You are on page 1of 16

c Pleiades Publishing, Ltd., 2014.

ISSN 0021-8944, Journal of Applied Mechanics and Technical Physics, 2014, Vol. 55, No. 4, pp. 634–649. 
c V. Sivakumar, S. Sivasankaran.
Original Russian Text 

MIXED CONVECTION IN AN INCLINED LID-DRIVEN CAVITY


WITH NON-UNIFORM HEATING ON BOTH SIDEWALLS

V. Sivakumara and S. Sivasankaranb UDC 536.2; 536.25

Abstract: The present work reports a numerical simulation of mixed convection in an inclined square
cavity. The vertical sidewalls are assumed to have a nonuniform temperature distribution. The finite
volume method is used to solve dimensionless governing equations. Simulations are performed for
different Richardson numbers, amplitude ratios, phase deviations, and cavity inclination angles. The
results are presented graphically. The mean heat transfer significantly increases in the buoyancy-
dominated mode on increasing cavity inclination angle if both walls have identical heating and cooling
zones.
Keywords: mixed convection, lid-driven cavity, nonuniform heating, finite volume method, inclined
cavity.
DOI: 10.1134/S0021894414040105

INTRODUCTION

Mixed convection heat transfer in lid-driven cavities is often encountered in many engineering and industrial
applications, such as cooling of electronic systems, nuclear reactors, chemical processing equipment, lubricating
grooves, coating and industrial processes, float glass manufacturing, and solidification processes [1, 2]. The flow
and convection processes are influenced by two driving forces. One is the buoyancy force due to the temperature
gradient, and the other is the shear force due to lid movement. Mixed convection occurs if either the effect of
buoyancy dominates over the forced flow or the effect of the forced flow is significant as compared to the buoyant
flow. Mixed convection in a shallow square lid-driven cavity with a stable vertical temperature gradient was
studied by Mohammad and Viskanta [3]. They reported that multicellular flow regimes and a transition to a three-
dimensional flow occur in the cavity. Chamkha [4] made an analysis for mixed convection in a square cavity in the
presence of a magnetic field and internal heat generation or absorption. The result showed that the average Nusselt
number for the aiding flow increases on increasing Grashof number with a fixed value of the Reynolds number.
Seddeek [5] studied the effect of a magnetic field and variable viscosity on a forced non-Darcy flow past a flat plate
with a variable wall temperature in porous media in the presence of suction and blowing. He observed that the
magnetic field increases the wall skin friction, while the heat transfer rate decreases.
Jeng and Tzeng [6] investigated numerically the heat transfer in a lid-driven cavity filled with water-saturated
aluminum foam. Their result indicated that the local heat transfer rate increases with the Reynolds number for
a constant Grashof number. Numerical computations of two- and four-sided lid-driven cavity flows were carried
out by Perumal and Dass [7]. The effect of heating location and size on mixed convection in a lid-driven cavity
was investigated by Sivakumar et al. [8]. They observed that the heat transfer rate is enhanced on reducing the
heating portion length in the hot wall. The effect of the nonmonotonic temperature dependence of water density on

a
Department of Mathematics, Knowledge Institute of Technology, Salem 637 504, India. b Institute of Math-
ematical Sciences, University of Malaya, Kuala Lumpur 50603, Malaysia; sd.siva@yahoo.com. Translated from
Prikladnaya Mekhanika i Tekhnicheskaya Fizika, Vol. 55, No. 4, pp. 97–114, July–August, 2014. Original article
submitted November 19, 2012.
634 0021-8944/14/5504-0634 
c 2014 by Pleiades Publishing, Ltd.
free convection from a linear heat source was examined by Bukreev et al. [9]. They revealed that the height of the
convective plume is limited because water in the plume head reaches the maximum density and becomes heavier
than the surrounding water. A numerical study on natural convection in a rectangular porous enclosure was carried
out with five different locations of heating and cooling zones by Bhuvaneswari et al. [10]. They found that the heat
transfer rate is increased by increasing the Grashof number. Thandapani et al. [11] performed an analysis to study
the heat transfer characteristics of natural convection over a vertical cone under combined effects of a magnetic field
and thermal radiation. They found that an increase in the strength of the applied magnetic field decelerates the
fluid motion along the wall of the cone inside the boundary layer and increases the temperature of the boundary
layer.
Having either one sidewall or two sidewalls with sinusoidal temperature profiles in rectangular enclosures
has received considerable attention over the last several years due to the wide range of engineering applications.
Abourida et al. [12] studied natural convection in a square cavity with a periodic temperature distribution on its
horizontal walls. The temperature of the bottom surface was varied sinusoidally with time, while that of the top
surface was maintained constant or varied sinusoidally. They concluded that the heat transfer rate can be enhanced
or reduced with respect to the case of constant temperatures at high Rayleigh numbers. Saeid and Yaacob [13]
studied natural convection in a square cavity with a sinusoidally heated vertical wall. They found that the average
Nusselt number varies sinusoidally on increasing the wave number, and it increases on increasing the amplitude.
Numerical simulation on natural convection in a rectangular enclosure with a sinusoidal temperature profile on one
sidewall was carried out by Bilgen and Yedder [14]. They determined that the thermal penetration approaches
100% if the lower part is heated and the upper part is cooled, whereas it is limited to 70% if the upper half is heated
and the lower half is cooled for high Rayleigh numbers.
Deng and Chang [15] investigated natural convection in a rectangular enclosure with sinusoidal temperature
distributions on both side walls. Their results showed that natural convection in enclosures with two sinusoidal
temperature distributions on the sidewalls is superior to that with a single sinusoidal temperature profile on one side
wall. Al-Zoubi and Brenner [16] studied a laminar flow of a Newtonian fluid pressurized through a two-dimensional
channel with one sinusoidal wall. They observed that the wave length, the wave amplitude, and the Reynolds
number affect the flow fields. A numerical study on mixed convection in a lid-driven cavity with a sinusoidal
temperature distribution on vertical walls was performed by Sivasankaran et al. [17]. Their result showed that
nonuniform heating on both walls provides a higher heat transfer rate than nonuniform heating of one wall. Nath
et al. [18] investigated a combined effect of convective heat and mass transfer on a hydromagnetic electrically
conducting viscous incompressible fluid through a porous medium in a nonuniformly heated vertical channel.
Sivasankaran et al. [19] proposed a numerical approach on mixed convection in a lid-driven cavity with
sinusoidal boundary temperatures at the sidewalls in the presence of a magnetic field. Their investigation revealed
that the heat transfer rate increases with the phase deviation from zero to π/2, and then it decreases with a further
increase in the phase deviation. Bhuvaneswari et al. [20] conducted a numerical study on magneto-convection in a
square cavity with a sinusoidal temperature distribution on both walls. They concluded that the heat transfer rate
increases with increasing amplitude ratio. The effects due to the inclination of the cavity play an important role
in cooling of electronic equipment, especially inside laptop computers. These types of computers work at different
inclination angles on humans’ hands and in different environmental conditions. Natural convection in an inclined
enclosure was numerically studied by Rasoul and Prinos [21]. They observed that the mean heat transfer rate
increases on increasing Raleigh number for all inclination angles. Cianfrini et al. [22] numerically studied natural
convection in tilted square enclosures with two adjacent walls heated and two opposite walls cooled. They revealed
that heat transfer across the enclosure occurs as pure conduction for an inclination angle of 225◦ . A numerical
investigation on free convection heat transfer was carried out in a partially open inclined square cavity by Bilgen
and Oztop [23]. They concluded based on the results that the relationship between the heat transfer rate and the
inclination angle is nonlinear.
Sharif [24] studied combined convection in a shallow cavity for various Richardson numbers and inclination
angles ranging within 0–30◦ . He found that the average Nusselt number increases with increasing cavity inclination.
A laminar double-diffusive natural convective flow of a binary fluid mixture in inclined porous cavities in the presence
of temperature-difference-dependent heat generation (source) or absorption (sink) was studied by Chamkha and Al-
Mudhaf [25]. They found that the heat and mass transfer and the flow characteristics inside the enclosure depend

635
y U0

T(y) = T0 +Ar sin(2py + f)


g L

L
2py
T(y) = T0 +Al sin( )
L
x

Fig. 1. Physical configuration and coordinate system.

strongly on the buoyancy ratio, cavity inclination angle, and heat generation or absorption effects. Bairi [26]
experimentally and numerically investigated natural convection in a tilted square cavity. He found that convection
heat transfer is moderate due to fluid stratification affecting most of the cavity if the hot wall is at the top.
Varol et al. [27] performed a numerical study on natural convection in an inclined square enclosure with a corner
heater. They concluded that heat transfer is maximum for an inclination angle of around 135◦ and minimum at
60◦ . Ghasemi and Aminossadati [28] studied natural convection heat transfer in an inclined enclosure filled with
water–CuO nanofluid. They found that the heat transfer rate is maximized at a specific inclination angle depending
on the Rayleigh number and the solid volume fraction.
In most of the studies found in the literature, the main attention has been given in detail to problems
of a convective flow in vertical cavities with uniform or nonuniform varying thermal boundaries and in inclined
cavities with isothermal boundaries. To the best knowledge of the authors, the problem of mixed convection in an
inclined cavity subjected to nonuniform thermal boundaries has not yet been dealt in the literature. Hence, the
intention of the present study is to investigate the effects of sinusoidal boundary conditions at both sidewalls on
mixed convection in an inclined square cavity.

1. MATHEMATICAL FORMULATION

The physical model of the cavity is schematically shown in Fig. 1. The system is considered to be an unsteady
laminar incompressible mixed convective flow and heat transfer in a two-dimensional square cavity of length L filled
with air. The top horizontal wall (lid) of the cavity moves rightward with a constant speed U0 . Sinusoidally varying
temperature distributions are imposed on the left and right vertical walls of the cavity, respectively. Two horizontal
walls of the cavity are insulated. The cavity inclination angle γ varies from 0 to 90◦ . The fluid is treated as
incompressible. The thermophysical properties of the fluid are assumed to be constant, except for density. The
density varies linearly with temperature as ρ = ρ0 [1 − β(T − T0 )] (β is the coefficient of thermal expansion and the
subscript 0 denotes the reference state). The Boussinesq approximation is valid. It is further assumed that viscous
dissipation is neglected in this study. From the above-mentioned assumptions, the unsteady governing equations
for conservation of mass, momentum, and energy can be written as follows:
∂u ∂v
+ = 0,
∂x ∂y

∂u ∂u ∂u 1 ∂p  ∂ 2 u ∂ 2u 
+u +v =− +ν + 2 + gβ(T − T0 ) sin γ,
∂t ∂x ∂y ρ0 ∂x ∂x2 ∂y

∂v ∂v ∂v 1 ∂p  ∂2v ∂2v 
(1)
+u +v =− +ν 2
+ 2 + gβ(T − T0 ) cos γ,
∂t ∂x ∂y ρ0 ∂y ∂x ∂y
636
∂T ∂T ∂T k  ∂2T ∂2T 
+u +v = + .
∂t ∂x ∂y ρ0 cp ∂x2 ∂y 2
Here u and v [m/s] are the projections of the velocity vector onto the x and y axes, respectively, t [s] is the time,
ν = μ/ρ0 [m/s2 ] is the kinematic viscosity, μ [Pa · s] is the dynamic viscosity, ρ [kg/m3 ] is the density, g [m/s2 ]
is the gravitational acceleration, T [K] is the temperature, p [Pa] is the pressure, k [W/(m · K)] is the thermal
conductivity, and cp [J/(kg · K)] is the specific heat; the subscript 0 denotes the reference state.
The initial and boundary conditions of the problem can be written as follows:
at t = 0,
u = v = 0, T = 0, 0  x  L, 0  y  L;
at t > 0, ∂T
u = v = 0, = 0, y = 0,
∂Y
∂T
u = U0 , v = 0, = 0, y = L,
∂Y
u = v = 0, T (y) = T0 + Al sin (2πy/L), x = 0,

u = v = 0, T (y) = T0 + Ar sin (2πy/L + ϕ), x = L.


Here T (y) is the temperature of the vertical side walls, Al and Ar are the amplitudes of the sinusoidal temperature
of the left and right walls of the cavity, respectively. The phase deviations of the sinusoidal temperature of the left
and right walls are 0 and ϕ, respectively.
In order to reduce the number of equations and render them in dimensionless form, we introduce the stream
function ψ and vorticity Ω and apply the following transformations:
x y T − T0 tU0 ΩL ψ p
X= , Y = , θ= , τ= , ζ= , Ψ= , P = .
L L ΔT L U0 LU0 ρU02
Introducing these dimensionless variables into the governing equations (1), we obtain
∂ζ ∂ζ ∂ζ 1  ∂2ζ ∂2ζ   ∂θ ∂θ 
+U +V = + + Ri cos γ − sin γ ; (2)
∂τ ∂X ∂Y Re ∂X 2 ∂Y 2 ∂X ∂Y
∂2Ψ ∂2Ψ
+ = −ζ; (3)
∂X 2 ∂Y 2
∂θ ∂θ ∂θ 1  ∂2θ ∂2θ 
+U +V = + ; (4)
∂τ ∂X ∂Y Pr Re ∂X 2 ∂Y 2
∂Ψ ∂Ψ ∂V ∂U
U= , V =− , ζ= − . (5)
∂Y ∂X ∂X ∂Y
Here Re = U0 L/ν, Pr = ν/α, and Ri = Gr/ Re2 are the Reynolds number, the Prandtl number, and the Richardson
number, respectively, and Gr = gβΔT L3 /ν 2 is the Grashof number.
The dimensionless initial and boundary conditions of the problem can be described as follows:
at τ = 0,
U = V = Ψ = 0, θ = 0, 0  X  1, 0  Y  1;
at τ > 0,
∂θ
U = V = Ψ = 0, = 0, Y = 0,
∂Y
(6)
∂θ
U = 1, V = 0, Ψ = 0, = 0, Y = 1,
∂Y
U = V = Ψ = 0, θ = sin (2πY ), X = 0,

U = V = Ψ = 0, θ = ε sin (2πY + ϕ), X = 1.


637
Comparison of average Nusselt numbers for a lid-driven cavity with a heated top wall and a cooled bottom

Nu
Gr Re = 400 Re = 1000

Present work Data of [24] Data of [30] Present work Data of [24] Data of [30]
102 4.09 4.05 3.84 6.48 6.55 6.33
104 3.85 3.82 3.62 6.47 6.50 6.29
106 1.10 1.17 1.22 1.66 1.81 1.77

Here ε = Ar /Al is the amplitude ratio of the sinusoidal temperature on the right wall to that of the left wall of
the cavity. The heat transfer rate across the cavity plays an important role in thermal engineering applications.
Once the numerical values of the temperature function are known, it is possible to obtain the heat transfer rate in
terms of the Nusselt number. As both vertical walls contain heating and cooling regions, we have to calculate the
Nusselt number on both
 walls. The local Nusselt
 numbers along the left and right walls of the cavity are expressed
 
as Nul = (−∂T /∂x) and Nur = (−∂T /∂x) , respectively. In the heated half of the sidewalls of the cavity,
X=0 X=1
the fluid in the cavity gains heat from the side wall; thus, Nu > 0 in this region. The fluid loses heat in the cooled
half of the side wall; hence, Nu < 0 in this region. The total heat transfer rate across the cavity is the sum of the
averaged Nusselt numbers along the heated halves of both vertical sidewalls:
 
Nu = Nul dY + Nur dY.

2. METHOD OF THE SOLUTION

The finite volume method [29] is employed to discretize the governing equations (2)–(5) together with the
initial and boundary conditions (6). The QUICK scheme and the central difference scheme are taken for the
convective terms and the diffusion terms, respectively. The successive over relaxation (SOR) procedure is applied to
the stream function equation (3). The central difference approximation is used to solve the velocity components (5).
The boundary values of vorticity are calculated by the second-order expression for ζi obtained by expanding the
stream function Ψ by using the Taylor series:
Ψi+2 − 8Ψi+1
ζi = + o(Δη 2 ).
2(Δη)2
Here i denotes the boundary node and Δη is the interval in the direction normal to the boundary. The time
step is taken as Δτ = 10−5 . Uniform grids along the X and Y directions are used. The grid sizes are tested
from 21 × 21 to 121 × 121 for Ri = 0.01 and 100. From the grid-independence test, it is decided that an 81 × 81
grid is sufficient to solve the system of equations. An iterative process solves each variable obtained from the
resulting set of difference equations. The convergence criterion used in the study for all variables (ζ, Ψ, T ) is
|(Φn+1 (i, j) − Φn (i, j))/Φn+1 (i, j)|  10−5 (n is the iteration number). The trapezoidal rule is applied to find the
average Nusselt number. The validation of the present computational code is verified against the existing numerical
result of Sharif [24] and Iwatsu et al. [30] for a top-heated moving lid and bottom-cooled square cavity filled with
air (Pr = 0.71) (see the table). It is observed from the table that there is good agreement between the result of the
present code and the existing results.

3. RESULTS AND DISCUSSION

The effect of cavity inclination on combined convection in a lid-driven cavity with nonuniform heating on
both sidewalls is numerically examined. Two horizontal walls of the cavity are insulated. The Richardson number,
the amplitude ratio, the phase deviation, and the angle of inclination of the cavity are the main parameters to
control the flow and the temperature distribution in this study. The ratio Gr/ Re2 is a measure of the relative
strength of natural convection and forced convection. The Reynolds number is varied from 10 to 103 , while the
638
I II (a) III
1

-0.3
-0.9-0.7

-0.6
4

-0.9
-0.6 -0.

-0.9

-0.9.7

-0.4
-0.6
-0.4 -0.6

-0.4
-0.7

-0

-0.9
-0.7

-0.7
-0.4 -0 .67
-0.
-0.3

-0.6
-0.4.3
-0

-0.3
-0.3 -0.1
-0.1 0.0 0.0
0.1
0.1

0.3 0. 0
0.3 0.3

0.9 0.7 0.60.41


0.4 0.4

-0.1

0.6

0.7
0.9
0.7

0.6
0.7
0.6
0.6

0.4
0.4

0.6
0.7
0.3

0.7

0.4
0.1

0.3
0

0.1
0.1

0.3
0
-0.1
2
0

-0.1
-0.1
-0.1

-0.1
-0.1 -0.3
-0.1 -0.1
-0.1 -0.1
-0.3
-0.1 -0
-0.2
-0.1 -0
-0.1 -0
-0.1 -0.2
-0.2
-0.1 -0 -0 -0.2
-0 -0.1
-0 -0
-0 -0.1
-0.1 0
-0
-0
-0 0
-0 -0

I II (b) III
-0.4

1 -0.6
-0.3
-0.4
-0.9-0.7
-0. 6

-0.4

-0 .9.7
-0.9
-0.7
-0.7

-0
-0.6
-0.7

-0.6
-0.4 -0 .6

-0.4
.3.4 -0 .6

-0.7
-0-0

-0.3
-0.1
-0.3 -0.3 0
0.1
-0.1
0 0.3
0.1
0.3 0.4
0.3 0.0.-00.1

0.4
0.9 0.7 0.60.41

0.6

0.7
0.9

0.9
0.7
0.60.7

0.7

0.6
0.6
0.4
0.4 0.3
0.1

0.7

0.6

0.3
0.1

0.4
-0.1
0

0.3

2
-0

-0.1
-0.1
-0.1
-0.1
-0.1
-0.1
-0
0.4

-0
-0.1
-0
-0.1
-0.1 -0 0.4
-0.1 0.3
-0.1 -0.1 -0 0.3
0.3
-0.1 0.2
-0 0.2
-0 0.2
-0 0.1
-0 0.1
-0 0
0
-0 -0 0

I II (c) III
-0.4

1 -0.4
-0.6
-0.3
-0.4
-0.9-0.7
-0. 6

-0 .6
-0.6

-0.7
-0.9
-0.4 -0.6 -0.7

-0.9
-0.4
-0.7
-0.6

-0.7
-0.3

-0.7
-0 .9
-0.4

-0
-0.3 -0.1.3
-0.1 -0.3 0
0 0.1
0.1 0.3
0.3
0.3 0.10 .1

0.4 0.4
-0
0.9 0.7 0.60.4
0.9

0.9
0.6
0.7

0.6
0.7

0.7
0.6

0.6
0.6

0.7
0.7
0.4

0.9
0.3
0.1

-0.1

0.3
0

0.3

0.4 0.4

2
-0

-0.1
-0.1
-0.1
-0
-0
-0
-0
-0
-0
-0

-0.1
-0.1 0.5
0.4
-0.1
-0.1 -0.1 0.4
-0.1 0.4
0.3
-0.1 0.3
-0 0
0.2
-0 0.2
-0 0 0.2
-0 0.1
-0 0 0.1
-0

-0 0

Fig. 2. Isotherms (1) and streamlines (2) for ε = 1, ϕ = 0, and γ = 0◦ (a), 30◦ (b), 45◦ (c), 60◦ (d), and 90◦ (e):
Ri = 0.01 (I), 1 (II), and 100 (III) (see the final on the next page).

639
I II (d) III
-0.4

1 -0.4

-0.3

-0.6
-0.4
-0.9-0.76

-0.6
-0.

-0.9 .7
-0
-0. -0.7

-0.7
-0.7

-0.7
-0.6
-0.6
-0.3-0.46

-0.4

-0.4
-0.3 -0
-0 .3
-0.1 0 .1
0 -0.3
0.1
0.1
0.3
0.3
0.4 0.4

0.3 0. -0.1
0.9 0.7 0.60.41
0

0.7
0.9
0.79

0.7
0.
0.7

0.6

0.6
0.9
0.6
0.6
0.40.3

0.6
0.7
0.1
0.3 0.3

-0.1
0.4

0
0.4

2
-0.1
-0.1
-0
-0
-0
-0
-0
-0

-0.1
-0.1 0.5 0.5
-0.1
-0.1 -0.1 0.4
-0.1 0 0.4
0.3 0.3
-0.1
-0 0 0.3
-0 0.2
-0 0 0.2
-0 0.1
-0 0 0.1
-0 0.1
0 0

I II (e) III
-0.4

1 -0.4
-0.7 -0.4
-0.3

-0.7
-0.9
-0.6
-0.7

.6

-0.9.7
-0.9

-0
-0.9

-0
-0.6
-0.6

-0.6
-0.7

-0.6
-0.7
-0 .4
.3.4
-0-0

-0.4

-0
.3
-0.1
-0.3
0 -0.1 0
-0.3
0.1 0.1
0.3
0.3
0.4
0

0.4
0.9 0.70.60.4
0.0.31

0.6
-0.1

0.9
0.6
0.7

0.7
0.9
0.70.6

0.6
0.6
0.40.3

0.7

0.7
0.1

0.3
0

0.3 0.4
-0.1

0.4

2 -0.1
-0.1
-0
-0
-0
-0
-0
-0

-0.1 -0.1
0.5 0.4
-0.1
-0.1
-0.1 0.4
-0.1 0 0.4
-0.1 0.3
-0 0 0.3
-0 0.2
-0 0 0.2
-0 0.2
0 0.1
-0
0 0.1
-0 0
0

Fig. 2. Final (see the beginning on the previous page).

Grashof number is fixed as 104 . The Richardson number is varied from 0.01 to 100.00. The phase deviations are
varied from 0 to π. The values of the amplitude ratio are 0, 0.5, and 1. The angle of inclination is taken to be 0,
π/4, π/2, 3π/4, and π.
3.1. Effect of Cavity Inclination

The angles for the cavity inclination under investigation are considered to be 0, 30, 45, 60, and 90◦ . The
phase deviations are fixed as 0 and π. Similarly, the amplitude ratio is taken to be 0 and 1. Therefore, three
cases are taken into account in this study. In the first case, the phase deviation and the amplitude ratio are taken
as 0 and 1, respectively. Hence, both walls have identical heating and cooling zones in the absence of the phase
deviation (ϕ = 0). The isotherms and streamlines for different inclination angles of the cavity with ε = 1 and
ϕ = 0 are shown in Fig. 2. In the forced convection-dominated mode and γ = 0◦ , the isotherms are clustered
along the heating and cooling zones on the both walls of the cavity. It indicates that thermal boundary layers with
high temperature gradients are formed. High heat transfer rates are observed in both zones (see Fig. 2a). The
isotherms do not penetrate significantly into the middle of the cavity. On the other hand, the fluid flow contains
a single clockwise-rotating eddy that occupies almost the whole cavity. Changing the cavity inclination angle does
not make any variation in the heat distribution and flow pattern at Ri = 0.01. There are minor changes in the
isotherm pattern for all cavity inclination angles. However, two eddies appear in the mixed convection mode at
γ = 0◦ and Ri = 100. The secondary eddy is formed at the lower bottom of the cavity. For the cavity inclination
640
I II (a) III
0.4

0.3
0.6

0.1
0 .3
0.70.6 0.4

0.6.4
0
-0 .9

-0.7.9
-0
-0.7

0.9

-0.6
0.9
0.7
-0 -0.6
-0.7
.1.3 -0.-0.6

0.7
-0.3.4
4

-0.1
0.1 0
-0-0

0
-0.4 .3 0.
-0 0.3 4
-0.1 0.1
0
0

0.1
.4 -0.1
-0.1
0

-0.3
-0.1

-0
-0.9 -0.7-0.6

-0.4.3
0

-0.6
-0.3

-0

-0.7 .6
-0.4
-0

-0 .9
0.7

-0.7
-0.9
0.6

0.7
0.4

0.4
0.6 0.7
0.3

0.4
0.3
0.6

0.3
0.1
0.1
-0.4

2 -0

-0.1
-0.1 0.1
0.1
-0.1 0.1
-0.1 0
0
-0.1
-0.1
-0.1 -0

-0.1
-0
-0.1
-0.1 -0
-0.1 -0
-0.1
-0.1 -0.2
-0 -0 -0
-0.1
-0 -0 -0.1
-0 -0.1

0
-0 -0 -0.1
-0.1
-0 -0
-0

I II (b) III
0.4

1
0.3

0.6

0.1
0.3
0.7 0.6
0.70.6 0.4

-0.7

-0.9
-0.9
0.9
4 -0.7

-0.6
0.4

-0.7
0.9
-0.6
.1.3 -0.-0.6

-0.4
-0.3

0.7
-0 .4
-0.1
0.1 0

0
-0-0

-0.3 0.3
-0.1 0.1
0
-0 -0 .4 -0.1.1

0.1
0.0 0

0
-0.9 -0..67 -0.3

.1
-0 0.3

-0 -0 .4
-0.9 .7 .6
-0

.4

-0 .9.7
-0.3

0.9

-0

-0
0.70.6

0.9

0.7

-0.6
0.6
0.4

0.4
0.6

0.3
0.7

0.3
0.3

0.4
0.1

0.1

0
-0.4

2 -0

0
-0.1
-0.1
-0.1 .1 0.1
-0
-0.1 0
0
-0.1 -0.1
-0.1 -0.1
-0.1
-0.1
-0.1
-0.1 -0
-0 -0
-0 -0
-0 -0.2
-0 -0.2
-0 -0.2
-0 -0.2
-0 -0.1
-0 -0.1
-0 -0.1
-0 -0.1
-0
0

-0

I II (c) III
0.4

1
0.3
0.70.6 0.4

0.6 0.4
0.3
0.1

0.90.7
-0.7 9

-0.7
-0.9
-0.

0.6
0.9
4 -0.7

-0.6
-0.4 -0.6
.1.3 -0.-0.6

0.7
-0.1
-0.3
0.1 0
-0-0

0 -0.4 -0.3
0.3
-0.1
0.1
0
0.1
-0 -0 .4 -0.1

0 -0.1
-0.9 -0..67 -0.3

-0.3
-0 .1
0

-0.4.3
-0

-0.7 -0.6
0.7

-0.7
0.7

-0.9

-0.4
-0.6
0.6

0.9
0.9
0.6

0.7
0.4

0.4 6
0.3
0.
0.4
0.3

0.3
0.1

0.1

-0.4

2
-0.1

-0.1 0.1
0
-0.1
-0.1
-0.1

-0.1
-0.1
-0.1
-0.1
-0.1 -0
-0.1 -0
-0.1 -0.3
-0.1 -0 -0.2
-0 -0.2
-0 -0
-0.2 -0.2
-0 -0 -0.1
-0 -0.1
-0 -0 -0.1
-0 -0.1
-0 -0
-0

Fig. 3. Isotherms (1) and streamlines (2) for ε = 1, ϕ = π, and γ = 0◦ (a), 30◦ (b), 45◦ (c),
60◦ (d), and 90◦ (e): Ri = 0.01 (I), 1 (II), and 100 (III) (see the final on the next page).

641
I II (d) III
1 0.4

0.1
0.3

0.40.3
0.70.6 0.4

-0 .7
-0.9
0.1 -0.1-0.3-0.4-0.6-0.7

0.9

0.7
0.6

-0.6
-0.7
6

0.6
-0.4 -0.

0.7
-0.1
-0.3
0
0.3
0 -0.4
-0.3
-0.1
0.1
0

.4 -0.1
0 0.1
0 -0.1

-0.3
-0
-0.9 -0.7-0.6

-0.3
-0
.1 -0.3

-0.9 .7
-0 .7.6

-0
-0

-0.4
-0.4
0.7

-0.6
0.6

0.9
0.6
0.7

0.7
0.6
0.4

0.4

0.4
0.3
0.3

0
0.1
0.1

0.3
-0.4

2
-0.1
-0.1
0

-0.1 0
-0.1

-0.1
-0.1
-0.1
-0.1
-0.1
-0.1 -0
-0.1 -0.3
-0.1 -0
-0.1 -0.3
-0 -0.2
-0 -0 -0.2
-0 -0.2
-0 -0 -0.2
-0 -0.1
-0 -0 -0.1
-0 -0.1
-0 -0
-0

I II (e) III
1

0.4 0.3
0.3

0.1
0.70.6 0.4

0.4
0.9 7 0.6
-0.7 -0.9

-0.9
0.9
-0.7
-0.1-0.3 -0.4 -0.6

0.

0.9
0.7
0.6
-0.7
-0.3 -0.4
-0.6

-0.4
-0.6
-0.1
0.1 0

-0.3
0.3
-0.1
0.1
0
-0.1 0 0.1

-0.1
0 -0.3
.4
.6.3

-0.6 0.4 .3
-0
-0-0

- -0
-0.1

-0.70.6
0.7

0.9

-
-0.9 .7

0.7
-0.7

-0
-0.4
0.70.6

-0.9
0.4
0.3

0.4
0.6

0.6
0.3

0.3

0.1

0
0.1

0.4
-0.4

2
-0.1
-0.1
.1
-0

.1
-0

-0.1 -0.1
-0.1

-0.1 -0.1 0.1


-0.1 -0 -0.2
-0.1 -0 0.1 -0.2
-0.1
-0.1 -0 -0.2
-0 0 -0.1
-0
-0 -0.1
-0 -0 0
-0 -0.1
-0 -0.1
-0
-0 -0
-0
-0

Fig. 3. Final (see the beginning on the previous page).

angles γ = 0–90◦ , the size of the primary eddy decreases. On the other hand, the secondary eddy grows in size.
Finally, clockwise and counterclockwise eddies are formed when the right wall of the cavity becomes the top wall
at γ = 90◦ .
The heat distribution remains unchanged at γ = 0–90◦ . A tri-cellular flow structure appears in the cavity
at γ = 0◦ and Ri = 100. A primary clockwise eddy diagonally elongates in the middle of the cavity. Two
counterclockwise secondary eddies appear on both sides of the diagonal eddy. At γ = 0–30◦ , the primary eddy is
totally reduced in size and moves to the moving wall. Moreover, two secondary eddies merge together and rotate
in the counterclockwise direction. On increasing γ from 30 to 45◦ , the secondary eddy elongates along the moving
wall, and one more secondary eddy occurs at the right bottom corner. At γ = 45–90◦ , the shape of the primary
eddy becomes circular, and both secondary eddies disappear. Accordingly, changing cavity inclination affects the
fluid flow significantly in the buoyancy-dominated mode.
In the second case, the phase deviation and the amplitude ratio are taken as π and 1, respectively. Figure 3
displays the isotherms and streamlines for different inclination angles of the cavity at ε = 1 and ϕ = π. Changing
the cavity inclination angles does not make any variation in the temperature distribution and flow pattern for
Ri  1. On the other hand, in the buoyancy-dominated mode, three eddies are formed at γ = 0◦ . At γ = 0–30◦ , the
central eddy is broken into two eddies, one being greater in size. At γ = 45◦ , the smaller eddy disappears and the
top and bottom eddies merge together. At γ = 45–90◦ , the secondary eddy grows in size. Finally, two horizontal
eddies are formed at Ri = 100.
642
Nu (a) Nu (b)
12 1
4 1
11 2
6
3
5

2
2 4 3
3
3 4
4 5
1 5
2
0 10 20 30 40 50 60 70 80 90 g 0 10 20 30 40 50 60 70 80 90 g

Nu (c)
9
1
8
2

5
3
4

4
3
5
2
0 10 20 30 40 50 60 70 80 90 g

Fig. 4. Average Nusselt number versus the angle of inclination γ for different values of the phase
deviation ϕ and amplitude ratio ε: (a) ε = ϕ = 0; (b) ε = 1 and ϕ = 0; (c) ε = 1 and ϕ = π;
Ri = 0.01 (1), 0.1 (2), 1 (3), 10 (4), and 100 (5).

The average heat transfer rate across the cavity as a function of γ for different values of the amplitude
ratio and phase deviation is shown in Fig. 4. At Ri = 0.1, increasing cavity inclination does not influence the
average heat transfer rate in the forced convection mode. At Ri = 1, the average Nusselt number increases as the
cavity inclination angle is increased. On the other hand, fluctuations are observed in the mean heat transfer rate
at γ = 0–90◦ in the buoyancy-dominated mode. It is found from Fig. 4b that slight changes happen in the average
Nusselt number at Ri < 1 on increasing cavity inclination. At Ri = 1 and 10 and γ = 0◦ , the average heat transfer
rate is high. However, as γ is increased from 0 to 90◦ , the average heat transfer rate decreases and then increases
again. At Ri = 100 and γ = 0–90◦ , the average heat transfer rate gradually increases in the buoyancy-dominated
mode.

3.2. Effect of the Amplitude Ratio

Figure 5 shows the isotherms and streamlines for different amplitude ratios with a fixed phase deviation
and inclination of the cavity. The nonuniform temperature is maintained on the left wall, whereas the right wall is
changed from uniform to nonuniform temperature according to changes in the amplitude ratio. Different amplitude
ratios are taken only on the right wall of the cavity. The right wall of the cavity has a uniform temperature
distribution at ε = 0. Figure 5a reveals that a thermal boundary layer with a high temperature gradient is formed
along both zones of the left wall of the cavity in the forced convection-dominated mode at ε = 0. Heat transfer
occurs only near the left wall. The corresponding streamlines show that the clockwise rotating single cell occupies
the major portion of the cavity.
643
I II (a) III
1

-0.7 .9
-0.7

-0.6
-0.9

-0
-0.7
64
-0 .1-0.3-0.-0.

-0.4
-0.3 .6
-0

-0.1
-0.4
3
-0.
-0.1
0

0.9

0.9
0.7

0.7

0.7
0.6

0.6

0.6
0.4
0.4
0.3

0.1
0.3

0
0.1
0.4

0.1

0.3
0
0
0
2

0
-0.1
-0.1 0
-0.1
-0.1
-0.1 -0.1
-0.1

-0.1 -0.1
-0.1 -0
-0.1 -0
-0.1
-0.1 -0 -0.2
-0.1 -0 -0.2
-0 -0.1
-0 -0.1
-0 -0.1
-0 -0.1
-0 -0 -0.1
-0 -0.1
-0
-0 -0
-0

I II (b) III
1

-0.7
-0.7

-0.3

-0 .3
-0.4

-0.4
-0.9

-0.4 -0.6
-0.3

-0.6
-0.7
-0.6
.3.4
-0-0

-0.3

-0.4 .3
-0

-0.1

-0.1 0
-0.1

0
0.1
0.30.1 0

0.1

0.3

0.3
0.7
0.9
0.7

0.6
0.4

0.6

0.7

0.4
0.6

0.4
0.4
0.3

0.3

0.4
0.1

0.3
0

0
0.1

0.1

2
-0.1 0
-0.1

-0.1 -0.1
-0.1
-0.1 -0.1
-0.1 -0.1 -0.1
-0
-0.2 -0.2
-0.1 -0
-0.1 -0 -0.2 .2
-0.1 -0
-0.1 -0
.
-0 -0.2
.
-0 -0.1
-0 -0.1
-0 -0.1
-0 -0.1
-0 -0 -0.1 0
-0 -0
-0

I II (c) III
1
-0.3
-0.9-0.7

-0.6
4

-0.9

-0.9
-0.6 -0.

-0 .9.7
-0.6

-0.4
-0.6

.4

-0
-0.4 -0.7

-0.9
-0

-0.7

-0.7
-0 .7
-0.4 -0 .6
-0.3

.6
-0

-0.4.3
-0
-0.3
-0.3 -0.1
0
-0.1 0 0.1
0

0.1
0.3 0.3
0.9 0.7 0.60.41

0.4
0.3 0.

0.4
0.6
-0.1

0.7
0.9
0.7

0.6
0.7
0.6
0.6

0.4
0.4

0.6
0.7
0.3

0.7

0.4
0.1

0.3
0

0.1

0.1
0.3
-0.1

2
0

-0.1
-0.1
-0.1
-0.1
-0.1 -0.3
-0.1 -0.1
-0.1
-0.1 -0.3
-0.1 -0 -0.2
-0.1 -0
-0.1 -0 -0.2
-0.1 -0.2
-0.1
-0 -0 -0.2
-0 -0.1
-0 -0.1
-0 -0
-0 -0.1 0
-0
-0 0
-0 -0

Fig. 5. Isotherms (1) and streamlines (2) for ϕ = 0, γ = 0◦ , and ε = 0 (a), 0.5 (b), and 1 (c):
Ri = 0.01 (I), 1 (II), and 100 (III).

644
Nu (a) Nu (b)
11
11

9
9

7 7
3 3
5 5
2 2

3 1 3 1

1 -2 1 -2
10 10-1 1 10 102 Ri 10 10-1 1 10 102 Ri

Fig. 6. Average Nusselt number versus the Richardson number at ϕ = 0 and γ = 0◦ (a) and 90◦ (b):
ε = 0 (1), 0.5 (2), and 1 (3).

Moreover, on increasing the Richardson number from 0.01 to 1.00, neither forced convection nor natural
convection dominates; the strong boundary layer disappears in this mode. The center of the eddy moves toward
the moving wall. Moreover, the shape of the fluid flow is slightly changed at Ri = 1. As the Richardson number
is increased from 1 to 100, the isotherms penetrate into the middle of the cavity at ε = 0. As the buoyancy force
dominates over the shear force, the single cell flow pattern is broken into a multi-cell flow pattern. It is found that
the thermal distribution and the fluid flow are not affected by changes in the amplitude ratio at Ri = 0.01.
On changing the amplitude ratio, the right wall of the cavity is changed from uniform to nonuniform tem-
perature. So, both walls have nonuniform temperature distributions at ε > 0. However, changing the amplitude
ratio from zero to one does not make any impact on the fluid flow at Ri = 0.01 because the shear force dominates
over the buoyancy force in the forced convection mode. With the effect of the amplitude ratio, a secondary eddy
develops in the right bottom corner of the cavity in the mixed convection mode. An increase in the amplitude ratio
leads to an increase in the size of the secondary eddy. At Ri = 100 and ε = 0, two eddies are observed in the middle
of the cavity (see Fig. 5a). As the amplitude ratio is increased, the primary eddy elongates diagonally from the left
bottom corner to the right top corner. Moreover, the flow is of a tri-cellular structure with one large diagonal cell
and two smaller corner cells.
The average Nusselt number versus the amplitude ratio with a fixed phase deviation and different inclination
angles of the cavity is illustrated in Fig. 6. On increasing the amplitude ratio, the average heat transfer rate increases.
On the other hand, the average Nusselt number decreases as the Richardson number increases from 0.01 to 100.00
for a given amplitude ratio. The mean heat transfer rate seems to be high at Ri = 0.01, whereas a low heat transfer
rate is observed in the buoyancy-dominated mode (Ri = 100). The nonuniform temperature on the both walls plays
a significant role in heat transfer than a uniform temperature at γ = 0◦ , as well as at γ = 90◦ . As both side walls
have sinusoidally varying temperature distributions, heat transfer is enhanced at all inclination angles of the cavity.
This clearly shows that the nonuniform temperature on both walls is beneficial for enhancement of convection heat
transfer in enclosures.

3.3. Effect of the Phase Deviation

Figure 7 exhibits the isotherms and streamlines for different phase deviations with a fixed amplitude ratio and
inclination of the cavity. Heating and cooling zones are formed on the both walls at sinusoidal heating temperature
distributions. These zones on the right wall change as the phase deviation ϕ is changed from 0 to π. Both walls
have identical heating and cooling zones at ϕ = 0. In the forced convection mode, thermal boundary layers with
a high temperature gradient are formed along the heating and cooling zones on both walls of the cavity for all
phase deviations ϕ. No considerable changes in the temperature gradient are observed in the middle of the cavity.
Moreover, the clockwise rotating single eddy resides in the major part of the cavity at Ri = 0.01 (see Fig. 7a).
645
I II (a) III
1

-0.3
-0.9-0.7

-0.6
4

-0.9
-0.6 -0.

-0.4
-0.9

-0.9.7
-0.6
-0.4 -0.6

-0.4
-0.7

-0

-0.9
-0.7

-0.7
-0.4 -0 .67
-0.
-0.3

-0.6
-0.4.3
-0

-0.3
-0.3 -0.1
0
-0.1 0
0.1

0
0.1
0.3 0.3

0.9 0.7 0.60.41


0.4

0.3 0.
0.4

-0.1

0.6

0.7
0.6
0.9
0.7

0.7
0.6
0.6

0.4
0.4

0.6
0.7
0.3

0.7

0.4
0.1

0.3
0

0.1
0.1

0.3
-0.1
2
0

-0.1
-0.1
-0.1
-0.1
-0.1 -0.3
-0.1 -0.1
-0.1
-0.1 -0.3
-0.1 -0
-0.2
-0.1 -0
-0.1 -0
-0.1 -0.2 -0.2
-0.1 -0 -0 -0.2
-0 -0.1
-0 -0.1
-0 -0
-0 -0.1 0
-0
-0 0
-0 -0

I II (b) III

0.3 0.1

0.3
0.1
0

0
1 -0.1

-0.7 .9
-0.1
-0.9
-0 .4 -0.1

-0
-0.6 -0.7

-0 -0 -0.7
-0 .4

.3
-0.9 -0.7 -0.6

-0
-0.3

-0.3.4 .6

-0.6 .4
-0.9-0.7

-0.3
-0

-0.6

-0.7
-0.6
-0.3

-0.4

-0.4
-0.3
-0.1

-0.1
0 0
0.1
0 -0.1
0.1
0.7

0.9
0.1
0.7
0.9

0.3 0.3 0.4


0.6

0.4

0.7
0.4

0.6
0.7 .40.6
0.40.3

0.6

0.6
0.1

0.3

0.6

0.4
0

0.3

0.3
0
0

0.1

0.7
0.1

0.7
0.9
-0.1

2
.1
-0

-0.1
-0.1
-0.1
-0.1 -0.1
-0.1 -0.4
-0.1 -0.4
-0.1
-0.1
-0 -0.3
-0.1 -0.3
-0.1 -0
-0.1 -0.3
-0.1 -0 -0.2
-0 -0 -0.2
-0 -0.2
-0 -0 -0.1
-0 -0.1
-0 -0.1
-0 -0
-0
-0

I II (c) III
0.7

0.9 0.7
0.3
0.6
0.4

1
0.4

0
0.6

0.4 .6
0.3

-0 -0

0.3
.7
0.1
-0.7

.9
-0.4
-0.6

0.1
-0.6

-0.7

0.1
-0 .4

0
-0.1

0
43

-0.6
-0.3

-0 -0.4
-0.6-0.-0.
0

-0 .6 -0.4 -0.3
-0 .3

-0.1
.1

.1 -0.3
-0
-0.9 -0.7

-0.7
-0.1

-0.3

-0.9 .7

-0.1
-0.4
-0.6
-0
0.9
0.70.6

0.7
0.9

0.6

0
0.7

0
0.4

0.6
0.4

0.1
0.40.3

0.1
0.3

0
0.1

0.3

0.1

0.4 0.4
0

0.6
0.7 .6
0

0.3
0.1

0
0.1

0.4
0.3

-0.1
-0.1
-0.1
-0.1

-0.1
-0.1 -0.1

-0.1
-0.1
-0.1 -0.3
-0.1 -0.1 -0
-0.3
-0 -0.3
-0.1 -0.3
-0.1 -0 -0.2
-0 -0.2
-0
-0 -0.1 -0.2
-0 -0 -0.1
-0 -0.1
-0 -0 -0.1
-0 -0
-0

Fig. 7. Isotherms (1) and streamlines (2) for ε = 1, γ = 0◦ , ϕ = 0 (a), π/4 (b), π/2 (c), ϕ = 3π/4 (d), and π (e):
Ri = 0.01 (I), 1 (II), and 100 (III) (see the final on the next page).

646
I II (d) III

0.7
1

0.7
0.9

0.9
0.4 0.6

0.4

0.7
-0 .9
-0.7
-0.9

-0.6
0.4 0.6

-0.7
0.6

.4 -0.6
-0.7
-0-0.3 -0.4-0.6

-0.1
0
0.3

.1

-0.3 -0
0.3 0.1
-0.4

0.1
0 0
-0.1

0.1 0.3
0.1 -0.3

-0.3.1
64
-0-0.

-0.6 .4 -0.3
0
-0.1

-0.9 -0.7-0.

.3
-0
.1

-0.7
-0.6
-0
-0.7
-0
0

-0.4
0.9

0.9
0.60.4

0.6 0.7
0.7

0.4

0.7
0.4
0.6
0.3

0.3

0.1
0.3

0
-0.1

0.1
0 0
0.1 0.1 0.1

-0

2 -0
0

.1
-0
0

0
-0.1
-0.1 -0.1

-0.1 -0.1
-0.1
-0.1
-0.1
-0.1 -0
-0.1 -0.1 -0.2
-0.1 -0 -0.2
-0 -0 -0.2
-0 -0.2
-0 -0 -0.1
-0 -0.1
-0.1
-0 -0 -0.1
-0.1
-0 -0
-0

I II (e) III
0.4

1
0.3

0.6

0.1
0 .3
0.70.6 0.4

0.6.4
0
-0.9

-0.7.9
-0
-0.7

0.9

-0.6
0.9
0.7
-0 -0.6
-0.7
.1.3 -0.-0.6

0.7
-0 .3.4
4

-0.1
0.1 0
-0-0

-0.4 0.4
-0.3 0.3
-0.1 0.1
0
0
0.1
.4 -0.1

-0.1
0
-0.3

-0.1
-0
-0.9 -0.7-0.6

-0.4.3
0

-0 .7 -0.6
-0.3

-0

-0.7 .6
-0.4
-0
-0.9
0.7

-0.9
0.6

0.7
0.4

0.4
0.6 0.7
0.3

0.4
0.3
0.6

0.3
0.1
0.1

-0.4

2 -0

-0.1
-0.1 0.1
-0.1 0.1
0.1
-0.1 0
0
-0.1
-0.1
-0.1 -0
-0.1
-0
-0.1
-0.1 -0
-0.1
-0.1 -0
-0.1 -0 -0.2
-0 -0
-0.1
-0 -0 -0.1
-0 -0.1
-0.1
0

-0 -0 -0.1
-0 -0
-0

Fig. 7. Final (see the beginning on the previous page).

As the phase deviation ϕ is varied from 0 to π, the flow pattern does not alter in the forced convection-dominated
mode. In the mixed convection mode, the isotherms are distributed in the middle of the cavity at ϕ = 0 (see
Fig. 7a).
At Ri = 1, there is a significant change in the heat distribution along the right wall on changing the phase
deviation ϕ from 0 to π. Two eddies appear in the mixed convection mode at ϕ = 0 (see Fig. 7a). As the phase
deviation ϕ is varied from 0 to π, the size of the secondary cell is reduced, and it disappears at the end. Figure 7a
shows that the isotherms are well established in the middle of the cavity in the natural convection-dominated mode
at ϕ = 0. In the corresponding streamlines, three eddies are formed at ϕ = 0 in the buoyancy-dominated mode,
and both walls have identical heating and cooling zones. The primary eddy diagonally elongates from the left
bottom corner to the right top corner. The secondary eddies disappear on changing ϕ from 0 to π/2. Moreover, on
increasing ϕ from π/2 to 3π/4, a multi-cellular flow pattern is formed in the cavity. Two secondary eddies merge
together and grow in size at ϕ = π, and a horizontal multi-cellular flow pattern is formed.
The average Nusselt number for different phase deviations with a fixed amplitude ratio and inclination of
the cavity is shown in Fig. 8 as a function of the Richardson number. It is clearly understood that the average
heat transfer rate seems to be high at Ri = 0.01 and is gradually reduced as Ri is changed from 0.01 to 100.00. At
Ri < 1, a high heat transfer rate is reached at ϕ = π/4, whereas the heat transfer rate is low at ϕ = 0 or ϕ = π.
On the other hand, in the natural convection-dominated mode, a high heat transfer rate is observed at ϕ = π/2,
647
Nu
12

10
4
8 2
5
1
6
3
4

2 -2
10 10-1 1 10 102 Ri

Fig. 8. Average Nusselt number versus the Richardson number for ε = 1, γ = 0◦ , and ϕ = 0 (1),
π/4 (2), π/2 (3), 3π/4 (4), and π (5).

whereas a low mean heat transfer rate is found at ϕ = 0 or ϕ = π. It is concluded that the heat transfer rate is
very low at ϕ = 0 or ϕ = π for all values of the Richardson number.

CONCLUSIONS

The effect of cavity inclination on combined convection in a lid-driven cavity with nonuniform heating on
both sidewalls is numerically studied. Two horizontal walls are insulated. The top wall of the cavity moves at
a constant speed in its own plane. The Richardson number, the amplitude ratio, the phase deviation, and the
cavity inclination angle are considered as the governing parameters for this study. The major findings are as
follows. The mean heat transfer rate significantly increases in the buoyancy-dominated mode on increasing the
cavity inclination if both walls have identical heating and cooling zones. The heat transfer rate is not affected
by changing the inclination angle of the cavity in the forced convection mode, whereas this effect is significant
in the natural convection mode. Tilting the cavity does not make much impact on heat transfer if the walls have
different heating and cooling zones. A unicellular flow pattern is observed in the forced convection-dominated mode,
whereas a multi-cellular flow pattern occurs in the natural convection mode for all examined values of the amplitude
ratio, cavity inclination angle, and phase deviation. The heat transfer rate decreases on increasing the Richardson
number from 0.01 to 100.00. Nonuniform temperature distributions on both walls are beneficial for enhancement
of convection heat transfer in enclosures, as compared to a uniform wall temperature.

REFERENCES

1. A. J. Chamkha, S. H. Hussain, and Q. R. Abd-Amer, “Mixed Convection Heat Transfer of Air Inside a Square
Vented Cavity with a Heated Horizontal Square Cylinder,” Numer. Heat Transfer A 59, 58–79 (2011).
2. P. Talukdar, “Mixed Convection and Non-Gray Radiation in a Horizontal Rectangular Duct,” Numer. Heat
Transfer A 59, 185–208 (2011).
3. A. A. Mohammad and R. Viskanta, “Flow and Heat Transfer in a Lid-Driven Cavity Filled with Stably Stratified
Fluid,” Appl. Math. Model. 19, 465–472 (1995).
4. A. J. Chamkha, “Hydromagnetic Combined Convection Flow in a Vertical Lid-Driven Cavity with Internal
Heat Generation or Absorption,” Numer. Heat Transfer A 41, 529–546 (2002).
5. M. A. Seddeek, “Effects of Magnetic Field and Variable Viscosity on Forced Non-Darcy Flow about a Flat Plate
with Variable Wall Temperature in Porous Media in the Presence of Suction and Blowing,” J. Appl. Mech. Tech.
Phys. 43 (1), 13–17 (2002).
648
6. T.-M. Jeng and S.-C. Tzeng, “Heat Transfer in a Lid-Driven Enclosure Filled with Water-Saturated Aluminum
Foams,” Numer. Heat Transfer A 54, 178–196 (2008).
7. D. A. Perumal and A. K. Dass, “Multiplicity of Steady Solutions in Two-Dimensional Lid-Driven Cavity Flows
by Lattice Boltzmann Method,” Comput. Math. Appl. 61, 3711–3721 (2011).
8. V. Sivakumar, S. Sivasankaran, P. Prakash, and J. Lee, “Effect of Heating Location and Size on Mixed Con-
vection in Lid-Driven Cavities,” Comput. Math. Appl. 59, 3053–3065 (2010).
9. V. I. Bukreev, N. V. Gavrilov, and A. V. Chebotnikov, “Effect of the Nonmonotonic Temperature Dependence
of Water Density on Free Convection from a Linear Heat Source,” J. Appl. Mech. Tech. Phys. 52 (1), 24–30
(2011).
10. M. Bhuvaneswari, S. Sivasankaran, and Y. J. Kim, “Effect of Aspect Ratio on Convection in a Porous Enclosure
with Partially Active Thermal Walls,” Comput. Math. Appl. 62, 3844–3856 (2011).
11. E. Thandapani, A. R. Ragavan, and G. Palani, “Finite-Difference Solution of Unsteady Natural Convection
Flow Past a Nonisothermal Vertical Cone under the Influence of a Magnetic Field and Thermal Radiation,”
J. Appl. Mech. Tech. Phys. 53 (3), 408–421 (2012).
12. B. Abourida, M. Hasnaoui, and S. Douamna, “Transient Natural Convection in a Square Enclosure with Hori-
zontal Walls Submitted to Periodic Temperature,” Numer. Heat Transfer A 36, 737–750 (1999).
13. N. H. Saeid and Y. Yaacob, “Natural Convection in a Square Cavity with Spatial Sidewall Temperature Varia-
tion,” Numer. Heat Transfer A 49, 683–697 (2006).
14. E. Bilgen and R. B. Yedder, “Natural Convection in Enclosure with Heating and Cooling by Sinusoidal Tem-
perature Profiles on One Side,” Int. J. Heat Mass Transfer 50, 139–150 (2007).
15. Q.-H. Deng and J.-J. Chang, “Natural Convection in a Rectangular Enclosure with Sinusoidal Temperature
Distributions on Both Side Walls,” Numer. Heat Transfer 54, 507–524 (2008).
16. A. Al-Zoubi and G. Brenner, “Simulating Fluid Flow Over Sinusoidal Surfaces using the Lattice Boltzmann
Method,” Comput. Math. Appl. 55, 1365–1376 (2008).
17. S. Sivasankaran, V. Sivakumar, and P. Prakash, “Numerical Study on Mixed Convection in a Lid-Driven Cavity
with Non-Uniform Heating on Both Sidewalls,” Int. J. Heat Mass Transfer 53, 4304–4315 (2010).
18. P. R. Nath, P. M. V. Prasad, and D. R. V. Prasada Rao, “Computational Hydromagnetic Mixed Convective
Heat and Mass Transfer Through a Porous Medium in a Non-Uniformly Heated Vertical Channel with Heat
Sources and Dissipation,” Comput. Math. Appl. 59, 803–811 (2010).
19. S. Sivasankaran, A. Malleswaran, J. Lee, and P. Sundar, “Hydro-Magnetic Combined Convection in a Lid-Driven
Cavity with Sinusoidal Boundary Conditions on Both Sidewalls,” Int. J. Heat Mass Transfer 54, 512–525 (2011).
20. M. Bhuvaneswari, S. Sivasankaran, and Y. J. Kim, “Magnetoconvection in a Square Enclosure with Sinusoidal
Temperature Distribution on Both Walls,” Numer. Heat Transfer A 59, 167–184 (2011).
21. J. Rasoul and P. Prinos, “Natural Convection in an Inclined Enclosure,” Int. J. Numer. Methods Heat Fluid
Flow 7, 438–478 (1997).
22. C. Cianfrini, M. Corcione, and P. P. Dell’Omo, “Natural Convection in Tilted Square Cavities with Differentially
Heated Opposite Walls,” Int. J. Therm. Sci. 44, 441–451 (2005).
23. E. Bilgen and H. Oztop, “Natural Convection Heat Transfer in Partially Open Inclined Square Cavities,” Int.
J. Heat Mass Transfer 48, 1470–1479 (2005).
24. M. A. R. Sharif, “Laminar Mixed Convection in Shallow Inclined Driven Cavities with Hot Moving Lid on Top
and Cooled from Bottom,” Appl. Therm. Eng. 27, 1036–1042 (2007).
25. Chamkha and A. Al-Mudhaf, “Double-Diffusive Natural Convection in Inclined Porous Cavities with Various
Aspect Ratios and Temperature-Dependent Heat Source or Sink,” Heat Mass Transfer 44, 679–693 (2008).
26. A. Bairi, “Nusselt–Rayleigh Correlations for Design of Industrial Elements: Experimental and Numerical In-
vestigation of Natural Convection in Tilted Square Air Filled Enclosures,” Energy Convers. Management 49,
771–782 (2008).
27. Y. Varol, F. Oztop, A. Koca, and F. Ozgen, “Natural Convection and Fluid Flow in Inclined Enclosure with a
Corner Heater,” Appl. Thermal Eng. 29, 340–350 (2009).
28. B. Ghasemi and S. M. Aminossadati, “Natural Convection Heat Transfer in an Inclined Enclosure Filled with
a Water–CuO Nanofluid,” Numer. Heat Transfer A 55, 807–823 (2009).
29. S. V. Patankar, Numerical Heat Transfer and Fluid Flow (Hemisphere, Washington, 1980).
30. R. Iwatsu, J. M. Hyun, and K. Kuwahara, “Mixed Convection in a Driven Cavity with a Stable Vertical
Temperature Gradient,” Int. J. Heat Mass Transfer 36, 1601–1608 (1993).

649

You might also like