Mueller 2006

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Applied Thermal Engineering 26 (2006) 1662–1668

www.elsevier.com/locate/apthermeng

Prediction of the temperature in a fin cooled by natural


convection and radiation
D.W. Mueller Jr., H.I. Abu-Mulaweh *

Department of Mechanical Engineering, Purdue University at Fort Wayne, Fort Wayne, IN 46805-1499, USA

Received 1 September 2005; accepted 3 November 2005


Available online 18 January 2006

Abstract

Predictions and measurements of the temperature along a fin cooled by natural convection and radiation are reported. The physical
situation considered is a horizontal fin with a cylindrical cross-section. One end of the fin is maintained at a constant elevated temper-
ature, and the fin is sufficiently long so that heat loss from the tip is negligible. Heat is transferred by conduction along the fin and dis-
sipated from the surface via natural convection and radiation. The effect of natural convection is described with a published correlation
for a horizontal cylinder, and a simple model is used for the radiative heat transfer. A finite difference formulation that allows for variable
fluid property effects is used to determine the temperature distribution along the fin. A comparison is made to experimental results, and
the agreement between the model and experiment is very good. Results show that the heat loss due to radiation is typically 15–20% of the
total.
 2005 Elsevier Ltd. All rights reserved.

Keywords: Fins; Convection; Radiation; Finite difference; Cooling systems

1. Introduction drawing, metal and plastic extrusion, and glass fiber form-
ing. Modeling such problems is useful to assess whether
Extended surfaces (i.e., fins) are used to enhance the rate additional heat removal mechanisms, such as forced con-
of heat transfer from surfaces, especially in thermal engi- vection, are necessary.
neering applications where increasing the heat transfer The physical situation considered in this study is a hori-
coefficient is not an option. Thus, in thermal engineering zontal cylindrical rod or pin fin. One end of the fin is main-
applications where cooling is required, it is essential to tained at a constant elevated temperature, and the fin is
understand the basic mechanism of heat transfer in such sufficiently long so that heat loss from the tip is negligible.
surfaces. These heat transfer applications appear in cooling Heat is transferred by conduction along the fin and, in accor-
systems for electronic equipment, chemical processes and dance with common practice, the temperature at each axial
energy systems equipment, and high performance heat location is assumed to be uniform. Thus, the heat transfer is
exchangers. Owing to this fact, fins have been the topic one dimensional. Heat is removed from the surface of the fin
of many studies (see e.g., Refs. [1–3]) and a typical analysis via natural convection and radiation. One assumption,
is found in many textbooks, e.g., [4,5]. The prediction of which often is made to simplify the analysis, is that the heat
the temperature distribution along a cylindrical geometry transfer coefficient along the fin is constant. In the case of
cooled by natural convection and radiation is important natural convection and radiation, the heat transfer coeffi-
in many manufacturing operations such as tube and wire cient depends on the surface temperature. Thus, the assump-
tion of a constant heat transfer coefficient is technically not
correct as the temperature along the fin varies.
*
Corresponding author. Tel.: +1 260 481 6357; fax: +1 260 481 5728. The effect of a temperature dependent heat transfer coef-
E-mail address: mulaweh@engr.ipfw.edu (H.I. Abu-Mulaweh). ficient has been investigated in the past [6–12]. Most studies

1359-4311/$ - see front matter  2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.applthermaleng.2005.11.014
D.W. Mueller Jr., H.I. Abu-Mulaweh / Applied Thermal Engineering 26 (2006) 1662–1668 1663

Nomenclature

Ac cross-sectional area, m2 Q_ b rate of heat transfer into the base, W


As lateral surface area, m2 Q_ s rate of heat transfer from the surface, W
D diameter, m or mm RaD Rayleigh number, gb(T  T1)D3/ma
h total heat transfer coefficient, hc + hr, W/m2 K T temperature along the fin, C

h average total heat transfer coefficient, W/m2 K Tb base temperature, C
hc convection heat transfer coefficient, W/m2 K Tf mean film temperature, (T + T1)/2, C
hr radiation heat transfer coefficient, W/m2 K Tsurr surroundings temperature, C
g acceleration of gravity, m/s2 T1 ambient temperature, C
k thermal conductivity of fin, W/m K x axial coordinate, m
kf thermal conductivity of fluid, W/m K
L length, m Greek symbols
Leff effective length, m a thermal diffusivity, m2/s
m dimensionless fin parameter v dimensionless axial coordinate, x/L

m average dimensionless fin parameter e emissivity
NuD Nusselt number, hcD/kf m kinematic viscosity, m2/s
P perimeter, m r Stefan–Boltzmann constant, W/m2 K4
Pr Prandtl number, m/a h excess temperature, T  T1, C
Q_ rate of heat transfer, W

consider the case in which the heat transfer coefficients


have a power-law temperature dependence and the expo-
nent determines the heat loss mechanism, e.g., free convec-
tion, forced convection, boiling, or radiation. Unal [6]
outlined a procedure to find the temperature distribution
and fin efficiency in terms of elliptic integrals. Sen and
Trinh [7] obtained a general expression for the temperature
distribution in terms of a hypergeometric function and pre-
sented results for several exponents. A more complete set
of results was presented by Yeh and Liaw [8]. Studies of
this type were later applied to the problem of optimization
[9–12]. In these studies [6–12], no comparison was made to
experimental results, and the mathematical analysis was
quite involved.
The objective of the present study is to investigate the Fig. 1. Cylindrical pin fin and a typical differential element.
use of a local heat transfer coefficient that accounts for
both natural convection and radiation and to predict the
temperature in a long horizontal fin rod. The effect of nat-
ural convection is described with a published correlation steady-state operation, an energy balance applied to a dif-
for a horizontal cylinder, while a simple model is used ferential element yields:
for radiative heat transfer. A finite difference formulation Q_ x  Q_ xþdx  dQ_ loss ¼ 0; ð1Þ
is developed, and a MATLAB computer code is employed
that allows for variable fluid property effects. Grid resolu- where dQ_ loss accounts for heat transfer due to convection
tion and convergence issues associated with the numerical and radiation from the surface. With the use of standard
procedure are discussed. A comparison is made to experi- expressions for conduction, convection, and radiation,
mental results. the energy balance can be written as
 
d dT dAs dAs  4 
2. Governing equations kAc  hc ðT  T 1 Þ  er T  T 4surr ¼ 0.
dx dx dx dx
ð2Þ
Consider a cylindrical pin fin, as shown in Fig. 1, that
extends into a fluid of temperature T1. The base is main- In this study, a cylindrical pin fin with constant cross-
tained at constant temperature Tb. With the assumption sectional area is considered so that Ac = pD2/4 and
of one-dimensional heat conduction along the fin and dAs = P dx, where P = pD is the perimeter.
1664 D.W. Mueller Jr., H.I. Abu-Mulaweh / Applied Thermal Engineering 26 (2006) 1662–1668

Z L
In theory, hc, e, and k can vary along the length of the
fin. In this study, the thermal conductivity k and emissivity Q_ s ¼ hP ðT  T 1 Þ dx; ð9Þ
0
e are assumed to be independent of the temperature and
thus constant. The convection heat transfer coefficient is which is equal to the rate of heat conduction into the fin at
the base, i.e.,
assumed to depend on the local surface temperature. 
hc is related to the Nusselt number. For a horizontal, dT 
Q_ b ¼ kAc  . ð10Þ
isothermal cylinder, the Nusselt number correlation recom- dx x¼0
mended by Churchill and Chu [13] for pure natural convec-
tion is Next, with the introduction of the excess temperature
! h = T  T1 and the dimensionless coordinate v = x/L,
1=4
0:518 RaD Eqs. (5), (6) and (8) become
NuD ¼ 0:36 ¼ ; ð3Þ
9=16 4=9
½1 þ ð0:559=PrÞ  d2 h
 m2 h ¼ 0; ð11Þ
d2 v
which is valid for 106 < RaD < 109. All thermodynamic
properties of the fluid evaluated at the film temperature, hjv¼0 ¼ hb ; ð12Þ
i.e., Tf = (T + T1)/2. Note that hc is a circumferentially and
averaged value, but to avoid confusion with the axially 
dh
averaged value, the term average will not be used. ¼ 0. ð13Þ
The heat loss term due radiation can be written in a dv v¼1
form similar to the convection term with the introduction The dimensionless parameter m is defined as
of the radiation coefficient [4], i.e., rffiffiffiffiffiffi
4h
hr ¼ erðT 2 þ T 2surr ÞðT þ T surr Þ. ð4Þ mL . ð14Þ
kD
Note that hr depends on the surface temperature, the sur- Note that m is a local value that varies along the length of
rounding temperature, and the fin surface (emissivity). In the fin. Eqs. (11)–(13) constitute a non-linear, two-point
this study, the surroundings are assumed to be completely boundary value problem.
absorbing, i.e., black and at the same temperature as the For the case of a uniform heat transfer coefficient, the
ambient air, i.e., Tsurr = T1. temperature distribution is given by [4,5]:
With the use of Eq. (4), the convection and radiation cosh ½m
 ð1  xÞ
loss terms can be combined into a single term, vis. hð vÞ ¼ hb ; ð15Þ
cosh m
d2 T 4h where
 ðT  T 1 Þ ¼ 0; ð5Þ
d2 x kD sffiffiffiffiffiffiffiffi
4h
where h is the total heat transfer coefficient that accounts  L
m . ð16Þ
for convection and radiation, i.e., h  hc + hr. kD
The temperature at the fin base is specified, so that the A bar over a quantity denotes average value over the
boundary condition at the fin base x = 0 is given by length. The rate of heat transfer conducted into the base
T jx¼0 ¼ T b . ð6Þ and from the surface of the fin is given by
pffiffiffiffiffiffiffiffiffiffiffiffi
An energy balance at the fin tip requires that the heat trans- Q_ b ¼ hb hkPAc tanh m ¼ Q_ s ; ð17Þ
fer conducted to the surface is equal to the heat transfer
which can be found either by differentiation or integration
from the tip due to convection and radiation. Thus, the
of Eq. (15).
boundary condition at the fin tip x = L is given by
 For a long fin (L ! 1) with a uniform heat transfer
dT  4 coefficient, the temperature distribution in Eq. (15) can
kAc  ¼ hc;tip Ac ðT jx¼L  T 1 Þ þ etip rAc ðT jx¼L  T 4surr Þ.
dx x¼L
be simplified to
ð7Þ hðvÞ ¼ hb emv

; ð18Þ
However, in this study, the fins under consideration are as- and the rate of heat transfer conducted into the base and
sumed to be sufficiently long so that the heat transfer out of from the surface of the fin is now given by
the fin tip is negligible which allows Eq. (7) to be simplified pffiffiffiffiffiffiffiffiffiffiffiffi
to Q_ b ¼ hkPAc hb ¼ Q_ s . ð19Þ

dT 
¼ 0. ð8Þ
dx x¼L 3. Finite difference equations

The total heat loss from the surface of the fin can be A standard finite difference scheme (see e.g., Ref. [1]) is
given by applied to Eqs. (11)–(13). The domain is discretized into N
D.W. Mueller Jr., H.I. Abu-Mulaweh / Applied Thermal Engineering 26 (2006) 1662–1668 1665

Fig. 2. Schematic for the finite difference model.


Fig. 3. Experimental setup.

nodal points with hi = h(vi). A schematic of the grid is Table 1


shown in Fig. 2. Experimental data for the fins
The equation for the excess temperature at a typical Fin D (mm) L (m) Tb (C) T1 (C)
interior node is given by
1 3.18 0.685 92.8 20.8
hiþ1 þ hi1 2 6.35 0.685 96.0 21.0
hi ¼ 2
; ð20Þ 3 9.53 0.90 89.8 20.9
2 þ m2 ðDvÞ
while an energy balance applied to the node at the tip of the
fin yields:
plate that serves as backing and support for the heated
hN 1
hN ¼ . ð21Þ plate structure.
2
1 þ m2 ðDvÞ =2 The aluminum plate can be heated and maintained at a
The rate of heat transfer from the fin can be found by inte- constant temperature by adjusting the level of electrical
gration over the surface, i.e. energy input to the heating pad using a variable voltage
transformer. The temperature of the aluminum plate is
Dv X
N 1
Dv measured by thermocouples that are inserted into the plate
Q_ s ¼ h1 PL h1 þ hi PL Dvhi þ hN PL hN ð22Þ from the backside. A portable thermocouple is also used to
2 2
iþ2 measure the ambient temperature.
Copper–constanan (Type T) thermocouples were
or by differentiation and Fourier’s Law at the base, i.e. inserted into the rod to measure the axial temperature dis-
 tribution. Thermal conductive adhesive was used to seal
kAc h3 þ 4h2 þ 3h1
Q_ b ¼  . ð23Þ and secure the measuring junction of the thermocouple
L 2 Dv
into the small hole in fin rod. The thermocouples were con-
nected to a digital readout, and once steady-state condi-
4. Experimental temperature measurement tions were achieved, the temperature along the fin was
recorded. Base and ambient temperatures for three fins of
The temperature along the fins was measured using different diameter are given in Table 1.
the experimental apparatus described in Ref. [14]. The fins used in this study were made from Al 2024-T4.
The apparatus consisted of a heating plate mounted The thermal conductivity and the emissivity of the fins were
on a stand to which the rod was attached as shown in assumed to be independent of temperature and position
Fig. 3. and known. The thermal conductivity of the aluminum
The heated aluminum plate was made of four composite alloy was taken to be 120 W/m K.1 The emissivity of the
layers that were held together by screws. The first layer was aluminum rods was estimated to be 0.35 using an Omega-
an aluminum plate (15.24 cm · 15.24 cm, and 0.95 cm scope 3000 Infrared Pyrometer.
thick). The second layer consists of a heating pad that
can be controlled for electrical energy input. The third
layer is a 0.64 cm thick Transite insulating material. The
1
bottom layer of the heated plate is a 1.6 cm thick aluminum Thermal conductivity of Al 2024-T4 found with www.matweb.com.
1666 D.W. Mueller Jr., H.I. Abu-Mulaweh / Applied Thermal Engineering 26 (2006) 1662–1668

5. Numerical solution Table 3


Comparison of rate of heat transfer (W) into the base and across the
surface of the fin
A MATLAB code was written to solve the finite differ-
ence model described above. The algorithm was con- N Fin 1 Fin 2 Fin 3
structed as follows: Q_ b Q_ s Q_ b Q_ s Q_ b Q_ s
21 0.8377 0.9287 2.3051 2.4148 3.6444 3.8198
1. Input all geometry (D and L) and material constants (k 41 0.8809 0.9071 2.3608 2.3910 3.7334 3.7817
and e) for the fin. 81 0.8946 0.09016 2.3771 2.3850 3.7594 3.7720
161 0.8984 0.9003 2.3814 2.3835 3.7664 3.7696
2. Input base and ambient temperatures (Tb and T1). 321 0.8995 0.8999 2.3826 2.3831 3.7682 3.7690
3. Calculate hc and hr at the base temperature and compute 641 0.8997 0.8998 2.3829 2.3830 3.7687 3.7689
a value of m at each node.
4. Calculate the excess temperature along the fin
using Gauss–Seidel iteration. Iteration was termi-
nated when the difference between successive values of 6. Results and discussion
the excess temperature at each node is less than
1 · 1014. In the standard analysis of fins, the fin parameter m is
5. Calculate new values of hc, hr, and m at each node with useful to interpret the results. However, in this study m
the updated temperatures. (actually h) is not specified and varies along the fin. h
6. Repeat Steps 4 and 5 until convergence is achieved. depends on the fin diameter, the surface emissivity, the
fin temperature, and the ambient temperature. A quantita-
The properties of air were calculated using a ‘‘table- tive discussion of the variation of h along the fin will be
lookup’’ interpolation scheme with data from Ref. [4]. given later, but at this point it is helpful to understand
Data was input at temperatures from 250 K to 450 K. A the functional form of h to assist in the interpretation of
spline function from MATLAB was used to determine the results. With Eq. (3), the definition of the Nusselt num-
properties at intermediate temperatures. ber, and Eq. (4), h can be written as
A numerical investigation was performed to study the
fins described in Table 1. First, the finite difference portion 0:36k f F ðT f Þh1=4  
h¼ þ 1=4
þ er T 2 þ T 21 ðT þ T 1 Þ; ð24Þ
of the code was validated for the case of Fin 2 with con- D D
stant h = 12.5 W/m2 K. For constant h an analytical solu-
where
tion (Eqs. (15) and (17)) is possible. As the number of
nodes increased, the numerical solution, found using Eqs. 0:518k f ðgb=maÞ1=4
(20) and (21), converges to the exact (analytical) solution, F ðT f Þ ¼ h i4=9 . ð25Þ
as can be seen in Table 2. Also, note that rate of heat con- 1 þ ð0:559=PrÞ9=16
ducted into the base of the fin (Eq. (23)) approaches the
rate of heat lost from the lateral surface (Eq. (22)) as the The first two terms in Eq. (24) are from hc and the third
number of nodes is increased. term is simply hr. F(Tf) is a grouping of the fluid properties
Next, to verify the proper treatment of the variable heat evaluated at the film temperature. For the range of temper-
transfer coefficient, a comparison was made between the atures considered, F is nearly constant. Note each of the
rate of heat conducted into the fin and rate of heat lost three terms in Eq. (24) increases as the surface temperature
from the fin surface. Results are shown in Table 3. As of the rod increases, and thus h increases. Also, for a given
the number of nodes increased, rate of heat conducted into surface temperature, as D increases, h decreases.
the base of the fin (Eq. (23)) approaches the rate of heat Fig. 4 shows the comparison of the dimensionless excess
lost from the lateral surface (Eq. (22)). It should be noted temperature along the fin as computed by the finite differ-
that all subsequent computer runs were made with ence solution and from the experimental data. Note that
N = 641. from Table 1 the excess temperatures at the bases of all
three fins are similar (70 C). As expected, the tempera-
Table 2
ture from both the experimental data and the numerical
Convergence and validation for Fin 2 with h = 12.5 W/m2 K solution is a maximum at the base and decays to nearly
N h(v)/hb Q_ b (W) Q_ s (W)
the ambient temperature at the tip. The temperature distri-
bution depends on the fin geometry (D and L), the fin
v = 0.25 v = 0.5 v = 0.75
material (k and e) and the temperatures (Tb, T1, and Tsurr).
21 0.250937 0.063188 0.016784 2.2535 2.3308 Agreement between the experiment and numerical solution
41 0.250109 0.062769 0.016612 2.2936 2.3143
81 0.249901 0.062664 0.016569 2.3048 2.3101
is very good. As is shown in the figure, the larger the diam-
161 0.249849 0.062638 0.016558 2.3077 2.3091 eter, the longer the fin rod needs to be for the temperature
321 0.249836 0.062631 0.016556 2.3085 2.3088 at the tip to reach the ambient temperature. This is
641 0.249832 0.062630 0.016555 2.3087 2.3088 because, for the same base excess temperature, larger diam-
Exact 0.249831 0.062629 0.016555 2.3087 2.3087 eters are associated with more heat conduction into the fin
D.W. Mueller Jr., H.I. Abu-Mulaweh / Applied Thermal Engineering 26 (2006) 1662–1668 1667

1.0 4
Fin 1
3.5
0.8 Fin 1
measurement
3 Fin 2
finite difference
0.6 2.5 Fin 3

Q cond (W)
θ/θb

2
0.4
1.5

0.2 1

0.5
0.0
0 0.2 0.4 0.6 0.8 1
x/L 0
0 0.2 0.4 0.6 0.8 1
1.0 x/L
Fin 2
Fig. 5. Heat transfer due to conduction along the fin.
0.8
measurement
finite difference
0.6
θ/θb

Table 4
0.4 Effective lengths and a comparison of heat loss due to convection and
radiation for the three fins

0.2
Fin Leff (m) hc (W/ hr (W/ Q_ loss (W)
m2 K) m2 K)
Base Tip Base Tip Convection Radiation
0.0
0 0.2 0.4 0.6 0.8 1
1 0.39 15.2 5.51 2.88 2.02 0.756 0.144
x/L
2 0.58 11.8 5.32 2.93 2.03 1.906 0.477
1.0
3 0.75 10.1 4.37 2.84 2.03 2.925 0.842
Fin 3

0.8
measurement
finite difference
0.6 Note that if the rod is longer than Leff no additional heat
θ/θb

loss will occur from the extra length.


0.4 The variation of the heat transfer coefficients along the
fins is shown in Fig. 6. The heat transfer coefficients are a
0.2 maximum at the base and decrease along the fin. This is
expected because the temperature along the fin decreases.
0.0
The maximum and minimum heat transfer coefficients
0 0.2 0.4 0.6 0.8 1
due to convection and radiation are given in Table 4.
x/L
The effect of the fin on hc is clearly shown with the values
Fig. 4. Temperature distribution along the fins. at the base, where the excess temperatures for all three fins
are similar. Near the tip, the effect of D is less evident. This
and lower values of h. Note that Fin 3 is physically longer is due to the fact that excess temperature near the tip is
than Fins 1 and 2. greater for Fin 3 than for Fin 1, and the effect of the tem-
The axial rate of heat conducted along the fin is pre- perature reduces the effect of the diameter. The Rayleigh
sented in Fig. 5 for the three different fins. As can be seen number for the three cases varies from a maximum of
from the figure, for a given fin, the rate of heat conducted 3572 at the base of the 9.53 mm-diameter fin to 0.428 at
along the fin is a maximum at the base and decays to zero the tip of the 3.18 mm-diameter fin. These values are well
at the tip. The figure also shows that at any given axial within the limits in Eq. (3). The heat transfer coefficient
location, the rate of heat conducted increases as the fin due to radiation does not depend on the fin diameter and
diameter increases. is therefore similar for all three fins.
In this study, all three fins can be classified as ‘‘long’’ by Note that the heat transfer coefficients as predicted by
inspection of the temperature and heat transfer distribu- Eqs. (3) and (4) do not go to zero at the tip of the fin. How-
tion. However, a more quantitative approach to the classi- ever, the heat transfer goes to zero because the temperature
fication can be made with introduction of an effective difference goes to zero. As an extreme example, consider
length, Leff. Leff is defined as the length of the fin from the case when T = T1 = 20 C. According to Eq. (3),
which 99% of the heat is lost. Effective lengths for the three hc ! 0.009288/D and with D = 0.00635 m, hc ! 1.46 W/
fins are given in Table 4. As expected, Leff increases with D. m2 K. According to Eq. (4), hr ! 2.0 W/m2 K.
1668 D.W. Mueller Jr., H.I. Abu-Mulaweh / Applied Thermal Engineering 26 (2006) 1662–1668

20 7. Concluding remarks
18
Fin 1
16 A published correlation for natural convection and a
14 simple model for radiative heat transfer have been used
to predict the temperature along a horizontal fin. The
h (W/m K)

12 h
agreement between the finite difference solution and exper-
2

10
imental results is good.
8 hc
Despite the relatively simple model proposed, the actual
6
physical situation is rather complicated. Effects such as (1)
4 hr currents in the air which indicate forced (or mixed) convec-
2 tion, (2) an interaction between the flow around the mount-
0
0 0.2 0.4 0.6 0.8 1
ing plate and the fin, and (3) radiative heat transfer
x/L between the mounting plate and the fin warrant additional
20 consideration.
18
Fin 2
16 References
14
[1] D.Q. Kern, A.D. Kraus, Extended Surface Heat Transfer, McGraw-
h (W/m K)

12 Hill, New York, 1972.


2

h
10 [2] A.D. Krause, A. Bar-Cohen, Design and Analysis of Heat Sinks,
Wiley, New York, 1995.
8
hc [3] P. Razelos, A critical review of extended surface heat transfer, Heat
6 Transfer Eng. 24 (6) (2003) 11–28.
4 [4] F.P. Incropera, D.P. Dewitt, Introduction to Heat Transfer, Wiley,
hr
2
New York, 1985.
[5] A.F. Mills, Heat Transfer, second ed., Prentice-Hall, New Jersey,
0 1999.
0 0.2 0.4 0.6 0.8 1
x/L [6] H.C. Unal, Determination of the temperature distribution in an
20 extended surface with non-uniform heat transfer coefficient, Int. J.
Heat Mass Transfer 28 (1985) 2279–2284.
18
Fin 3 [7] A.K. Sen, S. Trinh, An exact solution for the rate of heat transfer
16 from a rectangular fin governed by a power law-type temperature
14 dependence, J. Heat Transfer 108 (1986) 457–459.
[8] R.H. Yeh, S.P. Liaw, An exact solution for thermal characteristics of
h (W/m K)

12
fins with power-law heat transfer coefficients, Int. Commun. Heat
2

10 h Mass Transfer 17 (1990) 317–330.


8 [9] R.H. Yeh, Optimization of design parameters for spines of various
geometries, J. Mar. Sci. Technol. 3 (1995) 11–17.
6 hc
[10] R.H. Yeh, Optimum dimensions of longitudinal rectangular fins and
4 hr cylindrical pin fins with variable heat transfer coefficient, Can. J.
2 Chem. Eng. 74 (1996) 144–151.
[11] R.H. Yeh, An analytical study of the optimum dimensions of
0
0 0.2 0.4 0.6 0.8 1 rectangular fins and cylindrical pin fins, Int. J. Heat Mass Transfer 40
x/L (1997) 3607–3615.
[12] K. Laor, H. Kalman, Performance and optimum dimensions of
Fig. 6. Variation of the heat transfer coefficient along the fin.
different cooling fins with a temperature-dependent heat transfer
coefficient, Int. J. Heat Mass Transfer 39 (1996) 1993–2003.
Also shown in Table 4 is the rate of heat loss due to con- [13] S.W. Churchill, H.H.S. Chu, Correlating equations for laminar and
turbulent free convection from a horizontal cylinder, Int. J. Heat
vection and the rate of heat loss due to radiation. The heat
Mass Transfer 18 (1975) 1049–1053.
loss due to radiation ranges from 15% of the total for Fin 1 [14] H.I. Abu-Mulaweh, Integration a design of experiment in the heat
to 22% of the total for Fin 3. The heat loss due to radiation transfer laboratory, in: Proceedings of the ASEE Annual Conference,
is significant and must be included in the model. Nashville, TN, CD-ROM, Session 1426, 2003.

You might also like