Dimensionless molecular velocity

n Number density

T Temperature

q Heat flux vector

Pij Stress tensor

p Gas pressure

m Molecular mass of the gas

k Boltzmann constant

Pr Prandtl number

H Distance between the plates

Greek symbols

a Accommodation coefficient

d Rarefaction parameter

DT Temperature difference between the plates

s Relaxation time

U, W Reduced distribution functions

l Shear viscosity

Subscript

0 Values on the symmetry axis (y = 0)

w Wall quantities

FM Free molecular quantities

Superscript

mod Model distribution function

M Maxwellian distribution function

± Parameters of reflected/incident molecules

1 Introduction

A systematic research effort in micro mechanics devices

started in the late 1980s. Microducts, micronozzles,

micropumps, microturbines and microvalves are examples

of small devices involving liquid and gas flows. The

characteristic length scale of these devices is less than

1 mm but more than 1 micron. Therefore, even at atmo-

spheric conditions, the mean-free path to characteristic

dimension ratio cannot be neglected and so in the flow

dynamic associated with microelectromechanical systems

(MEMS) rarefied gas phenomena become apparent.

The correct description of the rarefied gas flows for the

complete Knudsen number range is based on the Boltzmann

kinetic equation. The difficulties associated with its resolu-

tion come from the large number of the independent vari-

ables (seven in the non stationary case), but also from a very

complicated structure of the collision integral. Therefore, in

the late sixties, several models of kinetic equations were

suggested (BGK, ES, S models), where the collision integral

I. A. Graur (&) � A. P. Polikarpov

Universite de Provence, Ecole Polytechnique Universitaire de

Marseille, UMR CNRS 6595, 5 rue Enrico Fermi,

13453 Marseille, France

e-mail: [email protected]

123

Heat Mass Transfer (2009) 46:237–244

DOI 10.1007/s00231-009-0558-x

is replaced by a relaxation term, which preserves various

important properties of the Boltzmann collision integral.

The main aim of the present paper is a test of validity

and of applicability of three kinetic models to describe

correctly the heat flux in the transitional flow regime. The

classical problem of a gas confined between two heated

plates is chosen as test case. This problem has been studied

by many authors, but especially using the linearized for-

mulation of the full Boltzmann or kinetic model equations

[3, 15, 17, 19]. But some of them have also solved the non-

linearized kinetic equations [12, 13]. In the present analysis

the numerical solutions of the heat transfer problem

between two parallel plates are obtained using three non-

linear model kinetic equations: BGK, S, ES models in the

transitional flow regime. These solutions are compared

with the numerical solution of the full Boltzmann equation

associated to the hard sphere model [12], with the DSMC

results [22] and with the experimental data [1, 18]. In

addition for the free molecular regime the numerical results

are compared with an analytical solution [10].

2 Kinetic models

To obtain the flow parameters in the transition regime the

Boltzmann equation should be solved:

of

otþ v � of

or¼ Qðf ; f 0Þ; ð1Þ

where t is the time, f(t, r, v) is the one particle distribution

function, r is the position vector, v is the molecular

velocity, Q(f, f0) is the collision integral. This equation

provides reliable numerical data but requires great

computational efforts. To reduce these efforts the

collision integral may be simplified and replaced by a

relaxation term, remaining its main properties:

of

otþ v � of

or¼ Iðf ; f modÞ; ð2Þ

where I (f, fmod) is the model collision integral

Iðf ; f modÞ ¼ f mod � f

s; ð3Þ

fmod, so called model distribution function, is an equilib-

rium distribution function which depends on the kinetic

model, s is a relaxation time. Different kinetic models were

proposed in the last sixties: BGK [4], S-model [14] and

ellipsoidal model (ES-model) [9]. In the present paper we

apply all these kinetic models for the steady flow simula-

tion and in the next section we will give a brief description

of each one.

When the distribution function is known, the macro-

scopic parameters may be calculated from the following

expressions:

number density and bulk velocity

nðrÞ ¼Zþ1

�1

f ðr; vÞdv; uðrÞ ¼ 1

n

Zþ1

�1

vf ðr; vÞdv; ð4Þ

temperature and heat flow vector

TðrÞ ¼ m

3nk

Zþ1

�1

V2f ðr; vÞdv;

qðrÞ ¼ m

2

Zþ1

�1

VV2f ðr; vÞdv;

ð5Þ

stress tensor

PijðrÞ ¼ m

Zþ1

�1

ViVjf ðr; vÞdv; ð6Þ

where V = v - u(r) is the peculiar velocity.

2.1 BGK

For the BGK kinetic equation [4] the equilibrium distri-

bution function fmod is the local Maxwellian distribution:

f mod ¼ f Mðn; T ; uÞ

¼ nðrÞ m

2pkTðrÞ

� �3=2

exp �mðv� uðrÞÞ2

2kTðrÞ

" #; ð7Þ

n is the number density, u is the bulk velocity, m is

the molecular mass of the gas, T is the gas temperature, k is

the Boltzmann constant. According to the construction

of the model kinetic equation the first five moments of

exact collision integral Q(f, f0) coincide with the first five

moments of the model collision integralZwQðf ; f 0Þdv ¼

ZwIðf ; f modÞdv ¼ MðwÞ;

where w ¼ 1; v; v2; ð8Þ

where here M(w) = 0. These properties derive from the

conservation of particles, momentum and energy in every

collision. Consequently, Eq. 2 provides the same macro-

scopic conservation laws as the Boltzmann equation 1 itself.

In this model the relaxation time is assumed to be

independent of the molecular velocity. The following

expression

s ¼ lp; ð9Þ

where p is the gas pressure and l is the shear viscosity,

provides the convenient expression of l. Since using

expression (9) in Eq. 2 and applying the Chapman–Enskog

method respectively to Eqs. 1 and 2 one obtains the same

expression of the viscosity coefficient.

238 Heat Mass Transfer (2009) 46:237–244

123

A number of important properties of the Boltzmann

collision integral are shared by the BGK model [9],

notably:

– The collision term drives the distribution function

toward equilibrium and vanishes, if and only if, the

distribution function fmod is equal to the local

Maxwellian equilibrium function;

– the collision term conserves molecular mass, momen-

tum, and kinetic energy;

– the collision term is a v-functional of f(r, t, v), i.e., it

depends upon the values of f over the entire range of the

v-space but only considering the values taken at a

single point in (r, t) space;

– the collision term is invariant in form under rotation or

translation of coordinates.

The main shortcoming of this kinetic model is that the

Prandtl number is obtained equal to 1.

2.2 S-model

The author of [14] proposed a technique for constructing an

approximate kinetic equation which is based on an

approximation of the collision integral. For this approxi-

mate equation the first few moment equations coincide

with the exact moment equations. This approximate

equation, called Shakhov model kinetic equation, is a

generalization of the BGK model equation in the sense that

the approximation condition (8) is satisfied not only for 1,

v, v2, but also for viv2 momentum which is not equal to

zero. Consequently this model provides the correct Prandtl

number equal to 2/3 for monoatomic gases. For this model

the equilibrium distribution function is given as

f mod ¼ f Sðn; T ; uÞ

¼ f M 1þ 2m

15nðrÞðkTðrÞÞ2V � qðrÞ mV2

2kTðrÞ �5

2

� �" #;

ð10Þ

where the bulk velocity u and the heat flux vector q are

defined by (4), (5).

Since the expression for fS contains the third-order

polynomial of v the distribution function may become

negative in the region of the large values of the velocities,

where the distribution itself is small. The H theorem is not

proved, but the author [20] gives some comments about the

correctness of the H-theorem for the S-model.

2.3 ES model

Holway [9] proposed a model based on the requirement for

the model distribution function to have the same first five

moments as the distribution function solution of the model

kinetic equation. In addition, the maximal probability

principle is used to obtain this ellipsoidal model [9]. This

model retains much of the mathematical simplicity of the

BGK model, but also yields the correct Prandtl number for

a monoatomic gas. In this model the equilibrium distribu-

tion function is replaced by the generalized Gaussian in the

following form [9]:

f mod ¼ f ESðn; T; uÞ ¼ n

p3=2ffiffiffiffiffiffiffiffiffiffiffiffidet Aij

p exp �X3

i;j¼1

A�1ij ViVj

" #;

ð11Þ

here Aij is a tensor with components

Aij ¼2kT

m

1

Prdij �

2ð1� PrÞnmPr

Pij; ð12Þ

Aij-1 is the tensor inverse to Aij, dij is the Kronecker symbol,

the stress tensor Pij is given by (6). The relaxation time for

the ES-model is defined as follows

s ¼ lpPr

: ð13Þ

The H-theorem was proved recently in [2] for this model.

3 Application to the heat flux problem

3.1 Physical conditions and normalization

A gas is confined between two infinite parallel plates,

y = -H/2 and y = H/2, the plates are considered at dif-

ferent but constant temperatures T�w ¼ T0 � DT and

Tþw ¼ T0 þ DT . We supposed here that the flow is sta-

tionary and depends only on the y-coordinate, notably

because the plates are infinite and at constant temperatures.

We assume also the bulk velocity u and the heat flux vector

q to have only two components: u = (ux, uy, 0) and

q = (qx, qy, 0). For some flow conditions, which will be

noted separately, we consider also the case where the

higher plate is moving with a relative constant velocity

uw = (uw, 0, 0) in the x direction and the lower plate with

the same velocity in the opposite direction. Then in the

following we consider only the stationary case.

The gas rarefaction is characterized by the parameter d:

d ¼ Hp

lvm; vm ¼

ffiffiffiffiffiffiffiffi2kT

m

r; ð14Þ

where vm is the most probable molecular speed. Since the

viscosity is proportional to the molecular mean free path,

the rarefaction parameter d is inversely proportional to the

Knudsen number.

The following dimensionless quantities have been

introduced:

Heat Mass Transfer (2009) 46:237–244 239

123

�r ¼ r

H; �T ¼ T

T0

; v0 ¼ffiffiffiffiffiffiffiffiffiffi2kT0

m

r; c ¼ v

v0

; �n ¼ n

n0

;

�f ¼ fv3

0

n0

; �s ¼ sv0

H; �u ¼ u

v0

; �q ¼ 2

n0mv30

q;

Pij ¼Pij

p0

; �l ¼ ll0

: ð15Þ

The reference pressure p0 is equal to n0kT0, where index

0 means the flow conditions on the symmetry axis

(y = 0).

The reference value of the viscosity coefficient l0 at T0

is taken from [5]. The hard-sphere model is used here, so

the viscosity coefficient is equal to �l ¼ffiffiffiffi�T

p:

In the following we omit the bars over the dimensionless

variables.

3.2 Reduced equations

The dimensionless form of the model Eq. 2 in this spatially

one-dimensional and steady case is written as

cyof

oy¼ d0n

ffiffiffiffiTpðf mod � f Þ: ð16Þ

Equation 16 is one-dimensional in the physical space,

but the molecular velocity space has three independent

variables. However, it is possible to reduce the

computational costs, introducing the reduced distribution

function. Since we supposed that the bulk velocity and the

heat flux vector have only two components (the third

components along the z-direction are equal to 0), we can

eliminate easily the dependence of the distribution function

on the molecular velocity cz. Multiplying Eq. 16 by 1 and

by cz2, and integrating it with respect to cz we obtain two

reduced equations

cyoUoy¼ d0n

ffiffiffiffiTpðUmod � UÞ;

cyoWoy¼ d0n

ffiffiffiffiTpðWmod �WÞ; ð17Þ

where the reduced functions U and W are introduced in

order to eliminate the variable cz

Uðy; cx; cyÞ ¼Z

f ðy; cÞcz;

Wðy; cx; cyÞ ¼Z

f ðy; cÞc2z cz:

ð18Þ

In the BGK model frame the reduced model distribution

functions, obtained from Maxwellian function (7),

according to (18) have the form

Umod ¼ UM ¼ n

pTexp �ðc� uÞ2

T

" #;

Wmod ¼ WM ¼ 0:5TUM :

ð19Þ

In the S-model frame the reduced model distribution

functions, obtained from (10), have the form

Umod ¼ US ¼ UM 1þ 4

15

ðc� uÞqnT2

ðc� uÞ2

T� 2

! !;

Wmod ¼ WS ¼ WM 1þ 4

15

ðc� uÞqnT2

ðc� uÞ2

T� 1

! !:

ð20Þ

If the ES-model is used, the dimensionless form of the

tensor (10) reads

Aij ¼T

Prdij �

1� Pr

nPrPij: ð21Þ

The reduced model distribution functions are written as

Umod ¼ UES ¼ n

pffiffiffiap exp �

X2

i;j¼1

A�1ij ðci � uiÞðcj � ujÞ

" #;

Wmod ¼ WES ¼ 0:5A33UES; ð22Þ

where a = A11A22 - A12A21.

3.3 Kinetic boundary conditions

The solution of the model kinetic equation needs a

boundary condition for the distribution function f. Despite

considerable efforts to understand the process of energy

and momentum transfer between gas flow and the solid

surface, models for detailed gas-surface interaction mech-

anism are still lacking [21]. In this paper we will use the

Maxwell-type diffuse-specular boundary condition at the

wall. For the down plate it reads

fþð�H=2; cx; cy; czÞ¼ ð1� aÞf�ð�H=2; cx;�cy; czÞ þ afw; ð23Þ

where, f? and f- are the distribution functions of the

incident and reflected molecules respectively, fw is distri-

bution function of particles fully accommodated with the

wall, fw ¼ f Mw ¼ nw

pT�wexp½�ðc�uwÞ2

T�w�; and where a is the

accommodation coefficient.

This formulation of the boundary condition does not

allow to distinguish the accommodated kinetic quantity, in

this theoretical frame a may represent either the accom-

modation coefficient of any molecule momentum compo-

nent or that of the molecule energy. In addition, if the

molecules with the internal degree of freedom are consid-

ered, this accommodation coefficient may be associated

240 Heat Mass Transfer (2009) 46:237–244

123

with the averaged quantity defined over all the energy

modes.

Therefore generally the accommodation coefficient adefined by the Maxwell scattering kernel can not be

directly related to a measured accommodation coefficient

without careful analysis: indeed the meaning and the value

of a experimental determination depend on the physical

conditions but also on the experimental technique. In the

present study we will associate the accommodation coef-

ficient to the energy accommodation, because we consider

here only heat transfer problem. Nevertheless the Maxwell

scattering kernel remains the basic and the most usual

model in numerical simulation owing to its simple phe-

nomenological description.

Moreover, using the impermeability condition (uy = 0)

the value of the number density nw at the wall may be

obtained as

nw ¼2ffiffiffippffiffiffiffiffiffiTw

pZ1

�1

Z1

0

cyUMdcxdcy; ð24Þ

where Tw is the temperature of the plate.

3.4 Numerical analysis

The distance between the plates -1/2 By B1/2 is divided in

Ny equal intervals. The second order central difference

scheme is used to approximate the transport terms of Eq.

17. In the molecular velocity space the discretization is

performed by the discrete velocity method. For the discrete

velocities the Cartesian coordinates (cx, cy) and the polar

coordinates (cp, u) are used for comparison. In the Carte-

sian coordinates the computational grid covers the interval

[-5, 5]. This finite interval is divided in Ncxand Ncy

uni-

form intervals for the velocities cx and cy respectively. In

the polar coordinates the angle u varies from 0 to 2p and

the uniform grid is implemented. For the variable cp the

Gaussian abscissas, calculated in [16] corresponding to the

weight function

wðcpÞ ¼ e�c2p cp ð25Þ

are used. All computations are performed with the accuracy

10-8 calculated for the temperature variations.

When the plates are at rest the solution of the present

two-surface problem is not determined uniquely by the

basic equation and boundary condition. Then in order to

determine as unique solution, either the value of density at

a certain position or that of average density between the

plates is required. Since the rarefaction parameter is

defined by using n0, i.e. the number density at y = 0, the

above boundary value problem is solved under the condi-

tion n(0) = 1 in the free molecular regime and under the

condition $-1/2?1/2ndy = 1 in the transitional regime [12].

4 Results and discussion

4.1 Free molecular regime

In the free molecular regime the molecular mean free path

tends to infinity and the rarefaction parameter tends to zero,

so the collision integral may be neglected. An analytical

solution of this problem subjected to the Maxwellian

specular-diffusion boundary condition with the same

accommodation coefficient a for the both plates was pro-

posed by Kogan in [10]. The distribution function depends

on the sign of the molecular velocity cy and does not

depend on the spatial coordinate. In [10] the macroscopic

parameters are expressed as follows:

n ¼ 1

2ðnþ þ n�Þ; ð26Þ

ux ¼ �uw

ffiffiffiffiffiffiTþp

�ffiffiffiffiffiffiT�p

ffiffiffiffiffiffiTþp

þffiffiffiffiffiffiT�p ; ð27Þ

T ¼ffiffiffiffiffiffiffiffiffiffiffiffiT�Tþp

þ 8

3u2

w

ffiffiffiffiffiffiffiffiffiffiffiffiT�Tþp

ðffiffiffiffiffiffiT�p

þffiffiffiffiffiffiTþp

Þ2; ð28Þ

Pxy ¼ �a

2� auwffiffiffipp n�

ffiffiffiffiffiffiT�p

þ nþffiffiffiffiffiffiTþp� �

; ð29Þ

qy ¼ �a

2� a1ffiffiffipp nþðTþÞ3=2 � n�ðT�Þ3=2� �

þ a2� a

u2wffiffiffipp nþ

ffiffiffiffiffiffiTþp

þ n�ffiffiffiffiffiffiT�p� � ffiffiffiffiffiffi

Tþp

�ffiffiffiffiffiffiT�p

ffiffiffiffiffiffiTþp

þffiffiffiffiffiffiT�p :

ð30Þ

Here n±, T ± are the number density and the

temperature of reflected molecules near the higher and

lower plates respectively. If the energy accommodation

coefficient is supposed equal to 1, the temperatures T ±

are equal to the temperatures T;w respectively. As it was

mentioned above, when the plates are at rest, the level

of the number density must be given. Here n(0) = 1 is

specified. One of the two values of the number density

may be found from the impermeability condition, namely

from nþffiffiffiffiffiffiTþp

¼ n�ffiffiffiffiffiffiT�p

. Therefore, Eqs. 26–30 define

completely the solution. It is to note that the macroscopic

parameters do not depend on the spatial coordinate y.

Equation 16 subjected to boundary condition (23) is

solved numerically. In order to verify the numerical

accuracy of the discretization in the molecular velocity

space some test cases are fulfilled. Two points in the

physical space are used in the numerical calculations. The

number of the required points in the velocity space depends

on the flow parameters (plate temperatures and velocities).

According to the analytical solution all parameters should

be constant between the plates. The accuracy of the

numerical calculations is determined through the control of

Heat Mass Transfer (2009) 46:237–244 241

123

the heat flux fluctuation. In order to achieve a variation of

the heat flux less than 0.1%, for a small temperature dif-

ference (DT ¼ 0:1) and for the walls at rest, 25 points in

both velocity directions are enough. But for an important

temperature difference (DT ¼ 0:9) and when the wall

velocities equal uw = 0.9, the number of points required in

the velocity space increases up to Ncx¼ 50 and Ncy

¼ 150

to obtain the same accuracy. The following temperature

differences are considered DT ¼ 0; 0:1; 0:14, 0.5764, 0.9.

The plate velocities are equal to uw = 0, 0.1, 0.9. The

values of the macroscopic quantities n, ux, T, pxy, qy

obtained in the numerical simulations coincide with the

analytical values given by Eqs. 26–30 until five significant

numbers.

It is to note that the commonly used analytical expres-

sion of the free molecular normal heat flux

qFM ¼ �a

2� a2DTffiffiffi

pp ð31Þ

is only valid for the small temperature difference between

the plates (DT�1). For example, when DT ¼ 0:14 the

difference between expressions (30) and (31) is 0.7%, but

for DT ¼ 0:5764 this difference grows up to 14%.

4.2 Transitional regime

4.2.1 Small temperature difference

In transitional flow regime the calculations are carried out

for d = 1.1906, 3.0150, 4.6483, 6.4710, 13.7189. The

temperature difference between the plates is equal to DT ¼0:14 and the both plates are at rest. The accommodation

coefficient for the both plates is supposed to be the same

and equal to 0.826. These conditions correspond to the

conditions of the density measurements of argon in [18]

and also to those of the numerical simulation using the full

non-linear Boltzmann equation for the hard-sphere mole-

cules in [12].

In the numerical simulations the distance between the

plates is divided into 100 uniform sections. In the velocity

space, when the Cartesian coordinates are used, the fol-

lowing number of points Ncx¼ 40 and Ncy

¼ 200 are

needed. The implementation of the polar coordinates sys-

tem and, especially, the Gaussian integration rule [16]

allows to reduce the number of computational points until

Ncp¼ 16 and Nu = 101. The accuracy of the present

simulation is confirmed by the constancy of the heat flux

qy. The variation of this flux should be theoretically zero,

but practically it is less than 0.1% in the numerical

calculations.

Let us to note that in spite of temperature difference that

does not seem small enough everywhere, the analytical

value (31) obtained from the linearized technique is used

here, as in [12], for the normalization of the calculated heat

flux. This normalized heat flux obtained from the numerical

simulation using respectively BGK, S-model and ES-

model is shown in Table 1. For comparison, the numerical

results from the full non-linear Boltzmann equation are

presented also in Table 1. The difference between the

results derived from the full non-linearized Boltzmann

equation [12] and the results of the ES kinetic model is less

then 3% for d = 13.7189 and less then 1% for

d = 1.19058. The results of S-model are also close to that

of ES model. But the BGK model gives values of heat flux

much smaller owing to the non-correct value of the Prandtl

number.

Figure 1(left) shows comparisons of calculated and

experimental measured density profiles. The density has

been normalized by the centerline value. Figure 1(left)

gives results for d = 1.1906 and 13.7189. The agreement

between numerical results given by both models is good for

d = 13.7189, but when d decreases (Kn number increases)

the discrepancy between the numerical results and the

experimental points increases, especially near the walls.

The temperature profiles calculated by S- and ES-models

are also shown in Fig. 1(right). The S and ES models

results are in good agreement. It is to note, that the bulk

velocity between the plates is close to zero.

In the heat-transfer measurements [18] the temperature

difference between the plates was about DT ¼ 0:013. The

calculations of the heat flux are also carried out for this

temperature difference and for seven values of the rare-

faction parameter listed in Table 2 and plotted in Fig. 2. In

the Fig. 2 the experimental values of the normalized heat

flux are shown and compared with the results of the S and

ES model simulations. The agreement between the

numerical results and the experimental data is not every-

where very good: the discrepancy varies between 4 and

12% and appears greater than the uncertainty of the mea-

surements evaluated as *3%.

One of the reasons of this discrepancy in the heat flux

may be the value of the accommodation coefficient derived

from inappropriate experiment [18]. In the experiment [18]

the thermal accommodation coefficient was derived from

Table 1 Normalized heat flux qy/qFM, comparison between different

models

d BGK S-model ES-model Full Boltzmann [12]

1.1906 0.6759 0.7393 0.7488 0.7558

3.0150 0.4774 0.5637 0.5684 0.5807

4.6483 0.3811 0.4694 0.4719 0.4843

6.4710 0.3117 0.3966 0.3980 0.4094

13.7189 0.1813 0.2462 0.2466 0.2538

242 Heat Mass Transfer (2009) 46:237–244

123

the heat flux measurements in free molecular regime and

for a small temperature difference. This coefficient a is

obtained using relation (31) that is correct in the conditions

of the experiments. But we use then this coefficient a,

which appears in condition (23), for the transitional flow

regime calculations, that is to say in rarefaction conditions

different from those of the experiments.

4.2.2 High temperature difference

Another comparison is fulfilled between the numerical

results and the experimental measurements of the density

profiles [1] for a larger temperature difference between the

plates DT ¼ 0:5764: where the hot plate is at the ambient

temperature (Tþw ¼ 294 K) and the cold plate is at

T�w ¼ 79 K. In addition to the density measurements, heat

transfer measurements were also made in [1] to evaluate the

thermal accommodation coefficients at the hot and cold

surfaces. But finally, for technical reasons (experimental

accuracy) the experimental accommodation coefficient was

derived only for a small temperature difference Tþw �T�w � 294 K and, as in [18], in free molecular regime. The

value of the accommodation coefficient a = 0.58 was

retained in the calculations for the both hot and cold plates

despite the fact that the two plates used in the experimental

study were of different materials and despite the fact that we

simulate the transitional flow regime with a cold plate very

much colder than in the experiment [1].

The density profiles, obtained from S and ES models,

are compared with the measurements of [1] in Fig. 3(left).

For both values of the rarefaction parameter (d = 2.2624

and d = 12.036) the agreement between the measured and

calculated values of density becomes worse near the high

temperature plate than near the colder one, that seems

surprising considering the previous remarks.

In Fig. 3(right) the temperature profiles from S and ES

models are plotted and compared with the results of the

DSMC simulation of [22]. The ES-model results seems to

be more close to the DSMC simulation, than the S-model

results. Finally it is to note the important temperature dif-

ference (jump) between the wall temperature and the gas

temperature near the wall. This temperature jump increases

when the gas rarefaction increases.

Y/H

n/n

0-0.5

-0.5

-0.25

-0.25

0

0

0.25

0.25

0.5

0.5

0.9 0.9

0.95 0.95

1 1

1.05 1.05

1.1 1.1δ=13.72

δ=1.19

Y/H

T/T

0

-0.5

-0.5

-0.25

-0.25

0

0

0.25

0.25

0.5

0.5

0.9 0.9

0.95 0.95

1 1

1.05 1.05

1.1 1.1

δ=13.72

δ=1.19

Fig. 1 Small temperature

difference DT ¼ 0:14. Left Fig:

Profiles of normalized density

for different models: solid lineES-model, dash-dotted lineS-model, circles: experiment

[18]. Right Fig: Profiles of

normalized temperature:

solid line ES-model,

dash-dotted line S-model

Table 2 Normalized heat flux qy/qFM, comparison with the experi-

mental data [18]

d [18] ES-model

0.2512 0.9654 0.9253

0.5844 0.9332 0.8517

0.8781 0.8871 0.7992

1.8840 0.7396 0.6686

4.6665 0.5346 0.4721

9.8330 0.3525 0.3100

16.0192 0.2442 0.2203

δ

Q/Q

FM

0

0

5

5

10

10

15

15

0.2 0.2

0.3 0.3

0.4 0.4

0.5 0.5

0.6 0.6

0.7 0.7

0.8 0.8

0.9 0.9

1 1

ExperimentS-modelES-modelBoltzmann eq. HS

Fig. 2 Comparison of the normalized heat flux qy/qFM for different

kinetic models and the measured values from [18]

Heat Mass Transfer (2009) 46:237–244 243

123

5 Conclusion

The numerical solutions of the heat transfer problem of

rarefied gas between parallel plates have been obtained for

different non-linear kinetic models (BGK, S-model and

ES-model) using the Maxwell boundary condition. Com-

parisons have been made with the numerical solution of the

full non-linear Boltzmann equation for the hard-sphere

molecules, with the results of the DSMC simulation and

with available experimental data. The conclusions are

summarized in follows:

– The ellipsoidal ES model gives the results more close

to the results from the non-linear Boltzmann equation

and to the DSMC results.

– The discrepancy between the experimental and numer-

ical values of the heat flux may be explained notably by

the ‘‘non-perfect’’ diffuse-specular Maxwell kernel.

As we said above in spite of its limits we used here only

the Maxwell scattering kernel, but the implementation of

another scattering models [6–8, 11] would be of great

interest for further comparison between various boundary

condition models.

Acknowledgments The research leading to these results has

received funding from the European Community’s Seventh Frame-

work Programme (ITN - FP7/2007-2013) under grant agreement

n 215504. We should also like to acknowledge J. Gilbert Meolans for

valuable discussions.

References

1. Alofs DJ, Flagan RC, Springer G (1970) Density distribution

measurements in rarefied gases contained between parallel plates

at high temperature difference. Phys Fluids 14(3):529–533

2. Andries P, Bourgat J-F, Le Tallec P, and Perthame B (2002)

Numerical comparison between the boltzmann and es-bgk models

for rarefied gases. Comput Methods Appl Mech Eng 191(31):

3369–3390

3. Bassanini P, Cercignani C, Pagani CD (1968) Influence of the

accommodation coefficient on the heat transfer in a rarefied gas.

Int J Heat Mass Transf 11:1359–1368

4. Bhatnagar PL, Gross EP, Krook M (1954) A model for collision

processes in gases. Part I: small amplitude processes in charged

and neutral one-component systems. Phys Rev 94(3):511–525

5. Bird GA (1994) Molecular gas dynamics and the direct simula-

tion of gas flows. Oxford Science Publications

6. Cercignani C, Lampis M (1971) Kinetic models for gas-surface

interactions. Transp Theory Stat Phys 1:101–114

7. Dadzie K, Meolans JG (2004) Anisotropic scattering kernel:

generalized and modified Maxwell boundary conditions. J Math

Phys 45(5):1804–1819

8. Epstein M (1967) A model of the wall boundary condition in

kinetic theory. AIAA J 5:1797–1800

9. Holway LH (1966) New statistical models in kinetic theory:

methods of construction. Phys Fluids 9(9):1658–1673

10. Kogan MN (1969) Rarefied gas dynamics. Plenum Press, New York

11. Nocilla S (1963) The surface re-emission law in free molecule

flow. In: Laurmann JA (ed) Rarefied gas dynamics. Academic

Press, New York, pp 327–346

12. Ohwada T (1996) Heat flow and temperature and density distribu-

tions in a rarefied gas between parallel plates with different tem-

peratures. Finite-difference analysis of the nonlinear Boltzmann

equation for hard-sphere molecules. Phys Fluids 8(8):2153–2160

13. Pantazis S, Valougeorgis D (2008) Heat transfer between parallel

plates via kinetic theory in the whole range of the knudsen

number. In: 5th European thermal-sciences conference, paper

409, Eindhoven, 18–22 May 2008

14. Shakhov EM (1968) Generalization of the Krook kinetic relax-

ation equation. Fluid Dyn 3(5):95–96

15. Sharipov F, Seleznev V (1998) Data on internal rarefied gas

flows. J Phys Chem Ref Data 27(3):657–706

16. Shizgal B (1981) A Gaussian quadrature procedure for use in the

solution of the Boltzmann equation and related problems.

J Comput Phys 41:309–328

17. Sone Y (2002) Kinetic theory and fluid mechanics. Birkhauser,

Boston

18. Teagan WP, Springer GS (1968) Heat-transfer and density-dis-

tribution measurements between parallel plates in the transition

regime. Phys Fluids 11(3):497–506

19. Thomas JR, Chang TS, Siewert CE (1973) Heat transfer between

parallel plates with arbitrary surface accommodation. Phys Fluids

16:2116

20. Titarev VA (2007) Conservative numerical methods for model

kinetic equations. Comput Fluids 36(9):1446–1459

21. Trott WM, Rader DJ, Gastaneda JN, Torczynski JR, Gallis MA

(2009) Measurements of gas-surface accommodation. In Abe T

(ed) Rarefied gas dynamic, 26th international symposium on

rarefied gas dynamics, vol 1084. AIP, New York

22. Wadsworth DC (1993) Slip effects in a confined rarefied gas. Part

I: temperature slip. Phys Fluids A 5(7):1831–1839

Y/H

n/n

(0)

-0.5

-0.5

-0.25

-0.25

0

0

0.25

0.25

0.5

0.5

0.8 0.8

1 1

1.2 1.2

1.4 1.4

1.6 1.6

1.8 1.8

δ=12.036

δ=2.2624

Y/H

T/T

0

-0.5

-0.5

-0.25

-0.25

0

0

0.25

0.25

0.5

0.5

0.6 0.6

0.8 0.8

1 1

1.2 1.2

1.4 1.4

δ=12.036

δ=2.2624

Fig. 3 High temperature

difference DT ¼ 0:5764. LeftFig: Profiles of normalized

density for different models:

solid line ES-model, dash-dotteded line S-model, circles:

experiment [18]. Right Fig:

Profiles of normalized

temperature: solid lineES-model, dash-dotted lineS-model, circles:

DSMC results [22]

244 Heat Mass Transfer (2009) 46:237–244

123