Sie sind auf Seite 1von 20

Numerical Heat Transfer, Part A: Applications

An International Journal of Computation and Methodology

ISSN: 1040-7782 (Print) 1521-0634 (Online) Journal homepage: http://www.tandfonline.com/loi/unht20

Natural convection heat transfer and entropy


generation inside porous quadrantal enclosure
with nonisothermal heating at the bottom wall

Shantanu Dutta, Arup Kumar Biswas & Sukumar Pati

To cite this article: Shantanu Dutta, Arup Kumar Biswas & Sukumar Pati (2018): Natural
convection heat transfer and entropy generation inside porous quadrantal enclosure with
nonisothermal heating at the bottom wall, Numerical Heat Transfer, Part A: Applications, DOI:
10.1080/10407782.2018.1423773

To link to this article: https://doi.org/10.1080/10407782.2018.1423773

Published online: 23 Jan 2018.

Submit your article to this journal

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=unht20
NUMERICAL HEAT TRANSFER, PART A
https://doi.org/10.1080/10407782.2018.1423773

none defined

Natural convection heat transfer and entropy generation


inside porous quadrantal enclosure with nonisothermal
heating at the bottom wall
Shantanu Duttaa, Arup Kumar Biswasa, and Sukumar Patib
a
Department of Mechanical Engineering, National Institute of Technology Durgapur, Durgapur, India; bDepartment
of Mechanical Engineering, National Institute of Technology Silchar, Silchar, India

ABSTRACT ARTICLE HISTORY


In this work, we study numerically the natural convection heat transfer Received 18 October 2017
and entropy generation characteristics inside a two-dimensional porous Accepted 1 January 2018
quadrantal enclosure heated nonuniformly from the bottom wall. The effect
of Darcy number is significant in dictating the Nusselt number only for
higher values of Rayleigh number and the variation is more profound for
larger values of Darcy number. The variation of entropy generation rate is
significant with the Darcy number only for higher values of Rayleigh number.
The entropy generation due to heat transfer is the significant contributor of
irreversibility at low values of Darcy number, while for larger values of Darcy
number and Rayleigh number entropy generation due to fluid friction
becomes dominant.

1. Introduction
The study of natural convection in an enclosure filled with porous medium has received considerable
attention from the research community in the last few decades owing to its manifold applications in
several engineering and environmental problems which include thermal insulation, electronic equip-
ment cooling, geothermal energy extraction, nuclear reactor system, solar heating, grain storage, food
processing, etc. A comprehensive documentation of these applications can be found in Nield and
Bejan [1], Bejan et al. [2], Ingham et al. [3], Pop and Ingham [4], Vafai [5], and Ingham and Pop [6].
Appreciating the importance of natural convection in an enclosure filled with porous medium, a
rich and variety of numerical and experimental investigations have been conducted by several
researchers to study the fluid flow and heat transfer characteristics. The geometry of enclosures
considered were square [7–18], rectangular [19–22], triangular [23–26], parallelogram [27–31],
trapezoidal [32–35], and rhombic [36–39]. Astanina et al. [15] in a recent work studied the transient
natural convection inside a square porous cavity with the consideration of temperature-dependent
viscosity. According to their findings, the convective flow and heat transfer become intensified with
an increase in the viscosity variation parameter for the porous media, while the effect is opposite in
the case of pure fluid. Chou et al. [16] analyzed the effects of temperature-dependent viscosity
on natural convection inside porous media between two concentric spheres. Pandit and
Chattopadhyay [22] used higher order compact scheme to investigate the transient natural convection
in a deep enclosure filled with porous medium. The works presented by researchers on different
geometries as mentioned mainly focused on the effects of different boundary conditions, aspect ratio
for rectangular enclosure, undulation on the wall, inclination angle for square and trapezoidal
enclosure, magnetic field, presence of heat sources, presence of block or geometrical obstruction, etc.

CONTACT Sukumar Pati sukumarpati@gmail.com; sukumar@mech.nits.ac.in Department of Mechanical Engineering,


National Institute of Technology Silchar, Silchar 788010, India.
Color versions of one or more of the figures in the article can be found online at www.tandfonline.com/unht.
© 2018 Taylor & Francis
2 S. DUTTA ET AL.

Nomenclature
cp specific heat of the fluid, J kg 1 K 1 U, V dimensionless x and y velocity components,
Da Darcy number respectively
g acceleration due to gravity, m s 2 u, v x and y velocity components,
h convective heat transfer respectively, m s 1
coefficient, W m 2 K 1 x, y axial and transverse coordinates, respect-
k thermal conductivity, W m 1 K 1 ively, m
K permeability, m2
L enclosure length, m Greek symbols
Nu local Nusselt number α thermal diffusivity
Nu average Nusselt number β coefficient of thermal expansion
p pressure, N m 2 θ dimensionless temperature
P nondimensional pressure µ dynamic viscosity
Pr Prandtl number υ kinematic viscosity
Re Reynolds number ρ density of fluid, kg m 3
S total dimensionless entropy Subscripts
Sθ dimensionless entropy generation due to avg average
heat transfer b bottom wall
Sψ dimensionless entropy generation due to c cold wall
fluid friction min minimum
T temperature, K max maximum

on the flow and heat transfer characteristics. Substantial advancements have also been reported in
modeling of porous media which include the effects of inertia, drag, acceleration, local thermal
nonequilibrium between fluid and solid matrices, temperature-dependent viscosity, etc.
It is also worth mentioning in this context that natural convection inside complex-shaped
enclosures has numerous applications including building systems with complex geometries, solar
collectors, lubrication systems, electric machinery, cooling system of microelectronic systems and
devices. Research investigations on the natural convection heat transfer inside complex-shaped
geometries, however, have been only a few in number [40–47]. Shiina et al. [40] experimentally
studied natural convection in a hemisphere heated from below and developed a correlation for aver-
age Nusselt number for a wide range of Rayleigh numbers and Prandtl numbers (106 � Ra � 1010 and
6 � Pr � 13,000). Chen and Cheng [41] performed a numerical and experimental investigation to
study the natural convection heat transfer in an inclined arc-shaped enclosure by varying Grashof
number (Gr) in the range 104 � Gr � 107 and the inclination angle (θ) in the range 0 � θ � π.
Kim et al. [42] analyzed natural convection heat transfer in a meniscus-shaped cavity for both
cylindrical and spherical geometries. Aydin and Yesiloz [43] performed an experimental and numeri-
cal investigation on natural convection heat transfer in a two-dimensional quadrantal enclosure
isothermally heated from the bottom and cooled from the vertical side for different values of Rayleigh
number, Ra in the range of 103–107. Yesiloz and Aydin [44] studied the buoyancy-induced flow and
heat transfer in an inclined quadrantal cavity both experimentally and numerically by varying the
inclination angle in the range of 0° � ϕ � 360° and the Rayleigh number in the range of 105–107.
Sen et al. [45, 46] performed the numerical investigation to analyze the natural convection heat
transfer in quadrantal cavity having hot bottom wall and cold curved wall and using heater on
adjacent walls for Rayleigh number in the range of 103 � Ra � 107 and found out that both flow
and temperature fields are affected by a changing Ra. They also found out that heat transfer increases
with an increase in Rayleigh number and the flow strength increases with an increase in the size of
heater on the vertical wall compared to the bottom wall and temperature fields are also affected. In
contrast, with an increase in size of heater on both sides of adjacent walls flow strength does not
change significantly. Bose et al. [47] performed numerical simulation to study the natural convection
heat transfer in quadrantal cavity having finned hot vertical wall and cold bottom wall for Rayleigh
number in the range of 104 � Ra � 106.
NUMERICAL HEAT TRANSFER, PART A 3

It is worth mentioning that the process of transfer of heat due to finite temperature difference is an
irreversible process causing the generation of entropy. To utilize the energy resources efficiently, the
entropy generation associated with transfer of heat during physical processes should be minimum. To
develop a thermodynamically efficient thermal system, Bejan [48] introduced the entropy generation
minimization based on second law of thermodynamics. Thus, proper attention should be made to
minimize the entropy generation while transferring heat during any physical process. The entropy
generation for natural convection in a porous enclosure of different geometries has been extensively
analyzed by several researchers [13, 14, 18, 23–24, 49–54]. The main focus of most of the works was to
identify the major sources of the entropy generation as well as the location of the local maxima of
entropy generation for different values of Rayleigh number and Darcy number.
It is important to mention in this context that buoyancy-driven flows in fluid-filled enclosures satu-
rated with porous medium and subjected to nonisothermal thermal boundary condition have a wide
variety of practical applications. Appreciating the usefulness of the thermal boundary conditions, a great
deal of research work has been reported in the literature [9–11, 18–20, 24, 25, 30, 34, 38–39, 49–53] on
natural convection heat transfer in a porous enclosure with nonuniform wall temperature.
To the best of the authors’ knowledge, no work has been reported in the literature on natural
convection inside a quadrantal enclosure filled with porous media induced by nonuniform heating.
Accordingly, the aim of the current investigation is to analyze the natural convection heat transfer
and entropy generation characteristics inside a two-dimensional porous quadrantal enclosure. The
bottom wall of the enclosure is heated sinusoidally and the vertical wall is cooled by a constant
temperature bath, while the curved wall is kept adiabatic. The results are illustrated in the form of
streamlines, isotherms, local Nusselt number along the walls, average Nusselt number, entropy
generation due to heat transfer and fluid flow, and average Bejan number for a wide range of Rayleigh
numbers (103–106) and Darcy numbers (10 5 to 10 3).

2. Theoretical formulation
Consider a two-dimensional quadrantal enclosure saturated with porous media as shown in Figure 1.
The vertical wall is cooled at constant temperature, while the curved wall is thermally insulated. The
bottom wall is nonuniformly heated as given in the following [23]:
� � ��
Th Tc 2px
T ð x Þ ¼ Tc þ 1 cos ð1Þ
2 L

where Th is the maximum temperature of the heated wall, Tc is the temperature at the cold wall, and L
is the height of the enclosure.
The major assumptions made to analyze the present problem are as follows:
i. The fluid confined within porous bed is Newtonian and the flow is steady, laminar, and
incompressible.
ii. The effect of viscous dissipation is neglected.
iii. The physical properties except the density in the body force term are considered to be constant.
The variation of density in the body force term with temperature follows Boussinesq
approximation.
iv. The radiation heat transfer is neglected.
v. Darcy–Forchheimer model is used to simulate the momentum transfer in the porous medium.
vi. The temperature of the fluid phase is equal to the temperature of the solid phase everywhere and
local thermodynamic equilibrium is applicable.
Under these assumptions, the governing transport equations of mass, momentum, and energy con-
servation pertinent to the analysis of the physical problem described as above can be written in
dimensionless form as follows:
4 S. DUTTA ET AL.

Figure 1. Schematic diagram of the physical model.

Continuity equation:
qU qV
þ ¼0 ð2Þ
qX qY
x-Momentum equation:
� 2 �
qU qU qP q U q2 U U
U þV ¼ þ Pr þ Pr ð3Þ
qX qY qX qX 2 qY 2 Da
y-Momentum equation:
� 2 �
qV qV qP q V q2 V Pr
U þV ¼ þ Pr þ V þ RaPrh ð4Þ
qX qY qY qX 2 qY 2 Da
Energy equation:
qh qh q2 h q2 h
þV U ¼ 2þ 2 ð5Þ
qX qY qX qY
In the above equations, various dimensionless parameters are defined as X ¼ Lx, Y ¼ Ly , U ¼ uL
a,
2 3
T Tc pL gbðTh Tc ÞL Pr
V ¼ vL K
a , h ¼ Th Tc , P ¼ qa2 , Da ¼ L2 , Ra ¼ n2 , where L is the length of the enclosure, u
and v are the dimensional velocities in x and y directions, respectively, α is the thermal diffusivity,
p is the dimensional pressure, ρ is the density of fluid, g is gravitational acceleration, β is the volu-
metric thermal expansion coefficient, ν is the kinematic viscosity, K is the permeability of the porous
media, Pr is Prandtl number, Da is Darcy number, and Ra is Rayleigh number.
The following boundary conditions are used to solve Eqs. (2)–(5):
U ¼ V ¼ 0 and h ¼ 0 ðalong ACÞ ð6bÞ

1
U ¼ V ¼ 0 and h ¼ ð1 cosð2pxÞÞ ðalong ABÞ ð6bÞ
2

U ¼ V ¼ 0 and n:rh ¼ 0 ðalong BCÞ ð6cÞ


where n denotes the normal direction to the wall.
NUMERICAL HEAT TRANSFER, PART A 5

2.1. Nusselt number calculation


The thermal field obtained from the numerical solution of the governing transport equations is
synthesized by a dimensionless parameter, called, Nusselt number. The local Nusselt number along
any wall of the enclosure can be calculated as follows:
hL qh
Nul ¼
¼ ð7Þ
k qn
where q=qn is the dimensionless derivative along the direction of the outward drawn normal to the
surface.
The average Nusselt number for a wall can be found as:
ZL
1
Nuavg ¼ Nul dl ð8Þ
L
0

2.2. Entropy generation


For the problem under consideration, the rate of entropy generation has two contributors, namely,
the entropy generation due to heat transfer (Sθ) originated due to diffusion of thermal energy caused
by local temperature gradient and the entropy generation due to fluid friction (Sψ) resulting from the
viscous effects within the fluid and at the fluid–solid interfaces. Thus, the local rate of entropy gen-
eration in dimensionless form [23] can be written as:
S ¼ Sh þ Sw ð9Þ
where
"� �2 � �2 #
qh qh
Sh ¼ þ ð10aÞ
qX qY
" ( � �2 � �2 ! � � )#
2 2
� qU qV qU qV 2
Sw ¼ n U þV þ Da 2 þ þ þ ð10bÞ
qX qY qY qX
where ξ is the irreversibility distribution ratio and is defined as:
� �2
mT0 a
n¼ pffiffiffi ð11Þ
k K DT
2
In the present work, the value of ξ is chosen as 10 [50]. One can obtain the global total entropy
generation (ST) in the enclosure as:
ST ¼ Sh;total þ Sw;total ð12Þ
where Sθ,total and Sψ,total are the total entropy generation due to heat transfer and fluid friction,
respectively, and are obtained as follows:
Z
Sh;total ¼ Sh dX ð13aÞ
X
Z
Sw;total ¼ Sw dX ð13bÞ
X
The Bejan number (Be) is also used to measure the irreversibility distribution and is defined as the
ratio of rate of entropy generation due to heat transfer to the total entropy generation rate. From the
6 S. DUTTA ET AL.

distribution of local entropy generation, one can estimate Bejan number as:

Sh
Be ¼ ð14Þ
Sh þ Sw

The relative global dominance of entropy generation due to heat transfer is quantified by the
average Bejan number (Beavg), which can be determined as follows:

Sh;total
Beavg ¼ ð15Þ
Sh;total þ Sw;total

The value of average Bejan number varies in the range 0 � Beavg � 1. The entropy generation due
to heat transfer dominates when Beavg ≫ 0.5, whereas the entropy generation due to fluid friction
becomes dominant when Beavg ≪ 0.5.

3. Numerical solution methodology and model validation


The governing transport equations (2)–(5) together with the boundary conditions (6a)–(6c) as
outlined in the previous section are solved numerically using finite element method. The Galerkin
weighted method is used to transform the governing equations into a system of integral equations.
The detailed description of the methodology can be found in Zienkiewicz and Taylor [55]. The
Gauss’ quadrature method [55] is used to perform the integration. The Newton–Raphson technique
is used
� to solve
�� the set of algebraic equations. The convergence criterion is set as
max �/nþ1 /n � /n � 10 6 , where / represents any transport variable.
A grid sensitivity analysis has been performed to confirm that the results presented herein are
independent of the grid used. Three different arrangements comprising of 33,744, 168,280, and
316,241 elements have been considered for the analysis. The values of averaged Nusselt number at
the bottom wall for different grid systems with Da ¼ 10 3, Pr ¼ 0.71, and four different values of
Rayleigh number (¼103, 104, 105, and 106) are presented in Table 1. The grid system having
168,280 elements has been used to simulate all the computations because on further refinement of
grids the change in the average Nusselt number is almost negligible (less than 1%).
To test the accuracy of the numerical code used in the current investigation, results presented
have been validated by Aydin and Yesiloz [43] and Lauriat and Prasad [56]. Figure 2 displays the
comparison of the results of streamline and isotherm contour for a quadrantal cavity filled with
water (Pr ¼ 6.62) when the bottom wall is uniformly heated and the vertical wall is cooled for
Ra ¼ 1.7 � 105. It becomes apparent from Figure 2 that the results are in good agreement with the
results available in the literature [43]. The results of average Nusselt number are found to be in good
agreement with Lauriat and Prasad [56] for porous vertical cavity with heated left wall, cooled right
wall, and top and bottom walls adiabatic. The comparisons are presented in Table 2. Furthermore, the
results on the local entropy generation due to fluid flow and heat transfer are also compared with the
results presented by Ilis et al. [51] and the same can be found in Bhardwaj and Dalal [24].

Table 1. Comparison of average Nusselt number on the bottom wall for different grid systems with Pr ¼ 0.71 and Da ¼ 10 3.
No. of elements
Ra 33,744 168,280 316,241 Relative error (%)
103 1.22 1.22 1.22 0
104 1.24 1.24 1.24 0
105 1.84 1.85 1.86 0.5
106 5.55 5.57 5.59 0.3
NUMERICAL HEAT TRANSFER, PART A 7

Figure 2. Streamline and isotherm contours for a quandrantal enclosure filled with a fluid (Pr ¼ 6.62) where bottom wall is heated
uniformly, vertical sidewall is cooled at constant temperature, while curved wall is adiabatic at Ra ¼ 1:7�105 .

Table 2. Comparison of average Nusselt number on the uniformly heated bottom wall from the present work with those reported
by Lauriat and Prasad [56].
Da Ra Present work Lauriat and Prasad [56] Difference (%)
2
10 103 1.00 1.02 1.9
104 1.62 1.70 4.7
105 4.10 4.26 3.7
4
10 105 1.03 1.06 2.8

4. Results and discussion


The prime focus of the current investigation is to analyze the transport characteristics of thermal
energy and entropy generation inside a two-dimensional porous quadrantal enclosure subjected to
nonuniform temperature distribution at the bottom wall. The numerical experiments are conducted
for Rayleigh number in the range of 103 � Ra � 106 and Darcy number 10 5 � Da � 10 3, which are
within the range of practical relevance [23, 50]. Important to note that the ability of flow through
porous media is represented by Darcy number, while Rayleigh number is used to describe the relative
ability to transport heat due to density difference caused by temperature gradient in the flow field to
that due to diffusion of heat. The working fluid is taken as air (Pr ¼ 0.71). The important findings
obtained from the current numerical simulation are presented in the form of streamlines, isotherms,
local Nusselt number along the heated as well as cooled walls and average Nusselt number.
Furthermore, in keeping view of identifying the thermodynamically efficient systems, local entropy
generation due to heat transfer and fluid flow and average Bejan number are also illustrated in detail.

4.1. Flow and temperature field


To elucidate the heat transfer and entropy generation characteristics for different values of Darcy
number and Rayleigh number, the analysis of flow and thermal field is very important. Accordingly,
the streamlines and isotherms are illustrated in the left and right panels in Figures 3–5 for three
different Darcy numbers (10 5, 10 4, and 10 3) and Rayleigh numbers in the range of 103–106.
For all values of Rayleigh number and Darcy number, two asymmetric circulating cells, rotating in
the opposite directions are observed within the enclosure. It is quite intuitive that the two cells are
not symmetric because of irregular nature of geometry and asymmetric boundary condition. The fluid
located in the region just above the midpoint of the bottom wall is heated maximum as compared to
the fluids located at other regions because of the imposed temperature distribution at the bottom wall.
The fluid after being heated from the bottom wall rises up and then descends along the vertical and
8 S. DUTTA ET AL.

5
Figure 3. Streamlines and isotherms for different Rayleigh numbers at Da ¼ 10 (a) Ra ¼ 103, (b) Ra ¼ 104, (c) Ra ¼ 105,
(d) Ra ¼ 106.

curved walls toward the edges of the bottom wall, thus forming two cells, left one rotating in antic-
lockwise direction and right one rotating in clockwise direction. For all cases, the strength as well as
the size of the left circulation cell is more as compared to that of the right cell. The reasons for such
NUMERICAL HEAT TRANSFER, PART A 9

4
Figure 4. Streamlines and isotherms for different Rayleigh numbers at Da ¼ 10 (a) Ra ¼ 103, (b) Ra ¼ 104, (c) Ra ¼ 105,
(d) Ra ¼ 106.

observation are the imposed boundary conditions and the profile of the walls adjacent to the bottom
one (vertical and curved). For Da ¼ 10 5, the strength of the vortex is negligibly small when Ra ¼ 103
because of combined confluence of high resistance to flow and low buoyancy force. With an increase
10 S. DUTTA ET AL.

3
Figure 5. Streamlines and isotherms for different Rayleigh numbers at Da ¼ 10 (a) Ra ¼ 103, (b) Ra ¼ 104, (c) Ra ¼ 105,
(d) Ra ¼ 106.

in Rayleigh number, the buoyancy force increases as a result of which the strength of the vortices
rotating in both anticlockwise and clockwise directions is enhanced. For example, the maximum
strength of vortex at Ra ¼ 103 is 0.001, while for Ra ¼ 106, it is 1.1. For Da ¼ 10 5, the isotherms
NUMERICAL HEAT TRANSFER, PART A 11

as displayed in the right panel of Figure 3, are distributed smoothly throughout the enclosure with
densely packed in a region close to middle of the heated bottom wall (0.25 � X ≤ 0.75) and the influ-
ence of Rayleigh number on the distribution of isotherms is not significant. This indicates that heat is
transferred mainly due to conduction because of the fact that resistance force due to fluid friction
dominates over the buoyancy force. Moreover, the temperature gradient on the bottom wall is
maximum at the middle for all values of Rayleigh number because of the fact that the temperature
on the bottom wall is maximum at the middle.
For Da ¼ 10 4, the distribution of streamlines and isotherms as shown in Figure 4 is qualitatively
similar to that for Da ¼ 10 5 up to Ra ¼ 105 except the fact of enhancement in the strength of the
vortices. Under these situations, the force due to fluid friction being dominant opposes buoyancy
force to establish any fluid motion in the enclosure and hence the mode of heat transfer is conduction.
However, when Rayleigh number is increased to 106, because of strengthening the buoyancy force, the
viscous resistance is overcome and the fluid motion is established. This is aptly reflected through the
enhanced strength in the vortices (maximum 14.5) (see Figure 4d). The corresponding isotherms
become distorted because of increased strength of convection and the principal mode of transport
of thermal energy within the enclosure is changed from conduction to convection.
It can be observed from the distribution of streamlines and isotherms as displayed in Figures 5a
and 5b that for Da ¼ 10 3, the buoyancy force is unable to overcome the viscous force for Rayleigh
number up to 104 and accordingly the dominant mode of thermal energy transport is diffusion of
heat. However, for higher values of Rayleigh number (105 and 106), the strength of the vortices is
increased to a great extent. The size of right circulation cell decreases, while that of left cell is
increased significantly with an increase in Rayleigh number. In contrast to the smoothly distributed
isotherms for lower values of Rayleigh number, isotherms are distorted at higher Rayleigh number
and moreover the thickness of the thermal boundary layer decreases resulting in steep temperature
gradient at the heated wall. Such observation can be elucidated from the fact that the permeability
of the medium being more at higher values of Darcy number, fluid can easily flow through the pores
and enhanced strength of buoyancy force over viscous force at higher Rayleigh number.
The combined effect of these two results in enhancement in the intensity of flow.

4.2. Nusselt number


4.2.1. Local Nusselt number
To explore the influences of Rayleigh number and Darcy number on the heat transfer characteris-
tics, the variation of local Nusselt number along the heated bottom and cooled vertical walls is
plotted in Figures 6 and 7, respectively. It can be mentioned in this context that local Nusselt num-
ber at the wall is dependent on the temperature gradient at the wall which in turn depends on the
thickness of the thermal boundary layer. Important to reiterate that as the thickness of thermal
boundary layer decreases, the temperature gradient increases and accordingly the local Nusselt
number increases. First of all, it can be seen from Figure 6a that for Da ¼ 10 5 the variation of local
Nusselt number at the bottom wall follows the same trend for all values of Rayleigh number con-
sidered. Furthermore, the local Nusselt number at any particular position at the bottom wall is
invariant with Rayleigh number as can be inferred from Figure 6a. This is because of the fact that
for lower values of Darcy number, the resistance to flow is more and hence fluid cannot flow easily
even for higher values of Rayleigh number which is clearly visible from streamlines contour (see
Figure 3). Accordingly, the mechanism of transfer of heat is the diffusion which solely depends
on the temperature differential. Important to notice that the variation of the local Nusselt number
at the bottom wall follows the same pattern as that of imposed temperature distribution on the wall.
The local Nusselt number is negative at both the corner representing that the heat is transferred
from the fluid to the wall at those location. The reverse heat transfer as seen can be explained from
the streamline and isotherm contours as follows: The temperature of the bottom wall is maximum
at the midpoint, while it is minimum at the two ends. The fluids located close to the middle of the
12 S. DUTTA ET AL.

Figure 6. Variation of local Nusselt number on heated bottom wall for different values of Rayleigh number (a) Da ¼ 10 5,
(b) Da ¼ 10 4, (c) Da ¼ 10 3.

bottom wall being heated maximum rises upward and then moves toward the edges of the bottom
wall thus forming two rotating cells.
It can be noticed from Figure 6b that for Da ¼ 10 4 the trend in the variation of local Nusselt
number at the bottom wall is same for Rayleigh number up to 105 and moreover the change in its
value with Rayleigh number at any position on the wall is negligibly small. This indicates that the
resistance to flow of fluid through the pores of the medium is still high because of low permeability
and accordingly the mechanism of transport of thermal energy is diffusion. However, when Rayleigh
number is increased to 106, there is an alteration in the variation of the local Nusselt number both in
its pattern and numerical value as noticed from the distribution of streamlines and isotherms. The
location of the maximum local Nusselt number has shifted from the middle of the bottom wall toward
the curved wall. Such observation can be explained as follows. The strength of the buoyancy force gets
enhanced and becomes dominant over viscous force in this situation and as a result the dominant
mode of heat transfer becomes convection. The temperature gradient at the heated wall becomes
steep at higher values of Rayleigh number because of thin thermal boundary layer thickness and this
augments the rate of heat transfer substantially.
As can be seen from Figure 6c that for Da ¼ 10 3 the variation of local Nusselt number at the
bottom wall with Rayleigh number is invariant both qualitatively and quantitatively for Rayleigh
NUMERICAL HEAT TRANSFER, PART A 13

number up to 104 owing to the fact that buoyancy force dominates over viscous force although
permeability of the medium is relatively high and accordingly the heat transfer is mainly dominated
by conduction. For Rayleigh number beyond 104, the local Nusselt number is strongly influenced by
Rayleigh number and the enhancement in the Nusselt number is significant for Ra ¼ 106. This can be
explained from the fact that the strength of the buoyancy force increases with an increase in Rayleigh
number and the same becomes dominant over viscous force once Rayleigh number is increased
beyond 104. As a result, the dominant mechanism of heat transfer is changed from conduction to
convection and the heat transfer increases with an increase in Rayleigh number and becomes
maximum when Rayleigh number is increased to 106.
The variation of local Nusselt number along the vertical cooled wall for different values of Rayleigh
number and Darcy number is plotted in Figures 7a–7c. One can see that local Nusselt number is
negative throughout indicating that the heat transfer is from the fluid to the wall. It can be seen from
Figure 7a that for Da ¼ 10 5, the change in Nusselt number with Rayleigh number is negligibly small
similar to bottom wall for the reasons already mentioned. It is further seen that the local Nusselt num-
ber increases initially, reaches its maximum value and thereafter decreases. This is owing to the fact
that heat transfer is mainly due to conduction and temperature gradient at the vertical wall initially
increases and thereafter decreases as can be seen from the isotherm contour (Figure 3). The similar

Figure 7. Variation of local Nusselt number on cooled vertical wall for different values of Rayleigh number (a) Da ¼ 10 5,
(b) Da ¼ 10 4, (c) Da ¼ 10 3.
14 S. DUTTA ET AL.

trend in the variation of Nusselt number is observed for Rayleigh number up to 105 when Da ¼ 10 4
and for Rayleigh number up to 104 when Da ¼ 10 3 for the reasons already enlightened. However,
the local Nusselt number increases monotonically with distance when Da ¼ 10 4 and Ra ¼ 106, which
is in stark contrast as compared to other values of Rayleigh number. Similarly, when Da ¼ 10 3, the
local Nusselt number increases monotonically for Ra ¼ 105 and 106. Such contrasting features can be
elucidated from the isotherm contour as can be seen from Figures 4d and 5c and 5d.

4.2.2. Average Nusselt number


Table 3 displays the values of average Nusselt number on the heated bottom wall and cooled vertical
wall for different values of Rayleigh number and Darcy number. It is noteworthy to mention that for
Ra ¼ 103, the average Nusselt number on both bottom and vertical walls is invariant with Darcy
number because the mode of heat transfer is conduction at Ra ¼ 103 and accordingly Darcy number
has almost no influence on the heat transfer characteristics. The trend in the variation of average
Nusselt number on either bottom or vertical wall with Darcy number for Ra ¼ 104 is similar to that
for Ra ¼ 103 implying that conduction-dominated heat transfer is still applicable. However, contrast-
ing features can be observed as the Rayleigh number is increased to 105 and 106. The average Nusselt
number is increased with an increase in Darcy number and the rate of enhancement is more profound
when Darcy number value is changed from 10 4 to 10 3. Therefore, the rate of heat transfer both
from heated wall to the fluid and from the fluid to the cooled wall is maximum when both Rayleigh
number and Darcy number are maximum within the range considered.

4.3. Entropy generation


As mentioned earlier, the total irreversibility for the present work is due to combined effects of
entropy generation due to heat transfer and fluid friction. It is important to mention in this context
that for efficient design of thermal systems one should have a fair idea about the distribution of
individual constituents of entropy generation inside the enclosure. Accordingly, the distribution of
local entropy generation due to heat transfer and fluid friction is plotted in Figures 8–10 for two
Rayleigh numbers (105 and 106) and Darcy numbers ranging from 10 5 to 10 3. One can observe
from Figure 8 that at low Darcy number, the irreversibility due to fluid friction is negligibly small
even for large Rayleigh number. This is attributed to the fact that at low values of Darcy number,
the porous medium offers high resistance to flow which causes very small flow velocity and thus
the induced flow is unable to penetrate the medium. It can further be observed that the entropy gen-
eration due to heat transfer is important in a region close to the bottom wall. The magnitude of Sθ is
maximum in the vicinity of the midposition of the bottom wall for all the cases considered because of
steep temperature gradient. The maximum value is almost independent of Rayleigh number when
Da ¼ 10 5, whereas for higher values of Darcy number, the value varies significantly with Rayleigh
number. The entropy generation due to fluid flow is very small for all Rayleigh numbers when
Da ¼ 10 5. As Darcy number is increased from 10 5 to 10 3, maximum entropy generation due
to both heat transfer and fluid flow is increased and the enhancement is more profound at Ra ¼ 106.
The irreversibility due to fluid flow is significant not only within a region close to the middle of the
bottom wall like the case of entropy generation due to heat transfer but also a region close to the
vertical wall and this region is more significant for higher values of Rayleigh number.

Table 3. Average Nusselt number on the bottom and vertical walls for different values of Rayleigh number and Darcy number.
Average Nusselt number on bottom wall Average Nusselt number on vertical wall
5 4 3 5 4 3
Ra Da ¼ 10 Da ¼ 10 Da ¼ 10 Da ¼ 10 Da ¼ 10 Da ¼ 10
3
10 0.774 0.776 0.778 0.775 0.778 0.778
104 0.775 0.797 0.799 0.776 0.796 0.798
105 0.776 0.816 1.18 0.779 0.82 1.18
106 0.798 1.319 3.55 0.80 1.33 3.57
NUMERICAL HEAT TRANSFER, PART A 15

5
Figure 8. Local entropy generation due to heat transfer (Sθ) and fluid friction (Sψ) for Da ¼ 10 (a) Ra ¼ 105 and (b) Ra ¼ 106.

4
Figure 9. Local entropy generation due to heat transfer (Sθ) and fluid friction (Sψ) for Da ¼ 10 (a) Ra ¼ 105 and (b) Ra ¼ 106.
16 S. DUTTA ET AL.

3
Figure 10. Local entropy generation due to heat transfer (Sθ) and fluid friction (Sψ) for Da ¼ 10 (a) Ra ¼ 105 and (b) Ra ¼ 106.

Table 4. Total entropy generation rate and average Bejan number for different Darcy numbers.
Ra Da Sθ,max Sψ,max ST Beavg
104 10 5
13.781 0.00008 13.78108 0.999
4
10 13.471 0.0011 13.4721 0.999
3
10 13.374 0.099 13.473 0.992
106 10 5
13.830 0.0837 13.9137 0.993
4
10 30.202 15.00 45.202 0.668
3
10 133.65 1041.00 1174.65 0.113

To identify the dominant source of irreversibility globally, average Bejan number is calculated for
different values of Rayleigh number and Darcy number and is presented in Table 4. It can be seen from
Table 4 that for Ra ¼ 104, average Bejan number is close to unity for all the values of Darcy number.
This indicates that for lower values of Rayleigh number, the irreversibility caused by fluid friction is
negligibly small in the enclosure and the entropy generation is mainly because of heat transfer due
to conduction. It can be further observed from Table 4 that the major contributor of irreversibility
is strongly dependent on Darcy number for Ra ¼ 106. The irreversibility due to fluid flow is negligibly
small for lower values of Darcy number (Da ¼ 10 5) because of low permeability of the medium. The
contribution of global fluid friction irreversibility increases with an increase in Darcy number and it
becomes the significant contributor for larger values of Darcy number (Da ¼ 10 3).

4. Conclusion
In the present work, we execute numerical experiments to study the natural convection heat transfer
and entropy generation characteristics inside a two-dimensional quadrantal enclosure filled with a
NUMERICAL HEAT TRANSFER, PART A 17

saturated porous medium. The enclosure is heated nonuniformly from bottom wall, while the vertical
wall is cooled to a constant temperature and the curved wall is kept adiabatic. The important findings
from the current investigation are summarized as follows:
For lower values of Darcy number (10 5), the heat transfer in the enclosure is mainly due to
conduction and it is not influenced by Rayleigh number. For higher values of Darcy number, the
dominant mode of heat transfer is dependent on Rayleigh number. The heat transfer is dominated
by conduction for lower values of Rayleigh number, whereas for higher values of Rayleigh number
heat transfer is mainly due to convection and the critical value of Rayleigh number for such transition
depends on Darcy number. Furthermore, the heat transfer is significantly altered at higher values of
Rayleigh number and Darcy number.
The entropy generation rate is almost invariant with Darcy number for lower values of Rayleigh
number (103, 104), whereas the rate of entropy generation varies significantly with the Darcy number
for higher values of Rayleigh number.
For all values of Rayleigh number, entropy generation due to heat transfer is the significant
contributor of irreversibility at low values of Darcy number, while for larger values of Darcy number
and Rayleigh number entropy generation due to fluid friction becomes dominant over that due to
heat transfer.

References
[1] D. A. Nield and A. Bejan, Convection in Porous Media, 3rd ed. New York: Springer, 2006.
[2] A. Bejan, I. Dincer, S. Lorente, A. F. Miguel, and A. H. Reis, Porous and Complex Flow Structures in Modern Tech-
nologies. New York: Springer, 2004.
[3] D. B. Ingham, A. Bejan, E. Mamut, and I. Pop (Eds.), Emerging Technologies and Techniques in Porous Media.
Dordrecht: Kluwer, 2004.
[4] I. Pop and D. B. Ingham, Convective Heat Transfer: Mathematical and Computational Modelling of Viscous Fluids
and Porous Media. Oxford: Pergamon, 2001.
[5] K. Vafai, Handbook of Porous Media. New York: Marcel Dekker, 2000.
[6] D. B. Ingham and I. Pop, Transport Phenomena in Porous Media. Oxford: Pergamon, 1998.
[7] N. H. Saeid and I. Pop, “Transient free convection in a square cavity filled with a porous medium,” Int. J. Heat
Mass Transfer, vol. 47, pp. 1917–1924, 2004. DOI:10.1016/j.ijheatmasstransfer.2003.10.014.
[8] N. H. Saeid and I. Pop, “Non-darcy natural convection in a square cavity filled with a porous medium,” Fluid Dyn.
Res., vol. 36, pp. 35–43, 2005. DOI:10.1016/j.fluiddyn.2004.10.004.
[9] N. H. Saeid, “Natural convection in porous cavity with sinusoidal bottom wall temperature variation,” Int.
Commun. Heat Mass Transfer, vol. 32, pp. 454–463, 2005. DOI:10.1016/j.icheatmasstransfer.2004.02.018.
[10] T. Basak, S. Roy, T. Paul, and I. Pop, “Natural convection in a square cavity filled with a porous medium: effects of
various thermal boundary conditions,” Int. J. Heat Mass Transfer, vol. 49, pp. 1430–1441, 2006. DOI:10.1016/j.
ijheatmasstransfer.2005.09.018.
[11] M. Sathiyamoorthy, T. Basak, S. Roy, and I. Pop, “Steady natural convection flow in a square cavity filled with a
porous medium for linearly heated side wall(s),” Int. J. Heat Mass Transfer, vol. 50, pp. 1892–1901, 2007.
DOI:10.1016/j.ijheatmasstransfer.2006.10.010.
[12] S. A. Khashan, A. M. Al-Amiri, and I. Pop, “Numerical simulation of natural convection heat transfer in a porous
cavity heated from below using a non-Darcian and thermal non-equilibrium model,” Int. J. Heat Mass Transfer,
vol. 49, pp. 1039–1049, 2006. DOI:10.1016/j.ijheatmasstransfer.2005.09.011.
[13] P. Datta, P. S. Mahapatra, K. Ghosh, N. K. Manna, and S. Sen, “Heat transfer and entropy generation in a porous
square enclosure in presence of an adiabatic block,” Trans. Porous Med., vol. 111, pp. 305–329, 2016. DOI:10.1007/
s11242-015-0595-5.
[14] P. A. K. Lam and K. A. Prakash, “A numerical study on natural convection and entropy generation in a porous
enclosure with heat sources,” Int. J. Heat Mass Transfer, vol. 69, pp. 390–407, 2014. DOI:10.1016/j.
ijheatmasstransfer.2013.10.009.
[15] M. S. Astanina, M. A. Sheremet, and J. C. Umavathi, “Unsteady natural convection with temperature-dependent
viscosity in a square cavity filled with a porous medium,” Trans. Porous Med., vol. 110, pp. 113–126, 2015.
DOI:10.1007/s11242-015-0558-x.
[16] H. Chou, H. Wu, I. Lin, W. Yang, and M. Cheng, “Effects of temperature-dependent viscosity on natural convec-
tion in porous media,” Numer. Heat Transfer A, vol. 68, pp. 1331–1350, 2015. DOI:10.1080/10407782.2015.
1012864.
18 S. DUTTA ET AL.

[17] L. Tian, C. Ye, S. Xue, and G. Wang, “Numerical investigation of unsteady natural convection in an inclined
square enclosure with heat-generating porous medium,” Heat Transfer Eng., vol. 35, pp. 620–629, 2014.
DOI:10.1080/01457632.2013.837676.
[18] P. Meshram, S. Bhardwaj, A. Dalal, and S. Pati, “Effects of the inclination angle on natural convection heat transfer
and entropy generation in a square porous enclosure,” Numer. Heat Transfer A, vol. 70, pp. 1271–1396, 2016.
DOI:10.1080/10407782.2016.1230433.
[19] Y. Varol, H. F. Oztop, and I. Pop, “Numerical analysis of natural convection for a porous rectangular enclosure
with sinusoidally varying temperature profile on the bottom wall,” Int. Commun. Heat Mass Transfer, vol. 35,
pp. 56–64, 2007. DOI:10.1016/j.icheatmasstransfer.2007.05.015.
[20] H. F. Oztop, “Natural convection in partially cooled and inclined porous rectangular enclosures,” Int. J. Thermal
Sci., vol. 46, pp. 149–156, 2007. DOI:10.1016/j.ijthermalsci.2006.04.009.
[21] S. Sivasankaran, Y. Do, and M. Sankar, “Effect of discrete heating on natural convection in a rectangular porous
enclosure,” Trans. Porous Med., vol. 86, pp. 261–281, 2011. DOI:10.1007/s11242-010-9620-x.
[22] S. K. Pandit and A. Chattopadhyay, “Higher order compact computations of transient natural convection in a
deep cavity with porous medium,” Int. J. Heat Mass Transfer, vol. 75, pp. 624–636, 2014. DOI:10.1016/j.
ijheatmasstransfer.2014.03.079.
[23] S. Bhardwaj and A. Dalal, “Effect of undulations on the natural convection heat transfer and entropy generation
inside a porous right-angled triangular enclosure,” Numer. Heat Transfer A, vol. 67, pp. 972–991, 2015.
DOI:10.1080/10407782.2014.949152.
[24] S. Bhardwaj, A. Dalal and S. Pati, Influence of wavy wall and non-uniform heating on natural convection heat
transfer and entropy generation inside porous complex enclosure,” Energy, vol.79, pp. 467–481, 2014.
DOI:10.1016/j.energy.2014.11.036.
[25] Y. Varol, H. Oztop, M. Mobedi, and I. Pop, “Visualization of natural convection heat transport using
heatline method in porous non-isothermally heated triangular cavity,” Int. J. Heat Mass Transfer, vol. 51,
pp. 5040–5051, 2008. DOI:10.1016/j.ijheatmasstransfer.2008.04.023.
[26] M. Zeng, P. Yu, F. Xu, and Q. W. Wang, “Natural convection in triangular attics filled with porous medium
heated from below,” Numer. Heat Transfer A, vol. 63, pp. 735–754, 2013. DOI:10.1080/10407782.2013.756702.
[27] A. C. Baytas and I. Pop, “Free convection in oblique enclosures filled with a porous medium,” Int. J. Heat Mass
Transfer, vol. 42, pp. 1047–1057, 1999. DOI:10.1016/s0017-9310(98)00208-7.
[28] E. Baez and A. Nicolas, “2D natural convection flows in tilted cavities: Porous media and homogeneous fluids,”
Int. J. Heat Mass Transfer, vol. 49, pp. 4773–4785, 2006.
[29] Q. W. Wang, M. Zeng, Z. P. Huang, G. Wang, and H. Ozoe, “Numerical investigation of natural convection in
an inclined enclosure filled with porous medium under magnetic field,” Int. J. Heat Mass Transfer, vol. 50,
pp. 3684–3689, 2007. DOI:10.1016/j.ijheatmasstransfer.2007.01.045.
[30] G. Wang, Q. Wang, M. Zeng, and H. Ozoe, “Numerical study of natural convection heat transfer in an inclined
porous cavity with time periodic boundary conditions,” Trans. Porous Med., vol. 74, pp. 293–309, 2008.
[31] S. E. Ahmed, H. F. Oztop, and K. Al-Salem, “Natural convection coupled with radiation heat transfer in an
inclined porous cavity with corner heater,” Comput. Fluids, vol. 102, pp. 74–84, 2014. DOI:10.1016/j.
compfluid.2014.06.024.
[32] A. C. Baytas and I. Pop, “Natural convection in a trapezoidal enclosure filled with a porous medium,” Int. J. Eng.
Sci. vol. 39, pp. 125–134, 2001.
[33] T. Basak, S. Roy, A. Singh, and A. R. Balakrishnan, “Natural convection flows in porous trapezoidal enclosures
with various inclination angles,” Int. J. Heat Mass Transfer, vol. 52, pp. 4612–4623, 2009. DOI:10.1016/j.
ijheatmasstransfer.2009.01.050.
[34] E. Natarajan, T. Basak, and S. Roy, “Natural convection flow in a trapezoidal enclosure with uniform and
non-uniform heating of bottom wall,” Int. J. Heat Mass Transfer, vol. 51, pp. 747–756, 2008. DOI:10.1016/j.
ijheatmasstransfer.2007.04.027.
[35] H. Saleh, R. Roslan, and I. Hashim, “Natural convection in a porous trapezoidal enclosure with an inclined
magnetic field,” Comput. Fluids, vol. 47, pp. 155–164, 2011. DOI:10.1016/j.compfluid.2011.03.002.
[36] F. Moukalled and M. Darwish, “Natural convection heat transfer in a porous rhombic annulus,” Numer. Heat
Transfer A, vol. 58, pp. 101–124, 2010. DOI:10.1080/10407782.2010.497322.
[37] R. Anandalakshmi and T. Basak, “Heatline based thermal management for natural convection in porous rhombic
enclosures with isothermal hot side or bottom wall,” Energy Convers. Manage., vol. 67, pp. 287–296, 2013.
DOI:10.1016/j.enconman.2012.11.022.
[38] R. Anandalakshmi and T. Basak, “Heatline analysis for natural convection within porous rhombic cavities with
isothermal/nonisothermal hot bottom wall,” Ind. Eng. Chem. Res., vol. 51, pp. 2113–2132, 2012. DOI:10.1021/
ie2007856.
[39] R. Anandalakshmi and T. Basak, “Heat flow visualization for natural convection in rhombic enclosures due to
isothermal and non-isothermal heating at the bottom wall,” Int. J. Heat Mass Transfer, vol. 31, pp. 1325–1342,
2012. DOI:10.1016/j.ijheatmasstransfer.2011.09.006.
NUMERICAL HEAT TRANSFER, PART A 19

[40] Y. Shiina, K. Fujimura, T. Kunugi, and N. Akino, “Natural convection in a hemispherical enclosure heated from
below,” Int. J. Heat Mass Transfer, vol. 37, pp. 1605–1617, 1994. DOI:10.1016/0017-9310(94)90176-7.
[41] C. L. Chen and C. H. Cheng, “Buoyancy-induced flow and convective heat transfer in an inclined arc-shaped
enclosure,” Int. J. Heat Fluid Flow, vol. 23, pp. 823–830, 2002. DOI:10.1016/s0142-727x(02)00189-3.
[42] C. J. Kim, and S. T. Ro, “Heat transfer correlation for natural convection in a meniscus-shaped cavity and its
application to contact melting process,” Int. J. Heat Mass Transfer, vol. 39, pp. 2267–2270, 1996. DOI:10.1016/
0017-9310(95)00285-5.
[43] O. Aydin, and G. Yesiloz, “Natural convection in a quadrantal cavity heated and cooled on adjacent walls,” J. Heat
Transfer, vol. 133, pp. 052501–052507, 2011. DOI:10.1115/1.4003044.
[44] G. Yesiloz and O. Aydin, “Natural convection in an inclined quadrantal cavity heated and cooled on adjacent
walls,” Exp. Therm. Fluid Sci., vol. 35, pp. 1169–1176, 2011. DOI:10.1016/j.expthermflusci.2011.04.002.
[45] D. Sen, P. K. Bose, R. Panua, and A. Das, “Numerical analysis of laminar natural convection in a quadrantal cavity
with a hot bottom and cold curved walls,” Heat Transfer Res., vol. 46, no. 7, pp. 631–641, 2015. DOI:10.1615/
heattransres.2015004629.
[46] D. Sen, P. K. Bose, R. Panua, A. K. Das, and P. Sen, “Laminar natural convection study in a quadrantal cavity using
heater on adjacent walls,” Front. Heat Mass Transfer, vol. 4, no. 1, pp. 1–7, 2013. DOI:10.5098/hmt.v4.1.3005.
[47] P. K. Bose, D. Sen, R. Panua, and A. K. Das, “Numerical analysis of laminar natural convection in a quadrantal
cavity with a solid adiabatic fin attached to the hot vertical wall,” J. Appl. Fluid Mech., vol. 6, pp. 501–510, 2013.
[48] A. Bejan, Entropy Generation Minimization. Boca Raton: CRC Press, 1996.
[49] I. Zahmatkesh, “On the importance of thermal boundary conditions in heat transfer and entropy generation for
natural convection inside a porous enclosure,” Int. J. Thermal Sci., vol. 47, pp. 339–346, 2008. DOI:10.1016/j.
ijthermalsci.2007.02.008.
[50] T. Basak, R. S. Kaluri, and A. R. Balakrishnan, “Entropy generation during natural convection in a porous cavity:
Effect of thermal boundary conditions,” Numer. Heat Transfer A, vol. 62, pp. 336–364, 2012. DOI:10.1080/
10407782.2012.691059.
[51] G. G. Ilis, M. Mobedi, and B. Sunden, “Effect of aspect ratio on entropy generation in a rectangular cavity with
differentially heated vertical walls,” Int. Commun. Heat Mass Transfer, vol. 35, pp. 696–703, 2008. DOI:10.1016/j.
icheatmasstransfer.2008.02.002.
[52] T. Basak, P. Gunda, and R. Anandalakshmi, “Analysis of entropy generation during natural convection in porous
right-angled triangular cavities with various thermal boundary conditions,” Int. J. Heat Mass Transfer, vol. 55,
pp. 4521–4535, 2012. DOI:10.1016/j.ijheatmasstransfer.2012.03.061.
[53] Y. Varol, H. F. Oztop, and I. Pop, “Entropy generation due to natural convection in non-uniformly heated porous
isosceles triangular enclosures at different positions,” Int. J. Heat Mass Transfer, vol. 52, pp. 1193–1205, 2009.
DOI:10.1016/j.ijheatmasstransfer.2008.08.026.
[54] V. M. Rathnam, P. Biswal, and T. Basak, “Analysis of entropy generation during natural convection within
entrapped porous triangular cavities during hot or cold fluid disposal,” Numer. Heat Transfer A, vol. 69,
pp. 931–956, 2016. DOI:10.1080/10407782.2015.1109362.
[55] O. C. Zienkiewicz and R. L. Taylor, The Finite Element Method, 4th ed. London: McGraw-Hill, 1991.
[56] G. Lauriat and V. Prasad, “Non-darcian effects on natural convection in a vertical porous enclosure,” Int. J. Heat
Mass Transfer, vol. 32, pp. 2135–2148, 1989. DOI:10.1016/0017-9310(89)90120-8.

Das könnte Ihnen auch gefallen