You are on page 1of 7

Rotation effects on non-Darcy convection in an enclosure lled with porous medium

H. Saleh
a
, A.Y.N. Alhashash
a
, I. Hashim
a, b,

a
School of Mathematical Sciences, Universiti Kebangsaan Malaysia, 43600 Bangi Selangor, Malaysia
b
Solar Energy Research Institute, Universiti Kebangsaan Malaysia, 43600 Bangi Selangor, Malaysia
a b s t r a c t a r t i c l e i n f o
Available online 8 February 2013
Keywords:
Natural convection
ForchheimerBrinkman model
Rotating enclosure
Natural convection heat transfer in a rotating, differentially heated enclosure is studied numerically in this
paper taking into consideration the ForchheimerBrinkman-extended Darcy model. The enclosure is lled
with a uid-saturated porous medium and executes a steady counterclockwise angular velocity about its
longitudinal axis. The staggered grid arrangement together with the Marker and Cell (MAC) method was
employed to solve the governing equations. The governing parameters considered are the porosity,
0.4 0.99, the Darcy number, 0.005Da0.01 and the Taylor number, 8.910
4
Ta3.810
5
, and the
centrifugal force is assumed weaker than the Coriolis force. It is found that higher porosities have weaker
ow circulation when the Coriolis effect is smaller than the buoyancy effect. The global quantity of the
heat transfer rate increases by increasing the porosity and the Darcy number and decreases by increasing
the Taylor number.
2013 Elsevier Ltd. All rights reserved.
1. Introduction
The uid ow and heat transfer characteristics of a rotating enclo-
sure were treated in detail according to its functional and practical
importance in the design of a guided missiles and a rotating electrical
machine. Consistently too high temperature occurring in the rotating
machine will reduce the lifetime of the machine and may lead to se-
rious failure, so that the heat sinks are needed. It is demonstrated
that the use of porous media as heat sinks is an effective way to re-
move heat [1]. Some theoretical and experimental studies can be
found in the literatures related to convection in enclosures that are
lled with a uid-saturated porous medium, but for a conguration
not closely similar to this work.
Areviewof the rotating enclosure problemlled with a viscous uid
will be given rst. [2] studied thermal cellular convection in a rotating
rectangular box and found that the roll cells changed orientation with
increasing Taylor number. The uid ow and heat transfer characteris-
tics of a rotating square enclosure were studied experimentally and
numerically by [3]. They concluded that the Coriolis force arising from
rotation may have a remarkable inuence on heat transfer when com-
pared with non-rotating results. [46] studied a differentially heated
rotating cubic enclosure. Signicant ow modication was obtained
when the rotational Rayleigh number is greater than the Rayleigh num-
ber. A signicant increasing or decreasing in heat transfer could be
achieved due to rotational effects as reported by [7,8]. [9] studied
numerically a rectangular enclosure with discrete heat sources and
found that rotation results in imbalance of clockwise and counter-
clockwise circulations. The effects of Coriolis force, centrifugal force
and thermal buoyancy force were segregated numerically by [10]
on a differentially heated square enclosure.
Many analyses of a differentially vertical or horizontal heated
rectangular enclosure that is lled with a porous medium are avail-
able in the literature. Here, we will restrict our discussion to the
ForchheimerBrinkman-extended Darcy model only. [11] showed
how the Darcy law could be extended in a buoyancy-driven ow to
include the inertia effects. [12] reported that the applicability of
Darcy formulation was highly restricted to the case of the Bnard con-
vection problem. [13] investigated the effects of quadratic inertia and
the viscous terms. [14] studied the case of a variable porosity and
found that the thickness of the porous layer and the nature of varia-
tion in porosity signicantly affect the natural convective ow pat-
tern as well as the heat transfer features. For the same model, [15]
showed that the progress of ow evolution is retarded by the lower
porosity and permeability of the medium, but the heat transfer rate
strongly depends on the Bnard convection ow pattern. Recently,
[16,17] studied a porous enclosure bounded by a conductive wall.
[18] studied an inclined rectangular enclosure with variable concen-
tration and reported that the heat transfer rate decreased when the
angle of inclination is increased.
Inthe present study, the problemof natural convectionheat transfer
in a rotating, differentially heated porous enclosure is studied numeri-
cally, considering the ForchheimerBrinkman-extended Darcy model.
The effects of porosity and permeability of the porous medium as well
as the rotational speeds on characteristics of uid ow as well as heat
transfer rate are considered.
International Communications in Heat and Mass Transfer 43 (2013) 105111
Communicated by A.R. Balakrishnan and T. Basak.
Corresponding author at: School of Mathematical Sciences, Universiti Kebangsaan
Malaysia, 43600 Bangi Selangor, Malaysia.
E-mail address: ishak_h@ukm.my (I. Hashim).
0735-1933/$ see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.icheatmasstransfer.2013.01.006
Contents lists available at SciVerse ScienceDirect
International Communications in Heat and Mass Transfer
j our nal homepage: www. el sevi er . com/ l ocat e/ i chmt
2. Mathematical formulation
A schematic diagramof a square porous enclosure of sidel execut-
ing a steady uniform counterclockwise angular velocity about its
longitudinal is shown in Fig. 1, with the geometric layout and the
Cartesian coordinates (x,y) rotating with the enclosure. The surface
at y l=2 has the constant hot temperature (T
h
) and the surface at
y l=2 has a constant cold temperature (T
c
). The temperatures
along the lateral wall are assumed to be linearly distributed between
T
h
and T
c
, i.e. T
h
T
c
=2 T
h
T
c
y=l to consider conjugate heat
transfer in the lateral wall. The shown in Fig. 1 is dened as an
angular position.
The uid is Newtonian and the ow is laminar and incompress-
ible. The density variation of the uid follows the Boussinesq approx-
imation and changes with temperature only. The porous medium is
considered homogenous, isotropic and is saturated with a uid that
is in local thermodynamic equilibrium with the solid matrix of the
porous medium. The terms representing the thermal buoyancy,
rotational buoyancy and Coriolis force are, respectively, equal to

f
g(TT
c
), o
f
g(TT
c
)(r), and 2
f
[1(TT
c
)] V.
The continuity, momentum and energy equations are:
u
x

v
y
0
1

u
t

2
x

v

2
y

1

f
P
M
x

2
u
x
2

2
u
y
2
_ _

K
F
f

u
2
v
2
_

K
p u 2v
2v TT
c

2
x TT
c

1
g TT
c
sin t
1

v
t

u

2
v
x

v

2
v
y

1

f
P
M
y

2
u
x
2

2
u
y
2
_ _

K
vF
f

u
2
v
2
_

K
p v 2u
2u TT
c

2
y TT
c

2
g TT
c
cos t 3
Cp
m
T
t
Cp
f
u
T
x
v
T
y
_ _
k
m

2
T
x
2

2
T
y
2
_ _
4
where P
M
is the motion pressure dened as:
P
M
p
1
2

2
x
2

1
2

2
y
2

f
gxsin t
f
gycos t : 5
Hence,

P
M
x

p
x

f

2
x
f
gsin t 6

P
M
y

p
y

2
y
f
gcos t : 7
Fig. 1. Schematic representation of the model.
Nomenclature
Da Darcy number
F inertia coefcient
Cp specic heat capacity (J/(kg K))
g gravitational acceleration (m/s
2
)
k thermal conductivity (W/(m K))
K permeability (m
2
)
l width and height of enclosure (m)
Nu Nusselt number
p & P pressure (N/m
2
) & dimensionless pressure
Pr Prandtl number
Ra Rayleigh number
Ra

rotational Rayleigh number


t time (s)
Ta Taylor number
T temperature (K)
u, v velocity components in the x- and y-directions (m/s)
x, y & X, Y space coordinates (m) & dimensionless space
coordinates
Greek symbols
thermal diffusivity (m
2
/s)
porosity
kinematic viscosity (m
2
/s)
dimensionless time

p
dimensionless time for one rotation
dimensionless temperature
magnitude of angular rotation rate (rpm, rad/s)
density (kg/m
3
)
angular position of enclosure (rad)
dynamic viscosity (N/s/m
2
)
Subscript
c cold
f uid
h hot
m uid-saturated porous medium
s solid
Superscript
- average
= global quantity
106 H. Saleh et al. / International Communications in Heat and Mass Transfer 43 (2013) 105111
The governing Eqs. (1)(4) can be converted to nondimensional
forms using the following nondimensional parameters:
X
x

; Y
y

;
t

2
; U
u

; V
v

;
TT
c
T
h
T
c
;
Pr
v

; Da
K

2
; Ra
g T
h
T
c

3

; P
P
M

2
; Ta
4
2

2
Ra



2
T
h
T
c

4

;
Cp
f
1 Cp
s
_ _
Cp
f
:
8
The Coriolis buoyancy force is neglected since |(TT
c
)| 1 in
the present study. The nondimensional continuity, momentum, and
energy equations are written as:
U
X

V
Y
0
1

2
U
X

V

2
U

2

P
X

Pr

2
U
X
2

2
U
Y
2
_ _
9

PrU
Da
F

U
2
V
2
_

Da
p U Ta
0:5
PrV
..
Coriolis force term
Ra

PrX
..
Centrifugal force term
RaPrsin 0:5Ta
0:5
Pr
_ _
..
Buoyancy force term
10
1

2
V
X

V

2
V
Y

P
Y

Pr

2
V
X
2

2
V
Y
2
_ _

PrV
Da
F

U
2
V
2
_

Da
p U Ta
0:5
PrU
..
Coriolis force term
Ra

PrY
..
Centrifugal force term
RaPrcos 0:5Ta
0:5
Pr
_ _
..
Buoyancy force term
11

X
V

Y


2

X
2

Y
2
: 12
We assume U=V=0 on the walls and the boundary conditions
for the non-dimensional temperatures are:
at X 0:5 and X 0:5 : Y 0:5; 0:5Y0:5 13
at Y 0:5 : 1 14
at Y 0:5 : 1: 15
The non-dimensional governing equations contain three indepen-
dent dimensionless parameters, namely Prandtl number, Rayleigh
number and Taylor number. In addition, rotational Rayleigh number
reects the effect of the centrifugal buoyancy force, but it depends
on the other dimensionless parameters too [10].
The uid motion is displayed using the stream function obtained
fromthe velocity components u and v. The physical quantities of interest
in this problemare the average Nusselt number, Nu
0:5
0:5
racY dX,
signifying the overall heat transfer performance on the heated or cooled
plates, and the global quantity of the average heat transfer rate over
one cycle, Nu
p
0
Nu d.
3. Numerical method and validation
Staggered grid arrangement together with the Marker and Cell
(MAC) method [19] is adopted to solve the governing Eqs. (9)(12)
subject to the boundary conditions (13)(15). Due to lack of
boundary conditions for pressure, the use of the staggered grid
and MAC formulation provides an advantage. That is, one may
locate the secondary grid along the boundaries of the domain
where only specication of velocity boundary conditions is re-
quired but not of the pressure. The ctitious values of velocity
outside the domain are obtained by extrapolation of the interior
points as given by [20]. A second-order central difference approxi-
mation is used for the space discretization and a rst-order
approximation is used for the temporal derivative. The solution of
the Poisson pressure equation is obtained by applying an iterative
20 40 60 80 100 120 140
2
2.05
2.1
2.15
2.2
2.25
2.3
2.35
2.4
2.45
2.5
Grid size
N
u
0 1 2 3 4 5 6 7 8 9 10
1
1.5
2
2.5
3
Number of rotations
N
u
(a) (b)
Fig. 2. (a) Grid independency study: Nu versus grid size inthe X- and Y- directions for =0.4, Da=0.01, Ta=8.9710
4
, Ra=1.210
5
at =5/4 and (b) typical computation process for
the present periodic oscillation problem.
107 H. Saleh et al. / International Communications in Heat and Mass Transfer 43 (2013) 105111
Gaussian-SOR method. The velocities are then computed by the
projection method.
In this study, the convergence criterion for the Poisson equation is
set at =10
5
and the chosen time stepping is =(1/4)(X)
2
Pr to
meet stability criteria, where is the safety factor with a value
between 0 and 1. Uniform grid distribution is used for the whole en-
closure. The effect of grid resolution was examined in order to select
the appropriate grid density as demonstrated in Fig. 2(a) for =0.4,
Da=0.01, Ta=8.9710
4
, Ra=1.210
5
at =5/4. The results indi-
cate that a 120120 mesh can be used in the nal computations. A
typical computation process for the present periodic oscillation prob-
lemis shown in Fig. 2(b), where after three rotations, the nal period-
ic oscillation of the Nu is obtained. Therefore, all computations in this
work were carried out beyond two rotations. As a validation, our re-
sults for the isotherms compare well with that obtained by [3,10]
for a pure uid taking =0.99, Da=10
7
, Pr =0.7, Ra=1.210
5
,
Ta=8.910
4
or =8.5 rpm as shown in Fig. 3.
4. Results and discussion
The analysis in the undergoing numerical investigation is performed
in the following ranges of the associated dimensionless groups: the
porosity of the porous medium, 0.4 0.99, the Darcy number,
0.005Da0.01 and the Taylor number, 8.910
4
Ta3.810
5
.
Other parameters are T
h
=282 (K), T
c
=273 (K), l 0:0508 (m),
Pr =0.7, Ra=1.210
5
, =8.5 to 17.5 (rpm) and Ra

=496 to 2104.
The rotational Rayleigh number, Ra

, was not discussed except to specify


it explicitly. This is due to the Ra

not being an independent parameter


but depending on Pr and Ra as well as Ta. We note that in the present re-
search the Coriolis force is stronger than the centrifugal force.
Fig. 4(a)(h) show the effects of rotation on the ow eld for
different porosities at Da=0.01 and Ta=8.910
4
. There are two
basic ow patterns, clockwise (negative sign) and counterclockwise
(positive sign) circulations, depending on the heated position as
well as the inertial forces. Starting from =/4 where the heated
(a)
(b)
(c)
Fig. 3. Comparison of computed isotherms of present work (left) with [10] (middle) and [3] (right) results for a pure uid, =0.99, Da=10
7
, Pr=0.7, Ra=1.210
5
, Ta=8.910
4
or Omega=8.5 rpm.
108 H. Saleh et al. / International Communications in Heat and Mass Transfer 43 (2013) 105111
(a)
(b) (h)
(c) (g)
(d) (f)
(e)
Fig. 4. Streamline [=0.4 (solid lines) and =0.45 (dashed lines)] evolution during one period for Da=0.01 and Ta=8.910
4
when (a) =0, (b) =/4, (c) =/2, (d) =3/4,
(e) =, (f) =5/4, (g) =3/2 and (h) =7/4.
109 H. Saleh et al. / International Communications in Heat and Mass Transfer 43 (2013) 105111
wall is above the cold wall, one observes a unicellular clockwise
motion induced by the buoyancy thermal force. This ow circulation
persists up to =, after the gravitational force reverses the ow
direction. At =5/4, subplot (f), the effect of buoyancy force now
is to induce a counterclockwise circulation, but the ow inertia from
the clockwise circulation still exists. The Coriolis force helps the
clockwise motion, yielding a clockwise ow in the center and two
smaller vortexes in the corners rotating in a counterclockwise, where
here the effect of the Coriolis force is small. With increasing angle the
gravitational force increases so that the size of the two vortexes grows
bigger as shown in the subplot (g). The effect of the gravitational force
neutralizes the effect of the inertial force of the clockwise motion. It is
interesting to note that the size of the two vortexes is much bigger for
higher porosities at =5/4, but the size of the two vortexes is much
smaller for higher porosities at =3/2, see subplot (g). These phe-
nomena indicate that higher porosities have weaker ow circulation
when the Coriolis effect is smaller than the buoyancy effect.
At =7/4 as shown in Fig. 4(h), the transition from clockwise to
counterclockwise motion is completed. The relatively weak motion is
observed in the center of the enclosure. It is interesting to note that
the Coriolis force now is directed outward from the core and tends
to stabilize the ow circulation. Finally at =2 or 0 rad, subplot
(a), the ow circulation is very weak, and two cell structures appear
in the core. Note that this sequence repeats itself, the fully clockwise
ow then transition ow and totally counterclockwise ow. It is ob-
served that during rotation, the strength of the ow circulation of
higher porosities is always stronger than lower porosities. Also, note
that the ow patterns between the =0.4 and the =0.45 were differ-
ent enough in some angular locations. The streamlines in all of the
subplots (a)(h) present centrosymmetric shape.
Fig. 5(a) and (b) depicts the variations of the average Nusselt num-
ber, Nu, over one cycle for different rotational speeds (Taylor numbers)
at =0.4 and Da=0.01 and different porosities at Ta=8.910
4
, respec-
tively. We can see that the average Nusselt number proles display
0 pi/4 pi/2 3pi/4 pi 5pi/4 3pi/2 7pi/4 2pi
1.2
1.4
1.6
1.8
2
2.2
2.4
N
u
Ta

10
4
Ta 10
5
Ta 10
5
Ta 10
5

0 pi/4 pi/2 3pi/4 pi 5pi/4 3pi/2 7pi/4 2pi
1.5
2
2.5
3
3.5
4
N
u
= 0.4
= 0.6
= 0.8
= 0.99
(a) (b)
= 8.9
= 1.3
= 2.2
= 3.8
Fig. 5. Plots of the average Nusselt number against for the different values of (a) Ta and (b) at Da=0.01.
0.4 0.5 0.6 0.7 0.8 0.9
1.2
1.4
1.6
1.8
2
2.2
2.4
2.6
N
u
Ta
Ta
Ta
Ta
0.4 0.5 0.6 0.7 0.8 0.9
1.6
1.8
2
2.2
2.4
2.6
2.8
N
u
Da=0 .005
Da=0 .01
Da=0 .05
Da=0 .1

10
4
10
5
10
5
10
5
= 8.9
= 1.3
= 2.2
= 3.8
(a) (b)
Fig. 6. Plots of the global quantity average Nusselt number against for the different values of (a) Ta and (b) Da.
110 H. Saleh et al. / International Communications in Heat and Mass Transfer 43 (2013) 105111
roughly similar characteristics for the various Taylor numbers as shown
in Fig. 5(a). For each value of the Taylor number, there exists only one
local maximum and minimum heat transfer rates. The angular location
of the local maximum heat transfer varies at different rotational speeds.
The local maximum point tends to move slightly to higher angle when
the enclosure is spun faster. We can conclude that increasing the rota-
tional speed stabilizes the heat transfer rate. It is observed in Fig. 5(b)
that higher porosity leads to stronger Nu except near =/4. At this
angular location, the heat transfer rate is stagnant by varying the poros-
ities and displays the minimum heat transfer. Fig. 5(b) also shows that
the angular location of the local maximum heat transfer varies at differ-
ent rotational speeds. It tends to move to higher angle when the porosi-
ties made larger. Decreasing the porous medium porosities can stabilize
the heat transfer oscillations.
Fig. 6(a) and (b) shows the global quantity average Nusselt num-
ber against the porosity for different Taylor numbers at Da=0.01 and
different Darcy numbers at Ta=8.910
4
, respectively. The heat
transfer performance increases by increasing the porosity as shown
in Fig. 6(a). This is expected because low porosity means more resis-
tance to uid ow. As a result of this phenomenon, for larger values of
porosity, isotherms are denser near the hot wall, which increase the
temperature gradient in this region. It is also observed that higher ro-
tational speeds lead to lower heat transfer rate at the considered
range. This result is consistent with the previous outcome. The results
demonstrated in Fig. 6(b) help compare the heat transfer appearance
by varying the permeability of the porous matrix. The lowest heat
transfer rate is obtained for the case Da=0.005 since the porous
matrix has a lower value of permeability that leads to lower uid
velocity which suppresses the heat transfer performance. Each Nu
plot displays similar trend. This indicates globally that the heat trans-
fer rate increases almost linearly by increasing the Darcy number.
5. Conclusions
Detailed computational results for ow elds and the heat transfer
performance of the rotating porous enclosure have been presented in
graphical forms. The periodic oscillation of the uid ow and tempera-
ture elds as well as the heat transfer was obtained. The main conclu-
sions of the present analysis are as follows:
1. A unicellular motion is obtained in the rotating enclosure at /4 rad
up to rad and a multicellular motion is obtained from 5/4 until
2. Higher porosities have weaker owcirculation when the Coriolis
effect is smaller than the buoyancy effect and the owstructure pre-
sents centrosymmetric shape during rotation.
2. Decreasing the porous medium porosity and/or increasing the ro-
tational speed enables the stabilization of the uid ow and heat
transfer oscillation.
3. The angular locations of the local maximumandminimumheat trans-
fer are sensitive to rotational speeds and porous medium porosity.
The local maximum or minimum point tends to move to higher
angle when the enclosure is spun faster and it moves to lower angle
when the porosity is made larger.
4. The global quantity of the heat transfer rate increases by increasing
the porosity and permeability and decreases by increasing the rota-
tional speeds.
Acknowledgments
This work was supported by the Universiti Kebangsaan Malaysia,
grant no. DIP-2012-61. The authors wish to thank the reviewers for
the constructive comments that led to an improvement in the paper.
References
[1] H.Y. Li, K.C. Leong, L.W. Jin, J.C. Chai, Transient two-phase ow and heat transfer
with localized heating in porous media, International Journal of Thermal Sciences
49 (2010) 11151127.
[2] K. Buhler, H. Oertel, Thermal cellular convection in rotating rectangular boxes,
Journal of Fluid Mechanics 114 (1982) 261282.
[3] F.J. Hamady, J.R. Lloyd, K.T. Yang, H.Q. Yang, A study of natural convection in a
rotating enclosure, Journal of Heat Transfer 116 (1994) 136143.
[4] T.L. Lee, T.F. Lin, Transient three-dimensional convection of air in a differentially
heated rotating cubic cavity, International Journal of Heat and Mass Transfer 39
(1996) 12431255.
[5] Y.T. Ker, T.F. Lin, A combined numerical and experimental study of air convection
in a differentially heated rotating cubic cavity, International Journal of Heat and
Mass Transfer 39 (1996) 31933210.
[6] Y.T. Ker, T.F. Lin, Time-averaged and reverse transitioninoscillatory air convectionin
a differentially heated rotating cubic cavity, International Journal of Heat and Mass
Transfer 40 (1997) 33353349.
[7] M.F. Baig, A. Masood, Natural convection in a two-dimensional differentially heated
square enclosure undergoing rotation, Numerical Heat Transfer Part A 40 (2001)
181202.
[8] M.F. Baig, M. Zunaid, Numerical simulation of liquid metals in differentially heated
enclosure undergoing orthogonal rotation, International Journal of Heat and Mass
Transfer 49 (2006) 35003513.
[9] L.F. Jin, S.K.W. Tou, C.P. Tso, Effects of rotation on natural convection cooling from
three rows of heat sources in a rectangular cavity, International Journal of Heat
and Mass Transfer 48 (2005) 39823994.
[10] C.P. Tso, L.F. Jin, S.K.W. Tou, Numerical segregation of the effects of body forces in
a rotating, differentially heated enclosure, Numerical Heat Transfer Part A 51
(2007) 85107.
[11] J.G. Georgiadis, I. Catton, Prandtl number effect on Bnard convection in porous
media, Journal of Heat Transfer 108 (1986) 284290.
[12] N. Kladias, V. Prasad, Flow transitions in buoyancy-induced non-Darcy convection
in a porous-medium heated from below, Journal of Heat Transfer 112 (1990)
675684.
[13] J.L. Lage, Effect of the convective inertia term on Bnard convection in a porous-
medium, Numerical Heat Transfer Part A 22 (1992) 469485.
[14] P. Nithiarasu, K.N. Seetharamu, T. Sundararajan Natural, Convective heat transfer
in an enclosure lled with uid saturated variable porosity medium, International
Journal of Heat and Mass Transfer 40 (1997) 39553967.
[15] T.C. Jue, Analysis of Bnard convection in rectangular cavities lled with a porous
medium, Acta Mechanica 146 (2001) 2129.
[16] A. Al-Amiri, K. Khanafer, I. Pop, Steady-state conjugate natural convection in a
uid-saturated porous cavity, International Journal of Heat and Mass Transfer 51
(2008) 42604275.
[17] K. Al-Farhany, A. Turan, Non-Darcy effects on conjugate double-diffusive natural
convection in a variable porous layer sandwiched by nite thickness walls, Interna-
tional Journal of Heat and Mass Transfer 54 (2011) 28682879.
[18] K. Al-Farhany, A. Turan, Numerical study of double diffusive natural convective heat
and mass transfer inaninclinedrectangular cavity lledwithporous medium, Inter-
national Communication of Heat Mass Transfer 39 (2012) 174181.
[19] F. Harlow, J.E. Welch, Numerical calculation of time-dependent viscous incompress-
ible ow of uid with a free surface, Physics of Fluids 8 (1965) 21822189.
[20] K.A. Hoffmann, S.T. Chiang, Computational Fluid Dynamics, vol. I, Engineering
Education System, Kansas, 2000.
111 H. Saleh et al. / International Communications in Heat and Mass Transfer 43 (2013) 105111

You might also like