Vous êtes sur la page 1sur 18

Int J Thermophys (2012) 33:924941

DOI 10.1007/s10765-012-1179-z
Analytical Solution for ConvectiveRadiative
Continuously Moving Fin with Temperature-Dependent
Thermal Conductivity
Mohsen Torabi Hessameddin Yaghoobi A. Aziz
Received: 21 July 2011 / Accepted: 24 February 2012 / Published online: 16 March 2012
Springer Science+Business Media, LLC 2012
Abstract In this article, heat transfer in a moving n with variable thermal con-
ductivity, which is losing heat by simultaneous convection and radiation to its
surroundings, is analyzed. The calculations are carried out by using the differen-
tial transformation method (DTM) which is an analytical solution technique that can
be applied to various types of differential equations. The effects of parameters such
as the Peclet number, Pe, thermal conductivity parameter, a, convectionconduction
parameter, N
c
, radiationconduction parameter, N
r
, dimensionless convection sink
temperature,
a
, and dimensionless radiation sink temperature,
s
, on the temperature
distribution are illustrated and explained. The analytical solution is found to be in good
agreement with the direct numerical solution. Moreover, the results demonstrate that
the DTM is very effective in generating analytical solutions for even highly nonlinear
problems.
Keywords Convectionradiation heat transfer Differential transformation method
Moving n Thermal analysis Variable thermal conductivity
M. Torabi H. Yaghoobi (B)
Young Researchers Club, Central Tehran Branch, Islamic Azad University, Tehran, Iran
e-mail: Yaghoobi.Hessam@gmail.com
A. Aziz
Department of Mechanical Engineering, School of Engineering and Applied Science,
Gonzaga University, Spokane, WA 99258, USA
123
Int J Thermophys (2012) 33:924941 925
List of Symbols
Variables
A Fin cross-sectional area (m
2
)
a Dimensionless thermal conductivity
c
p
Specic heat of the material (J kg
1
K
1
)
D Domain
H Constant
h Convection heat transfer coefcient (W m
2
K
1
)
k(T) Temperature-dependent thermal conductivity (W m
1
K
1
)
k
a
Thermal conductivity at the base temperature (W m
1
K
1
)
L Fin length (m)
N
c
Dimensionless convectionconduction parameter
N
r
Dimensionless radiationconduction parameter
P Fin perimeter (m)
Pe Peclet number
T Temperature (K)
T
a
Sink temperature for convection (K)
T
b
Fins base temperature (K)
T
s
Sink temperature for radiation (K)
U Speed of moving n (m s
1
)
X Dimensionless axial distance measured from the tip of the n
X(k) Transformed analytical function
x Axial distance measured from the tip of the n (m)
x(t ) Original analytical function
Greek Symbols
Thermal diffusivity of the material (m
2
s
1
)
Slope of the thermal conductivitytemperature curve (K
1
)
Emissivity
Dimensionless temperature

a
Dimensionless convection sink temperature

b
Dimensionless ns base temperature

s
Dimensionless radiation sink temperature
Density of material (kg m
3
)
StefanBoltzmann constant
1 Introduction
Extended surfaces are extensively used in various industrial applications. A compre-
hensive review on this topic is presented in a treatise by Kraus et al. [1]. The assump-
tions of constant thermophysical properties and uniform heat transfer coefcient
reduce the mathematical complexity of the energy equation and allow closed-form
analytical solutions for a number of cases as documented in Kraus et al. [1]. If a large
temperature difference exists within a n, the thermal conductivity varies from the
123
926 Int J Thermophys (2012) 33:924941
base to the tip of n, the variation being dependent on the material of the n. In real
operating conditions, the heat transfer coefcient also varies along a n. The variation
may be a function of the spatial coordinate along the n or the local temperature dif-
ference between the n surface and the surrounding uid. A brief review of published
work that is of immediate relevance to this article is now presented.
Sharqawy and Zubair [2] carried out an analysis to study the efciency of straight
ns with different congurations when subjected to simultaneous heat and mass trans-
fer mechanisms. Domairry and Fazeli [3] solved the nonlinear straight n differential
equation by the homotopy analysis method (HAM) to evaluate the temperature distri-
bution within a n. Arslanturk [4] developed correlation equations for the optimum
design of annular ns with a temperature-dependent thermal conductivity. Kulkarni
and Joglekar [5] proposed and implemented a numerical technique based on residue
minimization to solve the nonlinear differential equation governing the temperature
distribution in a straight convective n having a temperature-dependent thermal con-
ductivity. Khani et al. [6] used HAMto derive approximate analytical solutions for the
temperature distribution and efciency of a convective n with simultaneous varia-
tions of the thermal conductivity and heat transfer coefcient with temperature. Kundu
[7] described an analytical method for obtaining the performance characteristics of an
annular step n (ASF) with simultaneous surface heat and mass transfer. Fouladi et
al. [8] utilized the variational iteration method as an approximate analytical method
to overcome some inherent limitations arising as uncontrollability to the nonzero end-
point boundary conditions and used this method to solve some examples in the eld
of heat transfer. Khani and Aziz [9] used HAM to develop an analytical solution for
the thermal performance of a straight n of a trapezoidal prole when both the ther-
mal conductivity and heat transfer coefcient are temperature dependent. Recently,
thermal analyses of a convectiveradiative moving n with a temperature-dependent
thermal conductivity were conducted [10, 11].
The differential transformation method (DTM) is a semi-numerical-analytical
method. The DTM which is based on the Taylor series expansion was rst proposed
by Zhou [12] in 1986 for the solution of linear and nonlinear initial value problems
that appear in the analysis of electrical circuits. This method proposes a solution in
the form of a polynomial as will be seen in Sect. 3, where the concept of a differ-
ential transform is briey described. The method was subsequently used to obtain
analytical solutions of higher-order ordinary differential equations, partial differential
equations, difference equations, and integro-differential equations [1321]. Chu and
Chen [13] applied a hybrid method of a differential transform and nite difference
method to solve a transient heat conduction problem which had complex nonlinear
terms. Chu and Lo [14] applied the differential transformation technique for trans-
forming and discretizing the governing equations as well as the boundary conditions
arising in two other problems in nonlinear transient heat conduction. Lo and Chen [15]
proposed an alternative numerical method to investigate hyperbolic heat conduction
problems using the hybrid differential transfer/control-volume method. Joneidi et al.
[16] used the DTM to obtain an analytical solution for the efciency of a convec-
tive straight n with a temperature-dependent thermal conductivity. Jang et al. [17]
investigated a two-dimensional transient heat conduction problem with discontinuous
boundary and initial conditions. Rashidi et al. [18] applied the DTMto nd the analytic
123
Int J Thermophys (2012) 33:924941 927
solution for the problem of mixed convection about an inclined at plate embedded
in a porous medium. Rashidi and Mohimanian Pour [19] applied the DTM to steady
ow over a rotating disk in a porous medium with heat transfer. Kundu and Barman
[20] proposed the DTM to determine the temperature eld in wet ns of rectangular
and triangular geometries. Yaghoobi and Torabi [21] applied the DTM to solve the
problems of convective and convectiveradiative cooling of a lumped system with a
temperature-dependent specic heat.
The purpose of this article is to demonstrate the usefulness of the DTM to solve
problem of convectiveradiative heat transfer from a continuously moving n with a
temperature-dependent thermal conductivity. This problem is of fundamental impor-
tance in the thermal processing of materials [10, 11, 22]. Aziz and Khani [10] solved
the problem using the HAM with 20 terms of the series, while Aziz and Lopez [11]
utilized the numerical algorithm built into Maple 14 to generate numerical results. It
will be shown that with the DTM, a highly accurate analytical solution of the prob-
lem is achievable with only six easily computable terms. The accuracy of the DTM
will be demonstrated by comparing the DTM solution with a direct numerical solu-
tion [11]. The results to be presented will highlight the effects of the Peclet number,
Pe, thermal conductivity parameter, a, convectionconduction parameter, N
c
, radi-
ationconduction parameter, N
r
, dimensionless convection sink temperature,
a
, and
dimensionless radiation sink temperature,
s
, on the temperature distribution in the
moving n.
2 Problem Formulation
Consider the thermal processing of a plate or a rod of cross-sectional area A and
perimeter P as it moves horizontally with a constant speed U as shown in Fig. 1.
The hot plate or rod emerges from a die or furnace or an intermediate station at a
constant temperature T
b
. The motion of the plate or rod may induce a ow eld in an
otherwise quiescent surrounding medium or alternatively, the plate or rod may experi-
ence an externally driven ow over its surface. Either way, the plate or rod is exposed
to a colder surrounding medium and loses heat by convection and radiation. The
convection
q
radiation
q
0
d
d
=
x
T
Fig. 1 Geometry of a moving n
123
928 Int J Thermophys (2012) 33:924941
radiative component would play a more prominent role if the forced convection is
weak or absent or when only natural convection occurs. The convection and radiation
sink temperatures are taken to be different, and one can be varied independent of the
other. The surface of the moving material is assumed to be gray with a constant emissiv-
ity . The convective owin the surrounding mediumprovides a constant heat transfer
coefcient h over the entire surface of the moving material. As the material undergo-
ing the treatment experiences a large change in its temperature during the process, the
change in its thermal conductivity may be as much as 100%. For most materials, the
thermal conductivity of the material k increases linearly with temperature. Thus, we
assume
k(T) = k
a
[1 + (T T
a
)] (1)
where k
a
is the thermal conductivity of the n at ambient temperature T
a
and is
the slope of the thermal conductivitytemperature curve. It is assumed that the initial
transient, which occurs when the material is exposed to the coolant, dies out quickly
and the processing takes place under steady-state conditions. This assumption has
been made in several studies [10, 11, 22].
The steady-state energy balance for the material moving with a constant speed and
losing heat by simultaneous convection and radiation may be written as
d
dx
_
k(T)
dT
dx
_

hP
A
(T T
a
)
P
A
_
T
4
T
4
s
_
+
1

U
dT
dx
= 0 (2)
where = k
a
/(c
p
) is the thermal diffusivity of the material, is the density, and c
p
is the specic heat. The last term on the left in Eq. 2 is the advection term. The axial
coordinate x is measured from the tip of the moving n.
Introducing the following dimensionless parameters,
=
T
T
b

a
=
T
a
T
b

s
=
T
s
T
b
L

=
PL
A
X =
x L

L
N
c
=
h A
Pk
a
a = T
b
N
r
=
AT
3
b
Pk
a
Pe =
U A
P
(3)
Assuming L

= 1, the formulation of the moving n problemreduces to the following


equation:
d
2

dX
2
+a
d
2

dX
2
+a
_
d
dX
_
2
a
a
d
2

dX
2
N
c
(
a
) N
r
(
4

4
s
)
+ Pe
d
dX
= 0, 0 X 1 (4)
with the following boundary conditions:
123
Int J Thermophys (2012) 33:924941 929
d
dX

X=0
= 0 (5a)
|
X=1
= 1 (5b)
as measured from the point where the thermal processing terminates (see Fig. 1).
3 Fundamentals of DTM
Let x(t ) be analytic in a domain D and let t = t
i
represent any point in D. The function
x(t ) is then represented by one power series whose center is located at t
i
. The Taylor
series expansion function of x(t ) is in the form of
x(t ) =

k=0
(t t
i
)
k
k!
_
d
k
x(t )
dt
k
_
t =t
i
t D (6)
The particular case of Eq. 6 when t
i
= 0 is referred to as the Maclaurin series of x(t )
and is expressed as
x(t ) =

k=0
(t )
k
k!
_
d
k
x(t )
dt
k
_
t =0
(7)
As explained in [23], the differential transformation of the function x(t ) is dened as
follows:
X (k) =
(H)
k
k!
_
d
k
x(t )
dt
k
_
t =0
(8)
where x(t ) is the original function and X(k) is the transformed function. The differen-
tial spectrumof X(k) is conned within the interval t [0, H], where H is a constant.
The differential inverse transform of X(k) is dened as follows:
x(t ) =

k=0
_
t
H
_
k
X (k) (9)
It is clear that the concept of differential transformation is based upon the Taylor series
expansion. The values of the function X(k) at values of argument k are referred to
as discretes, i.e., X(0) is known as the zero discrete, X(1) as the rst discrete, etc.
The more discretes available, the more precise it is possible to restore the unknown
function. The function x(t ) consists of the T-function X(k), and its value is given by
the sum of the T-function with (t /H)
k
as its coefcient. In real applications, with the
right choice of constant H and sufciently large values of argument k, the discretes
decrease rapidly and the contributions of the subsequent terms become negligible. The
123
930 Int J Thermophys (2012) 33:924941
Table 1 Fundamental
operations of differential
transform method
Original function Transformed function
x(t ) = f (t ) g(t ) X(k) = F(k) G(k)
x(t ) =
d f (t )
dt
X(k) = (k +1)F(k +1)
x(t ) =
d
2
f (t )
dt
2
X(k) = (k +1)(k +2)F(k +2)
x(t ) = t
m
X(k) = (k m) =
_
1
0
k = m
k = m
x(t ) = exp(t ) X(k) =

k
k!
x(t ) = f (t )g(t ) X(k) =
k

l=0
F(l)G(k l)
function x(t ) may thus be expressed by a nite series and Eq. 9 can be written as
x(t ) =
n

k=0
_
t
H
_
k
X (k) (10)
Mathematical operations performed by the differential transform method are listed in
Table 1.
4 Solution with DTM
Now we apply the DTM to Eq. 4. Taking the differential transform of Eq. 4 with
respect to X, and considering H = 1 according to Table 1, gives
(k +2) (k +1) (k +2) +a
_
k

l=0
(l) (k +2 l) (k +1 l) (k +2 l)
_
+a
_
k

l=0
(l +1) (l +1) (k +1 l) (k +1 l)
_
N
c
(k) N
r
_
k

m=0
(k m)
_
m

v=0
(m v)
_
v

w=0
(v w)(w)
___
+Pe (k +1) (k +1) +(N
c

a
+ N
r

4
s
)(k) = 0 (11)
From the boundary condition in Eq. 5a, that we have at point X = 0, and exerting a
transformation,
(1) = 0 (12)
The other boundary conditions are considered as follows:
(0) = C (13)
123
Int J Thermophys (2012) 33:924941 931
where C is a constant, and we will calculate it while considering another boundary
condition in Eq. 5b at point X = 1.
Accordingly, from a process of an inverse differential transformation, in this prob-
lem we calculated (k +2) from Eq. 11 as follows:
(2) =
1
2
N
c
C + N
r
C
4
N
c

a
+ N
r

4
s
1 +Ca a
a
(14a)
(3) =
1
6
Pe
_
N
c
C + N
r
C
4
N
c

a
+ N
r

4
s
_
(1 +Ca a
a
)
2
(14b)
(4) =
1
24
_
N
c
C + N
r
C
4
N
c

a
+ N
r

4
s
_
(1 +Ca a
a
)
3

_
2aN
C
c
aN
r
C
4
3aN
r

4
s
2aN
c

a
N
c
Pe
2
4N
r
C
3
+4aN
r
C
3

a
_
(14c)
(5) =
1
120
Pe
_
N
c
C + N
r
C
4
N
c

a
+ N
r

4
s
_
(1 +Ca a
a
)
3

_
11aN
c
C +5aN
r
C
4
13aN
r

4
s
11aN
c

a
2N
c
Pe
2
8N
r
C
3
+8aN
r
C
3

a
_
(14d)
The above process may be continued further. Substituting Eq. 14 into the main equation
based on the DTM, the closed-form of the solutions is obtained as
(X) = C +
1
2
N
c
C + N
r
C
4
N
c

a
+ N
r

4
s
1 +Ca a
a
X
2

1
6
Pe
_
N
c
C + N
r
C
4
N
c

a
+ N
r

4
s
_
(1 +Ca a
a
)
2
X
3

1
24
_
N
c
C + N
r
C
4
N
c

a
+ N
r

4
s
_
(1 +Ca a
a
)
3

_
2aN
C
c
aN
r
C
4
3aN
r

4
s
2aN
c

a
N
c
Pe
2
4N
r
C
3
+4aN
r
C
3

a
_
X
4
(15)
+
1
120
Pe
_
N
c
C + N
r
C
4
N
c

a
+ N
r

4
s
_
(1 +Ca a
a
)
3

_
11aN
C
c
+5aN
r
C
4
13aN
r

4
s
11aN
c

a
2N
c
Pe
2
8N
r
C
3
+8aN
r
C
3

a
_
X
5
+. . .
To obtain the value of C, we substitute the boundary condition fromEq. 5b into Eq. 15
giving
123
932 Int J Thermophys (2012) 33:924941
(1) = C +
1
2
N
c
C + N
r
C
4
N
c

a
+ N
r

4
s
1 +Ca a
a

1
6
Pe
_
N
c
C + N
r
C
4
N
c

a
+ N
r

4
s
_
(1 +Ca a
a
)
2

1
24
_
N
c
C + N
r
C
4
N
c

a
+ N
r

4
s
_
(1 +Ca a
a
)
3

_
2aN
C
c
aN
r
C
4
3aN
r

4
s
2aN
c

a
N
c
Pe
2
4N
r
C
3
+4aN
r
C
3

a
_
(16)
+
1
120
Pe
_
N
c
C + N
r
C
4
N
c

a
+ N
r

4
s
_
(1 +Ca a
a
)
3

_
11aN
c
C +5aN
r
C
4
13aN
r

4
s
11aN
c

a
2N
c
Pe
2
8N
r
C
3
+8aN
r
C
3

a
_
+. . . = 1
Equation16 can be solved by using the NewtonRaphson iterative technique to deter-
mine the unknown tip temperature C. Substituting for C into Eq. 15, we determine
(X).
The calculations reported in this article use n = 10 which was found to be sufcient
to give an accurate solution. Because Eq. 4 requires the summation of a limited num-
ber of terms, the analytical solution can be computed without excessive computational
effort.
5 Results and Discussion
For checking the accuracy of the DTM solutions, numerical solutions were generated
using the symbolic algebra package Maple which offers a highly accurate and exten-
sively tested fourth to fth order RungeKuttaFehlberg algorithm. The accuracy and
robustness of Maples algorithm for solving the boundary value problems has been
repeatedly conrmed for various problems [24, 25].
Tables 2 and 3 compare the n-tip temperature predicted by the DTM with the
numerical solution (NS) for some combination values of the convectionconduction
parameter N
c
, radiationconduction parameter N
r
, Peclet number Pe, dimensionless
convection sink temperature,
a
, dimensionless radiation sink temperature,
s
, and
the thermal conductivity parameter a. The agreement between DTM and NS results
is excellent. This level of agreement will also be seen in Figs. 2, 3, 4, 5, 6, and 7. In
preparing Figs. 2, 3, 4, 5, 6, and 7, only one parameter is varied keeping the other
parameters xed. Moreover, for all numerical results reported here, the following
values of variables were used unless otherwise indicated by the graphs or tables.
a = 0.5, N
c
= 0.25, N
r
= 0.75, Pe = 0.5,
a
= 0.4,
s
= 0.4
Figure 2 shows the effect of a variable thermal conductivity, i.e., parameter a on the
temperature distribution in the n. The bottom curve corresponds to a = 0 (constant
thermal conductivity), and the top curve corresponds to a = 1 (thermal conductivity
123
Int J Thermophys (2012) 33:924941 933
Table 2 Comparison of the DTM results with numerical solution (NS) for
a
=
s
= 0 for n-tip temper-
ature
Pe a N
c
0.25 0.5
N
r
N
r
0.25 0.5 0.25 0.5
DTM NS DTM NS DTM NS DTM NS
0.25 0 0.8358 0.8358 0.7944 0.7944 0.7652 0.7652 0.7340 0.7340
0.25 0.8584 0.8584 0.8190 0.8190 0.7955 0.7955 0.7645 0.7645
0.5 0.8756 0.8756 0.8384 0.8384 0.8192 0.8192 0.7890 0.7890
0.75 0.8892 0.8892 0.8541 0.8541 0.8381 0.8381 0.8089 0.8089
1 0.9001 0.9001 0.8670 0.8670 0.8536 0.8536 0.8255 0.8255
0.5 0 0.8447 0.8447 0.8040 0.8040 0.7774 0.7774 0.7461 0.7461
0.25 0.8650 0.8650 0.8264 0.8264 0.8049 0.8049 0.7741 0.7741
0.5 0.8808 0.8808 0.8443 0.8443 0.8266 0.8266 0.7967 0.7967
0.75 0.8933 0.8933 0.8589 0.8589 0.8441 0.8441 0.8153 0.8153
1 0.9034 0.9034 0.8710 0.8710 0.8585 0.8585 0.8308 0.8308
Table 3 Comparison of the DTM results with numerical solution (NS) for n-tip temperature

a
0.2 0.8

s

s
0.2 0.5 0.8 0.2 0.5 0.8
DTM NS DTM NS DTM NS DTM NS DTM NS DTM NS
0.8180 0.8180 0.8278 0.8278 0.8815 0.8816 0.8281 0.8281 0.8392 0.8392 0.8991 0.8991
at the base temperature is 50% higher than the thermal conductivity at the environ-
ment temperature). As the parameter a increases, the average thermal conductivity of
the material increases, and as expected, the result is a gradual increase in the local
temperature.
Figure 3 illustrates the effect of the Peclet number on the temperature distribution
in the n. As Pe increases, i.e., as the material moves faster, the time for which the
material is exposed to the environment gets shorter, the temperatures are consequently
higher, and their variation is more gradual than when the material moves at a slower
speed. The effect of the convectionconduction parameter N
c
on the temperature
distribution is shown in Fig. 4. As N
c
increases, i.e., as convection gets stronger,
the cooling becomes more effective, promoting lower temperatures in the material. In
Fig. 5, we illustrate the effect of the radiationconduction parameter N
r
on the temper-
ature distribution in the n. As the radiative transport becomes stronger, the radiative
cooling becomes more effective, which in turn causes the lowering of temperatures in
123
934 Int J Thermophys (2012) 33:924941
Fig. 2 Effect of variable thermal conductivity on temperature distribution obtained by the DTM(solid line)
and NS (circle)
Fig. 3 Effect of Peclet number on the temperature distribution obtained by the DTM (solid line) and NS
(circle)
123
Int J Thermophys (2012) 33:924941 935
Fig. 4 Effect of convectionconduction parameter on the temperature distribution obtained by the DTM
(solid line) and NS (circle)
Fig. 5 Effect of radiationconduction parameter on the temperature distribution obtained by the DTM
(solid line) and NS (circle)
123
936 Int J Thermophys (2012) 33:924941
Fig. 6 Effect of convection sink temperature on the temperature distribution obtained by the DTM (solid
line) and NS (circle)
Fig. 7 Effect of radiation sink temperature on the temperature distribution obtained by the DTM (solid
line) and NS (circle)
123
Int J Thermophys (2012) 33:924941 937
Fig. 8 Effect of Peclet number on the n-tip temperature
Fig. 9 Effect of convectionconduction parameter on the n-tip temperature
123
938 Int J Thermophys (2012) 33:924941
Fig. 10 Effect of radiationconduction parameter on the n-tip temperature
the n. The effect is similar to what is shown in Fig. 4 with respect to the strengthening
of the convective transport mechanism.
Figure 6 shows that as the convection sink temperature increases, i.e., as
a
increases, the convective heat loss from the moving n decreases which is reected in
increasingly higher temperatures in the n. Compared to the rest of the material, the
tip temperature which is the temperature at which the thermal processing is terminated
appears to be most sensitive to the changes in the convection sink temperature. This
would be an important consideration in selecting the appropriate processing time.
Figure 7 which serves as a companion to Fig. 6 displays the effect of the radia-
tion sink temperature
s
on the temperature distribution in the moving material. As
the radiation sink temperature increases, the radiative heat loss decreases, resulting
in higher temperatures in the material. A comparison of Figs. 6 and 7 shows that the
radiation sink temperature has a more profound effect on the temperature distribu-
tion than the convection sink temperature. In particular, the terminal (tip) temperature
rises significantly with an increase in the radiation sink temperature. This leads to an
important practical conclusion that when the material is cooled by simultaneous con-
vection and radiation, the radiation heat loss exerts a comparatively stronger inuence
on the thermal processing time and the terminal temperature at which the material is
withdrawn after processing.
123
Int J Thermophys (2012) 33:924941 939
Fig. 11 Effect of convection sink temperature on the n-tip temperature
Figure 8 is a plot of the tip (terminal) temperature as a function of the Peclet number
Pe for parametric values of the thermal conductivity parameter a. The tip temperature
increases as the Peclet number and the thermal conductivity parameter increase. The
explanations offered for Figs. 2 and 3 in combination explain the results of Fig. 8.
In Fig. 9, we have plotted the n-tip (terminal) temperature as a function of the
convectionconduction parameter N
c
for parametric values of the thermal conductiv-
ity parameter a. When convection is stronger, it results in lower n-tip temperatures.
The higher average thermal conductivity is associated with the increase in parameter
a increases the n-tip temperature. The results of Fig. 10 in which the n-tip temper-
ature is plotted as a function of the radiationconduction parameter N
r
for parametric
values of the thermal conductivity parameter a may be similarly interpreted.
Figure 11 shows how the tip (terminal) temperature is affected by the variations
in the convection sink temperature and thermal conductivity. For the case of constant
thermal conductivity (a = 0), the tip temperature increases sharply as the surface
convection weakens due to the increase in the convection sink temperature. However,
for materials exhibiting stronger thermal conductivity variation with temperature, the
effect of the convection sink temperature on the tip temperature gets attenuated as
can be seen from curves corresponding to higher values of the thermal conductivity
parameter a. The convection sink temperature
a
and the thermal conductivity param-
eter a appear to exert competing inuences on the tip temperature. For a = 0.75, the
tip temperature (0) shows virtually no sensitivity to the changes in the convection
123
940 Int J Thermophys (2012) 33:924941
Fig. 12 Effect of radiation sink temperature on the n-tip temperature
sink temperature
a
, and for a = 1, the tip temperature (0) shows a slight decrease
with increasing convection sink temperature
a
. Figure 12 shows how the tip (termi-
nal) temperature increases with an increase in the radiation sink temperature
s
and
thermal conductivity parameter a, and that these parameters exert a particularly strong
inuence on the tip temperature beyond
s
= 0.4.
It may be pointed out that other analytical approaches such as the homotopy pertur-
bation method and variational iteration method have been successfully used in a series
of papers by Jalaal et al. [2629] dealing with the motion of spherical and nonspherical
particles in viscous uids.
6 Conclusions
In this article, the DTM is briey described and then used to derive approximate
explicit analytical expressions for the temperature distribution in a moving n with
a temperature-dependent thermal conductivity and experiencing simultaneous con-
vectiveradiative surface heat loss. The DTM provided a more rapidly convergent
series solution (only 10 terms) compared with other methods such as the HAM which
requires a minimumof 20 terms [10]. As the surface convection and radiation increase
either individually or in tandem, the effect is to lower the n temperature. However,
as the n moves faster, i.e., as the Peclet number Pe increases, the n experiences
123
Int J Thermophys (2012) 33:924941 941
slower cooling. Similarly, as the average thermal conductivity of the n increases, i.e.,
the parameter a increases, it promotes slower cooling accompanied by higher local
n temperatures.
References
1. A.D. Kraus, A. Aziz, J.R. Welty, Extended Surface Heat Transfer (Wiley, New York, 2002)
2. M.H. Sharqawy, S.M. Zubair, Appl. Therm. Eng. 28, 2279 (2008)
3. G. Domairry, M. Fazeli, Commun. Nonlinear Sci. Numer. Simul. 14, 489 (2009)
4. C. Arslanturk, Heat Mass Transf. 45, 519 (2009)
5. D.B. Kulkarni, M.M. Joglekar, Appl. Math. Comput. 215, 2184 (2009)
6. F. Khani, M. Ahmadzadeh Raji, H. Hamedi Nejad, Commun. Nonlinear Sci. Numer. Simul. 14,
3327 (2009)
7. B. Kundu, Int. J. Heat Mass Transf. 52, 2646 (2009)
8. F. Fouladi, E. Hosseinzadeh, A. Barari, G. Domairry, Heat Transf. Res. 41, 155 (2010)
9. F. Khani, A. Aziz, Commun. Nonlinear Sci. Numer. Simul. 15, 590 (2010)
10. A. Aziz, F. Khani, J. Franklin Inst. 348, 640 (2011)
11. A. Aziz, R.J. Lopez, Int. J. Therm. Sci. 50, 1523 (2011)
12. J.K. Zhou, Differential Transform and Its Applications for Electrical Circuits (Huazhong University
Press, Wuhan, 1986)
13. H.P. Chu, C.L. Chen, Commun. Nonlinear Sci. Numer. Simul. 13, 1605 (2008)
14. H.P. Chu, C.Y. Lo, Numer. Heat Transf. Part A 53, 295 (2008)
15. C.Y. Lo, B.Y. Chen, Numer. Heat Transf. Part B 55, 219 (2009)
16. A.A. Joneidi, D.D. Ganji, M. Babaelahi, Int. Commun. Heat Mass Transf. 36, 757 (2009)
17. M.J. Jang, Y.L. Yeh, C.L. Chen, W.C. Yeh, Appl. Math. Comput. 216, 2339 (2010)
18. M.M. Rashidi, N. Laraqi, S.M. Sadri, Int. J. Therm. Sci. 49, 2405 (2010)
19. M.M. Rashidi, S.A. Mohimanian Pour, Afr. J. Math. Comput. Sci. Res. 3, 93 (2010)
20. B. Kundu, D. Barman, Int. J. Heat Fluid Flow 31, 722 (2010)
21. H. Yaghoobi, M. Torabi, Int. Commun. Heat Mass Transf. 38, 815 (2011)
22. M.V. Karwe, Y. Jaluria, Trans. ASME J. Heat Transf. 108, 728 (1986)
23. A.H. Hassan, Appl. Math. Comput. 154, 299 (2004)
24. A. Aziz, Heat Conduction with Maple (R.T. Edwards, Inc., Philadelphia, PA, 2006)
25. R.E. White, V.R. Subramanian, Computational Methods in Chemical Engineering with Maple Appli-
cations (Springer, Berlin, 2010)
26. M. Jalaal, D.D. Ganji, G. Ahmadi, Adv. Powder Technol. 21, 298 (2010)
27. M. Jalaal, D.D. Ganji, Adv. Powder Technol. 198, 82 (2010)
28. M. Jalaal, D.D. Ganji, Adv. Powder Technol. 22, 58 (2011)
29. M. Jalaal, D.D. Ganji, G. Ahmadi, Asia Pac. J. Chem. Eng. (2011). doi:10.1002/apj.492
123

Vous aimerez peut-être aussi