Lea,b,*, Ngoc Anh Vuc, Le Tan Locd a Divison of Computational Mathematics and Engineering, Institute for Computational Science, Ton Duc Thang University, Ho Chi Minh City, Viet Nam b Fac
Trang 1Effect of the sliding friction on heat transfer in high-speed rare fied gas
flow simulations in CFD
Nam T.P Lea,b,*, Ngoc Anh Vuc, Le Tan Locd
a Divison of Computational Mathematics and Engineering, Institute for Computational Science, Ton Duc Thang University, Ho Chi Minh City, Viet Nam
b Faculty of Civil Engineering, Ton Duc Thang University, Ho Chi Minh City, Viet Nam
c Department of Aerospace Engineering, Ho Chi Minh City University of Technology, VNU-HCM, Viet Nam
d Faculty of Aviation Technology, Vietnam Aviation Academy (VAA), Ho Chi Minh City, Viet Nam
a r t i c l e i n f o
Article history:
Received 25 January 2016
Received in revised form
15 May 2016
Accepted 8 June 2016
Keywords:
Sliding friction
Heat transfer
Surface gas temperature
NACA0012 airfoil
Biconic
a b s t r a c t
Different nature of the heat transfer computation between the CFD and DSMC methods in rarefied gas flow simulations leads to the difference of their heat transfer results This paper will revisit a formulation
of heat transfer in CFD that includes the sliding friction part for a planar surface The sliding friction was first introduced by Maslen, and used to be omitted in calculating the CFD heat transfer This formulation will be extended to compute the heat transfer over the curved surfaces such as NACA0012 micro-airfoil and the sharp 25e55-deg biconic Two these configurations 1) NACA0012 micro-airfoil (M ¼ 2) with various Kn¼ 0.026, 0.1 and 0.26 with the angles-of-attack from 0-deg to 20-deg., and a sharp 25e55-deg biconic (M¼ 15.6) are simulated by the CFD method, which solves the Navier-Stokes equations within the OpenFOAM framework Moreover, the flat plat cases with various wall temperatures (Tw¼ 77 K and 294 K) are also undertaken to investigate the significance of the sliding friction on the heat transfer The CFD simulation results of the heat transfer involving the sliding friction give good agreements with the DSMC data, especially the NACA0012 micro-airfoil cases with high Knudsen numbers Kn¼ 0.1 and 0.26
© 2016 Elsevier Masson SAS All rights reserved
1 Introduction
The design of high speed vehicles requires accurate predictions
of the surface properties in theflight These quantities are typically
the surface temperature, the surface pressure, the shear stress, and
heat transfer The surface pressure and the skin friction forces
acting on the surface of the vehicles are integrated to calculate the
aerodynamic forces and the moments as lift, drag and pitching
moments The peak surface temperature or heatflux are integrated
over time and mapped over the vehicle surface as part of the
pro-cess to design the thermal protection system In this paper we will
focus on the wall heat transfer work, and try to reduce the gap
between the DSMC and CFD simulation results The typical
methods have been used for simulating rarefied gas flows as the
CFD and DSMC methods In the CFD method, in particular solving
the Navier-Stokes (NS) equations, the Fourier heat conduction law has been adopted for computing the wall heat transfer In the DSMC method the wall heat transfer has been calculated based on the velocities[1] The lack of the velocity term in CFD calculation al-ways results in the difference between the CFD and DSMC heat transfers In order to overcome this problem, Maslenfirst intro-duced the sliding friction part to compute the wall heat transfer in CFD for a planar surface[2]without any physical explanation The sliding friction (or shear work) had been employed in the energy equation in CFD for simulating the low-speed rarefied gas micro-flows, and gave good results[3e7] In high-speed rarefied gas flows, the wall heat transfer included the sliding friction was evaluated to compare with experimental data for rarefied gas flows over the flat plate[8], and agreed well with experiment In CFD rarefied gas simulations, we often use the slip velocity and temperature jump boundary conditions applied to the surface to improve the NS equations in the slip regime Since the slip velocity boundary condition is applied to the surfaces leading to the appearance of the sliding friction of gas molecules over the surface This slip will generate more heat transfer between gas molecules and the sur-face In the past, there were the work[8e12] using the sliding
* Corresponding author Divison of Computational Mathematics and Engineering,
Institute for Computational Science, Ton Duc Thang University, Ho Chi Minh City,
Viet Nam.
E-mail address: letuanphuongnam@tdt.edu.vn (N.T.P Le).
Contents lists available atScienceDirect International Journal of Thermal Sciences
j o u r n a l h o m e p a g e : w w w e l s e v ie r c o m / l o c a t e / i j t s
http://dx.doi.org/10.1016/j.ijthermalsci.2016.06.020
1290-0729/© 2016 Elsevier Masson SAS All rights reserved.
International Journal of Thermal Sciences 109 (2016) 334e341
Trang 2friction term proposed for planar surface to compute the wall heat
transfer However, many researchers have used the Fourier law of
heat conduction for calculating the heat transfer in their CFD high
speed rarefied gas flow simulations[13e22]that have omitted the
sliding friction part Recently Hong et al [9] pointed out the
physical reason of the inclusion of the sliding friction was
comprehensively explained in terms of both conservation law of
energy and the kinetic theory of gases Hadjiconstantinou [10]
derived an expression for fully developed slipflow and concluded
that the sliding friction (shear work) at the wall should be included
to calculate the heatflux from the wall when viscous heat
gener-ation needs to be considered Moreover, the boundary condition of
the heatflux of the higher moment method (R-13 equations)
pro-posed by Rana et al.[23]also included the slip velocity and stress
terms The simulation results in Refs.[9,10,23]gave good
agree-ment with DSMC and experiagree-mental data in internal rarefied gas
microflows
The heat transfer caused by the sliding friction part was
computed based on the slip velocity and the shear stress[2] Our
previous work[13]simulating rarefied gas flow of the NACA0012
airfoil the CFD heat transfers were calculated by the Fourier heat
conduction law, and did not match those of DSMC data They
showed that there were large differences between the DSMC and
CFD heat transfers since the gasflow velocity increases This can be
explained by the lack of the sliding friction part, which involves the
slip velocity and shear stress (e.g the velocity gradient) terms, in
the CFD heat transfer computation In the present work, we will
revisit a formulation of the wall heat transfer proposed by Maslen
[2]for a planar surface, and will extend this one to the curved
geometrical surfaces such as NACA0012 micro-airfoil and the sharp
25e55-deg biconic The NACA0012 micro-airfoil case with Mach
number, M¼ 2, is investigated with various angles-of-attack (AOA)
from 0-deg to 20-deg and Knudsen numbers from 0.026 to 0.26
The sharp 25e55-deg biconic case is simulated with M ¼ 15.6 The
flat plate cases [25](M ¼ 6.1) with different wall temperatures
Tw¼ 77 K and Tw ¼ 294 K are also simulated to investigate the
significance of the sliding friction on the wall heat transfer All
simulations are undertaken in the OpenFOAM framework[24] The
wall heat transfer involving the sliding friction is implemented in
OpenFOAM as a new post-processing utility The heat transfers
with and without the sliding friction in all CFD simulations will be
compared with those of the DSMC and experimental data
[13e15,26,27] We also investigate whether the heat transfer with
the sliding friction can capture that of DSMC at high Knudsen
numbers, Kn¼ 0.1 and 0.26 (i.e the transition regime)
2 CFD and DSMC heat transfers in rarefied gas flow
simulations
In CFD, the wall heat transfer is often calculated by the Fourier
heat conduction law that only depends on the temperature
gradient normal to the surface The DSMC wall heat transfer can be
obtained by sampling the difference of energyfluxes in which the
heat transfer is a function of molecular velocity, u[28]
qw DSMC¼tF$numDA
2
4XN
j¼1
1
2msimu
2 j
i
XN
j¼1
1
2msimu
2 j
r3
5 (1)
where Fnumis the number of real molecules each simulated particle
represents (weight factor); msim is the mass of the simulated
molecule in DSMC; The superscripts“i” and “r” denote the values of
before and after impacting the wall element, respectively; the
subscript“w” indicates the wall value; N is the total number of
molecules impacting the surface element during sampling time t;
andDA is the surface area
Due to different nature of the heatflux computation between two methods, there is always a gap between the CFD and DSMC wall heat transfer simulation results[13] We will therefore revisit the formulation of the heat transfer proposed by Maslen[2]that included the sliding friction part The heat transfer in the gas slip flow over the flat plate was first introduced by Maslen[2]including the Fourier heat conduction
kvT , and sliding friction
usmvu vy
parts,
q¼ kvTvy usmvuvy ; (2) where k is the thermal conductivity; T is the temperature; usis the slip velocity;mis the viscosity; and y is spatial coordinate in y di-rection (normal didi-rection) Maslen also stated that it was the cor-rect expression for heat transfer in slipflow without further proof
[2] It should be noted that this relation can be derived with the kinetic theory for a nonuniform gas Its derivation had been briefly presented in Ref.[8]and is now re-summarized here
In the gas kinetic theory the three dimensional motion of indi-vidual molecules between collisions is governed by the equation of motion which is a set of 3 s order differential equations They are expressed in both physical space (x, y, z) and velocity space (U, V, W) The velocity components (u, v, w) of a molecule can be referred
to their average values by writing,
u¼ u þ U; v ¼ v þ V; w ¼ w þ W[29], where u; v; w are the compo-nents of the mass velocity of the gas, and U, V, W are the compocompo-nents
of velocity of a molecule referred to the mass velocity[29] The next energyflux to a planar surface can be expressed as follows[8,29]
E¼1
2r Z
þ∞
∞
dU
Zþ∞
∞
dV
Zþ∞
∞
Vf
u2þ V2þ W2
where f is the Maxwell distribution function of molecular velocity;
r is density; and the definition of thermal velocity,
C2¼ U2þ V2þ W2[29] Since we attempt to calculate all the macroscopic properties of a gas from the motion of its molecules it
is clear that the function f is fundamental importance The role played by the law of distribution of molecular velocities becomes apparent when we consider the formation of the mean values which define the state of a gas [29] So the average value of a quantity Q is a function of the velocity components So we integral equation(3), and making the use of the integral definition of an average quantity Q[8,29]:
Q≡
Zþ∞
∞
dU
Zþ∞
∞
dV
Zþ∞
∞
The energyflux over a planar surface (flat plate) is expressed as:
[8,29]
E¼1
where u is now the velocity parallel to the surface (i.e the slip velocity, usat near the surface) By definition in Ref.[29], the second term is the component of the stress tensor such as shear stress From the momentum-transfer equations derived by the kinetic theory of gas and the equation of motion of thefluid element in the
x, y, and z directions according to Newton’s second law described in Ref.[29]we have
N.T.P Le et al / International Journal of Thermal Sciences 109 (2016) 334e341 335
Trang 3txy¼ rUV¼ mvvvx þvuvy; (6)
for the gasflow over a flat plate The gas flow over the flat plate
(planar surface) we then consider the flow in the horizontal
(streamwise) direction only So the termvvvxhas been neglected[8]
From the energy-transfer equations derived by the kinetic
the-ory of gas described in Ref.[29]we have
rVC2¼ 15
2mvyv C32
!
The definition of temperature in the kinetic theory of gases is
based on the assumption the molecules possess translational
en-ergy only in which the internal enen-ergy arises entirely from the
translational motion of the molecules[29]
So we have
rVC2¼ 15
2mRvT
where R is the gas constant, for a monatomic gas cv¼ 1.5 R the
coefficient of heat conduction k ¼ 2.5cvm[29], then1rVC2¼ kvT
So we have the heat transfer in slipflow over the flat plate is
presented[8]
q¼ kvTvy usmvuvy ¼ kvT
vy ustxy: (10) Derivation of heat transfer of the slip gasflow over the flat plate
figured out that there are two parts: 1) the usual contribution due
to conduction heatflux normal to the surfacekvT
, and 2) the second one due to sliding friction (or shear work), ustxy This
latter term is no zero if there is some slip at the surface[2] From
this one for gasflow over a flat plate we can express the general
formulation of the heat transfer of the slip gasflow over the curved
surface that we will consider the shear stresses and slip velocities in
three directions as follows:
q¼ kVnT us$S$ðn$PÞ (11)
where Vn≡n$V is the component of the gradient normal to
boundary surface; symbol“$” is inner product;Pis stress tensor;
tensor S¼ I¡nn, where n is the unit normal vector defined as
positive in the direction pointing out of the surface and I is identity
tensor, removes normal components of stresses Thefirst term on
the right hand side of equation(11)expresses the heatflux to the
gas, and the second term presents the sliding frictional work by the
slip over the curved surface Equation(11)is implemented as a new
post-processing utility in OpenFOAM to use with the solver
“rhoCentralFoam”
3 Nonequilibrium boundary conditions
As we consider the sliding friction part in the computation of
wall heat transfer, so the slip velocity and temperature jump
boundary conditions must be applied to the surfaces Thefirst order
Maxwell/Smoluchowski boundary conditions are adopted for our
CFD simulations Thefirst-order conventional slip condition was
developed by Maxwell in 1890 for aflat plate, based on his earlier
work on the theory of viscosity in gases Although it is not perfect, it
is still the simplest and most useful description of the slip condition that we have It is designed to work with the continuum governing equations in the CFD method, such as the NS equations Thefirst order Maxwell slip condition is expressed as follows[30,31]:
uþ
2su
su
lðS$VnuÞ ¼ uw
2su
su
l
mS$ðn$PmcÞ
3 4
m
wherelis the Maxwellian mean free path;gis the specific heat ratio;Pmc¼mðVuÞT
2
ItrðVuÞ
is the tensor; Pr is the Prandtl number;mis viscosity; p is the pressure and uwis the wall velocity The right-hand side of the equation(12)contains three terms that are associated with (in order): the surface velocity, the so-called curvature effect, and thermal creep The first-order slip condi-tions mentioned above, the accuracy of the simulation results is decided by a free parameter in the slip condition, namely the tangential momentum accommodation coefficient, which varies from 0 to 1 The tangential momentum accommodation coefficient,
su, determines the proportion of molecules reflected from the surface specularly (equal to 1su) or diffusely (equal tosu), and
0su 1
The temperature jump condition was an early development by Smoluchowski[32], and was driven by the heatflux to the surface
in the normal direction:
Tþ2sT
sT ðgþ 1ÞPr2g lVnT¼ Tw; (13) where Pr is the Prandtl number; Twis the wall temperature The temperature jump condition also depends on a free parameter,sT, namely the thermal accommodation coefficient, which varies from
0 to 1, and decides the accuracy of the simulation results The thermal accommodation coefficient is used to ascribe the temper-atures of the receding molecules Specularly reflected molecules recede from the surface with their original incident energy, and diffusely reflected molecules have their energies adjusted to those would arise in a mass of gas in equilibrium at the temperature of the surface Perfect energy exchange between the gas and the solid surface corresponds tosT¼ 1, and no energy exchange tosT¼ 0
If we make nondimensionalization with a reference length, ve-locity and temperature (such as free-stream veve-locity and temper-ature), these boundary conditions (Eqs (12) and (13)) can be expressed in the term of Kn as presented in Ref.[32] It is well known that the rarefaction effect depends on the Knudsen number
A rarefaction effect can be studied by solving the momentum and energy equations with the slip boundary conditions, in which the velocity slip, temperature jump, and shear stress work on the wall are taken into account[9]
The Maxwellian mean free path is defined as follows[33]:
l¼m r
ffiffiffiffiffiffiffiffiffi
p
2RT
r
where the viscositymis computed by the Maxwell model for two-dimensional gas[34]:
m¼
ffiffiffiffiffiffiffiffiffiffi
mkB
p
r
1
pd2T
where m is mass of a gas molecule, m¼ 46.5 1027kg for
nitro-gen, and 48.1 1027kg for air[1]; d is diameter of gas molecules, N.T.P Le et al / International Journal of Thermal Sciences 109 (2016) 334e341
336
Trang 4d¼ 4.17 1010m for nitrogen, and 4.19 1010m for air[1]; and
kBis the Boltzmann constant
Our CFD simulations are undertaken in OpenFOAM[24] The
open source CFD software, OpenFOAM, uses the finite volume
numeric to solve systems of partial differential equations ascribed
on any three-dimensional unstructured mesh of polygonal cells
The CFD solver“rhoCentralFoam” in the OpenFOAM which solves
the NS equations with the Maxwell slip and Smoluchowski jump
boundary conditions withsu¼sT¼ 1 is selected for performing all
our CFD simulations In this solver, the NS equations are
numeri-cally solved with the high-resolution central scheme described in
Ref.[17] The accommodation coefficients of unity are equivalent to
the diffuse reflections simulated in the DSMC data[13e15,26]
4 Numerical setups
The profile of the NACA 0012 airfoil is computed by the
expression below[35]:
y¼ 0:6
0:2969x
c
0:5
0:126x
c
0:3537x
c
2
þ 0:2843x
c
3
0:1015x
c
4
;
(16) where c is the chord length, c¼ 0.04 m; x is the running distance
along the chord (0 x c); y is the half thickness of the airfoil The
dimensions of geometries of the airfoil and biconic cases are
pre-sented in Figs 1a and 2a, respectively Numerical setups of
computational domains for these two cases are presented inFigs 1a
and 2b The computational domain of the biconic case is similar to that given in Ref.[13] The structured C-type mesh is used for the airfoil case, as seen inFig 1b, and thefinal mesh is adapted from our previous work in Refs.[13,14]with 600 cells on the airfoil surface; the values L¼ 6c and H ¼ 3c This mesh had been validated with DSMC and experimental data in the previous work[13,14] For the two-dimensional axisymmetric biconic case, the geometry is specified as a wedge of one cell thickness running along the plane
of geometry The axisymmetric wedge planes must be specified as separated patches of type“wedge”, as seen inFig 2b Structured mesh of the biconic case is presented inFig 2c Freestream con-ditions (indicated by the symbol“∞”), and working gas of all airfoil cases are presented inTable 1 The freestream conditions of the biconic case[26]are: p∞¼ 2.23 Pa, M∞¼ 15.6, T∞¼ 42.6 K, and the wall temperature Tw¼ 297.2 K with working gas as nitrogen A typical mesh for a flat plate simulation is regular rectangular A mesh independence analysis was completed tofind the final mesh (Dx ¼Dy¼ 0.117 103 m) for the convergence solution The
freestream conditions of theflat plate case[25]are: p∞¼ 2.97 Pa,
M∞¼ 6.1, T∞¼ 83.4 K, the length of plate L ¼ 0.0875 m, and the wall temperatures Tw¼ 77 K and 294 K with working gas as nitrogen
5 Heat transfer simulation results and discussions 5.1 Sharp 25e55-deg biconic case
Fig 3compares the CFD heat transfers with those of DSMC[26]
and experimental[27]data In 0< x < 0.05 m, the heat transfer with the sliding friction is close to DSMC data[26] The CFD simulations underpredict the heat transfers along the fore-cone in comparing with DSMC and experimental data There are increases in heat transfer across the recirculation zone The heat transfers signi fi-cantly drop in the range 0.072 m x 0.095 m within the area of the recirculation zone, and then rapidly increase the peak values of 1) 103.59 kW/m2at the location x¼ 0.1037 m for the heat transfer without sliding friction, 2) 98.88 kW/m2 at the location
x¼ 0.1045 m for that with the sliding friction, 3) 124.40 kW/m2at the location x¼ 0.1027 m for DSMC data, and 4) 112.93 kW/m2at the location x¼ 0.1023 m for experimental data in the area of the shock-shock interaction It is rational that DSMC and NS solutions disagree on some points due to following reasons, i.e enhanced non-equilibrium effects that undermine the NS basic assumptions
as continuumflow field Past these peak locations, the heat trans-fers rapidly drop and oscillate along the second cone Again the CFD results are lower than DSMC and experimental data than along this cone surface (0.105 m x 0.153 m) On the base surface, (0.153 m x 0.192 m), there is good agreement between all CFD, DSMC and experimental data Overall, the DSMC method predicts better heat transfer than the CFD method in comparing with the experimental data
5.2 NACA0012 airfoil cases 5.2.1 M∞¼ 2, AOA ¼ 0-deg, Kn ¼ 0.026 Considering the case with small Kn¼ 0.026 and AOA ¼ 0-deg shown inFig 4, DSMC data[13]reach a peak value of 240 W/m2
at the airfoil tip Past the peak value they gradually decrease to the location x/c¼ 0.1, and oscillate along the surface The heat transfer with the sliding friction obtains the peak value of 200 W/m2, and thereafter reduces gradually to the location x/c¼ 0.13 It is a nearly constantfinite value in 0.13 < x/c < 0.9, and increases in the range 0.9 x/c 1 There is a good agreement between the DSMC data and the heat transfer with the sliding friction The heat transfer without the sliding friction disagrees with the DSMC data and is the lowest values along the surface
Fig 1 a): Geometry, dimensions and numerical setups of the NACA 0012 airfoil case
N.T.P Le et al / International Journal of Thermal Sciences 109 (2016) 334e341 337
Trang 55.2.2 M∞¼ 2, AOA ¼ 10-deg, Kn ¼ 0.026 Continuing the case Kn¼ 0.026 with the change of AOA ¼ 10-deg, the calculated heat transfers are shown in Fig 5 for the lower surface The heat transfer without the sliding friction dis-agrees with the DSMC data [13] and predicts the lowest heat transfer values along the surface The DSMC data reach a peak value
c
Fig 2 a): Geometry and dimensions of the sharp 25e55-deg biconic case [26] b): Numerical setups of the sharp 25e55-deg biconic case [13] c): Structured mesh of the sharp 25e55-deg biconic case, 256 256 (every fourth line presented) [13]
Table 1
Freestream conditions and working gas of the NACA0012 micro-airfoil [14]
Case p ∞ (Pa) M ∞ u ∞ (m/s) T ∞ (K) T w (K) Gas AOA (deg.) Kn
Airfoil 2.78443 2 509 161 290 Air 0, 10 0.026
Airfoil 0.72395 2 509 161 290 Air 10 0.1
Airfoil 0.27844 2 509 161 290 Air 20 0.26
Fig 3 Heat flux distribution along the biconic surface.
Fig 4 Heat flux distribution over the airfoil surface of the case M ¼ 2, AOA ¼ 0-deg,
Kn ¼ 0.026.
N.T.P Le et al / International Journal of Thermal Sciences 109 (2016) 334e341 338
Trang 6of 290 W/m2 at the airfoil tip and then reduce gradually to the
location x/c¼ 0.1 Past this location they oscillate along the surface
The heat transfer with the sliding friction obtains the peak value of
125 W/m2 and thereafter decreases gradually to the location x/
c¼ 0.1 It is a nearly constant finite value in 0.1 < x/c < 0.9 before
slightly increasing in 0.9 x/c 1, and is close to the DSMC data
Fig 6presents the calculated heat transfer distributions over the
upper surface The DSMC data and the heat transfer with sliding
friction reach the peak values of 395 W/m2and 50 W/m2,
respec-tively at the airfoil tip Thereafter both of them rapidly decrease to
the location x/c¼ 0.4, and are nearly constant finite values along
the surface There is a good agreement between them The heat
transfer without the sliding friction obtains a negative value of
110 W/m2at the airfoil tip It rapidly reduces until the location x/
c¼ 0.5, and then increases gradually along the surface There is a
difference between the heat transfer without the sliding friction
and DSMC data
5.2.3 M∞¼ 2, AOA ¼ 0-deg, Kn ¼ 0.1
Fig 7presents the computed heat transfer distributions of the
case Kn¼ 0.1 with AOA ¼ 10-deg The heat transfer without the
sliding friction obtains the negative value at the airfoil tip, and then
rapidly increases to reach the peak value of 126.26 W/m2at the
location x/c¼ 0.012 Past the peak value, it gradually decreases along the surface In the range 0.05< x/c 1, its values are the lowest values, and have a gap with the DSMC data[14] The DSMC data and heat transfer with the sliding friction reach peak values of
150 W/m2and 169 W/m2, respectively, at the airfoil tip Past these peak values, they gradually drop to the location x/c¼ 0.15, and are nearly constantfinite values along the surface for both of them, except the latter one increasing in the range 0.9< x/c 1.0 The heat transfer with the sliding friction is close to the DSMC data
Fig 8 shows the simulation results of heat transfer over the upper surface, and is similar the heat transfers over the lower surface The heat transfer without the sliding friction is negative at the airfoil tip whereas that with sliding friction and the DSMC data
[14]are positive The heat transfer without the sliding friction does not give good agreement with DSMC data That with the sliding friction and the DSMC data reach the peak values of 126 W/m2and
179 W/m2, respectively and are close together in 0< x/c < 0.92
5.2.4 M∞¼ 2, AOA ¼ 0-deg, Kn ¼ 0.26 The heat transfer distributions over the lower surface for the case with high Kn¼ 0.26 and AOA ¼ 20-deg are presented inFig 9 The DSMC data[14]have the peak value of 91 W/m2at the airfoil tip Past this peak value they gradually decrease to the location x/
Fig 5 Heat flux distribution over the airfoil lower surface of the case M ¼ 2, AOA ¼
10-deg, Kn ¼ 0.026.
Fig 6 Heat flux distribution over the airfoil upper surface of the case M ¼ 2,
¼ 10-deg, Kn ¼ 0.026.
Fig 7 Heat flux distribution over the airfoil lower surface of the case M ¼ 2, AOA ¼ 10-deg, Kn ¼ 0.1.
Fig 8 Heat flux distribution over the airfoil upper surface of the case M ¼ 2,
¼ 10-deg, Kn ¼ 0.1.
N.T.P Le et al / International Journal of Thermal Sciences 109 (2016) 334e341 339
Trang 7c¼ 0.1 and are nearly constant a finite value along the lower
sur-face At the airfoil tip the heat transfer with the sliding friction has a
positive value while that without the sliding friction is a negative
value They increase to reach peak values of 34 W/m2at location x/
c¼ 0.01 for the heat transfer with the sliding friction, and 36 W/m2
at x/c¼ 0.04 for the other one Past the peak values that without the
sliding friction gradually decreases along the surface The heat
transfer with the sliding friction generally gives good agreement
with the DSMC data while that without the sliding friction does not
Fig 10shows the simulation heat transfers over the upper
sur-face Both of the DSMC data [14]and the heat transfer with the
sliding friction reach the peak values of 75 W/m2and 30 W/m2,
respectively, and they are generally close together along the
sur-face The heat transfer without the sliding friction does not give
agreement with DSMC data and there is gradually increase in the
range 0.84< x/c < 0.96 before decreasing rapidly to the location x/
c¼ 1.0
5.3 Flat plate cases, M∞¼ 6.1
All calculated heat transfers of theflat plate cases are presented
inFigs 11 and 12 The dsmcFoam solver[24]was used for DSMC
simulations on the same test cases, using the variable hard sphere
model The normal gradient of temperature, the slip velocity, and shear stress play a part in determining the wall heat transfer For the hot-wall case (Tw¼ 294 K), near the leading edge, the wall heat transfer without the sliding friction is negative while that is posi-tive even close to the leading edge of the cold-wall case (Tw¼ 77 K) The effect of the sliding friction on the wall heat transfers is
sig-nificant near the leading edge for both cases due to the large slip velocity and shear stress, which results in their positive values They obtain the peak values of 12.13 kW/m2and 10.33 kW/m2for the cold-wall and hot-wall cases, respectively, at near the leading edge Thereafter they gradually decrease along theflat plate In 0.005 m x 0.02 m the wall heat transfers with the sliding friction of two cases are close to DSMC data, seen inFigs 11 and 12
In 0.02 m x 0.0875 m all DSMC and CFD simulation results are close together This can be explained that the slip velocity and shear stress obtain the peak values near the leading edge, and thereafter they rapidly decrease to the almost low constant values along the flat plate in 0.02 m x 0.0875 m So the sliding friction does not strongly affect the wall heat transfers The wall heat transfers of the cold-wall case are greater than those of the hot-wall case
5.4 Discussions The CFD wall heat transfers with the sliding friction over the
Fig 9 Heat flux distribution over the airfoil lower surface of the case M ¼ 2,
AOA ¼ 20-deg, Kn ¼ 0.26.
Fig 10 Heat flux distribution over the airfoil upper surface of the case M ¼ 2,
¼ 20-deg, Kn ¼ 0.26.
Fig 11 Heat flux distribution over the flat plate, T w ¼ 77 K.
Fig 12 Heat flux distribution over the flat plate, T w ¼ 294 K.
N.T.P Le et al / International Journal of Thermal Sciences 109 (2016) 334e341 340
Trang 8micro-airfoil and biconic surfaces of all simulation cases give good
agreements with DSMC data, even for the airfoil cases with high
Kn¼ 0.1 and 0.26, while those without the sliding friction do not
This can be explained by the different nature of the computation of
the heat transfers between DSMC (Eq.(1)) and the CFD method
without the sliding friction (i.e the Fourier law of conduction
q¼ kVnT) The contribution of the velocity terms in equation(1)
such as molecular velocity results in the heat transfer of DSMC to be
positive at near the tip of the airfoil, while those of CFD without the
sliding friction are negative due to the temperature gradient only
Once the slip velocity boundary applying to the surface, the gas
molecules will slip over the surface to generate more the heat
transfer that was computed based on the slip velocity and shear
stress terms
In the previous work[13,14] and our CFD simulations of the
NACA0012 micro-airfoil cases, the shear stresses reached the peak
values at the tip of the airfoil Past these values, they gradually
decrease along the airfoil surface Corresponding to the shear stress
distributions the slip velocities were slow due to viscous stress at
the tip airfoil, and gradually increase to trailing edge of the airfoil
Since the heat transfers over the surface including the sliding
friction, which are computed by the inner product of the shear
stresses and slip velocities, are greater than those without the
sliding friction This results in the values of heat transfers including
the sliding friction are positive at the tip of airfoil, and close to the
DSMC data along the micro-airfoil surface
For the biconic case the slip velocity and shear stress reached
the peak values at the tip of biconic and then gradually decrease in
x< 0.05 m The wall heat transfer involving the sliding friction
gives good agreement with DSMC data at the tip of biconic
(x< 0.05 m) Past this location, the heat transfer with and without
the sliding friction are close together because the slip velocity and
shear stress are small values in 0.05 m< x < 0.2 m presented in our
CFD simulations This results in the sliding friction term does not
significantly affect the heat transfer calculation in
0.05 m x 0.2 m
For theflat plate case the sliding friction strongly affects the wall
heat transfer near the leading edge due to the slip velocity and
shear stress By involving the sliding friction part, the CFD heat
transfers give good agreements with DSMC data for the cases
Kn> 0.1
6 Conclusions
The present work revisited the computations of the CFD heat
transfer with and without the sliding friction in high-speed rarefied
gasflow simulations The formulation of the heat transfer including
the sliding friction was re-evaluated for the flat plate case and
extended to the curved surfaces such as NACA0012 micro-airfoil
and the sharp 25e55-deg biconic From all simulation results
ob-tainedfigured out the importance of the sliding friction part in
calculating the wall heat transfer in CFD This represents the work
done directly over the surface by the shear and is nonzero only if
the slip velocity condition is applied to the surface The addition of
the sliding friction part in the computation of CFD heat transfers
leads to good agreement with DSMC for all cases considered,
especially for the cases with high Knudsen numbers, Kn¼ 0.1 and
0.26 (i.e in the transition regime)
Acknowledgements
The authors thank Dr Craig White at the University of Glasgow
for providing data from his DSMC simulations of theflat plate cases
This research is funded by Vietnam National Foundation for Science
and Technology Development (NAFOSTED) under grant number 107.03-2015.16
References [1] Bird GA The DSMC method, Clarendon, Oxford 2013 [2] Maslen SH On heat transfer in slip flow J Aerosp Sci 1958;25:400e1 [3] Van Rij J, Ameel T, Harman T The effect of viscous dissipation and rarefaction
on rectangular microchannel convective heat transfer Int J Therm Sci 2009;48:271e81
[4] Hadjiconstantinou NG Dissipation in small scale gaseous flows J Heat Transf 2003;125:944e7
[5] Hong C, Asako Y Some considerations on thermal boundary condition of slip flow Int J Heat Mass Trans 2010;53:3075e9
[6] Ramadan K The role of the shear work in microtube convective heat transfer:
a comparative study J Heat Transf 2016;138:011701 [7] Colin S Gas microflows in the slip flow regime: a critical review on convective heat transfer J Heat Transf 2011;134:020908
[8] Vidal RJ, Bartz JA Surface measurements on sharp flat plates and wedges in low-density hypersonic flow AIAA J 1969;7:1099e109
[9] Hong C, Asako Y Heat transfer characteristics of gaseous flows in micro-channel with negative heat flux Heat Transf Eng 2008;29:805e15 [10] Hadjiconstantinou NG The effect of viscous heat dissipation on convective heat transfer small-scale slipping gaseous flows In: First international con-ference on microchannels and minichannels, ICMM2003-1029; 2003 [11] Davis RT Numerical solution of the hypersonic viscous shock layer equations AIAA J 1970;8:843e51
[12] G€okçen T, MacCormack RW Nonequilibrium effects for hypersonic transi-tional flows using continuum approach In: AIAA paper 1989-0461; 1989 [13] Le NTP, Shoja Sani A, Roohi E Rarefied gas flow simulations of NACA 0012 airfoil and sharp 25-55-deg biconic subject to high order nonequilibrium boundary conditions in CFD J Aerosp Sci Technol 2015;41:274e88 [14] Shoja-Sani A, Roohi E, Kahrom M, Stefanov S Investigation of aerodynamic characteristics of rarefied flow around NACA 0012 airfoil using DSMC and NS solvers Eur J Mech B Fluids 2014;48:59e74
[15] Fan J, Boyd ID, Cai C Computation of rarefied gas flows around a NACA 0012 airfoil AIAA J 2001;39:618e25
[16] Holden MS, Wadhams TP, Candler GV, Harvey JK Measurements in regions of low density laminar shock wave/boundary layer interaction in hypervelocity flows and comparison with NaviereStokes predictions In: AIAA paper 2003-1131; 2003
[17] Greenshields CJ, Weller HG, Gasparini L, Reese JM Non-oscillatory central schemes for high speed viscous flows Int J Numer Methods Fluids 2010;23: 1e21
[18] Lofthouse A, Scalabrin LC, Boyd ID Velocity slip and temperature jump in hypersonic aerothermodynamics J Thermo Heat Transf 2008;22:38e49 [19] Scalabrin LC, Boyd ID Numerical simulation of weakly ionized hypersonic flow for reentry configurations In: AIAA paper 2006-3773; 2006
[20] Lengrand JC, Allegre J, Chpoun A, Raffin M Rarefied hypersonic flow over a sharp flat plate: numerical and experimental results In: Proceedings of the 18th international symposium on rarefied gas dynamics; 1992 p 276e84 [21] Tannehill JC, Mohling RA Numerical computation of hypersonic viscous flow over a sharp leading edge AIAA J 1974;12:129e30
[22] Roy CJ, Gallis MA, Bartel TJ, Payne JL Navierestokes and direct simulation Monte Carlo predictions for laminar hypersonic separation AIAA J 2003;41 [23] Rana AS, Mohammadzadeh A, Struchtrup H A numerical study of the heat transfer through a rarefied gas confined in a microcavity Contin Mech Thermodyn 2015;27:433e46
[24] OpenFOAM, http://www.openfoam.org , 12/2015.
[25] Metcalf SC, Lillicrap DC, Berry CJ A study of the effect of surface temperature
on the shock-layer development over sharp-edge shapes in low-Reynolds-number high-speed flow In: Proceedings of the seventh international sym-posium on rarefied gas dynamics; 1969 p 619e34
[26] Moss JN, Bird GA Direct simulation Monte Carlo simulations of hypersonic flows with shock interactions AIAA J 2005;43:2565e73
[27] Holden MS, Wadhams TP Code validation study of laminar shock/boundary layer and shock/shock interactions in hypersonic flow Part A: experimental measurements In: AIAA paper 2001e1031; 2001
[28] Liou WW, Fang Y Microfluid mechanics McGraw-Hill; 2006 [29] Patterson GN Molecular flow of gases Wiley; 1956 [30] Le NTP, Roohi E A new form of the second order temperature jump boundary condition in the low speed nano/microscale and hypersonic rarefied gas flow simulations Int J Therm Sci 2015;98:51e9
[31] Maxwell JC On stresses in rarefied gases arising from inequalities of tem-perature Phil Trans Roy Soc 1879;170:231e56
[32] von Smoluchowski M Überw€armeleitung in verdünntengasen Ann Phys Chem 1898;64:101e30
[33] Karniadakis G, Beskok A, Aluru N Microflows and nanoflows: fundamentals and simulation Springer; 2005
[34] Kennard EH Kinetic theory of gases McGraw-Hill; 1938 [35] Abbott IH, von Doenhoff AE Theory of wing sections, including a summary of airfoil data McGraw-Hill; 1949
N.T.P Le et al / International Journal of Thermal Sciences 109 (2016) 334e341 341