Anda di halaman 1dari 8

International Journal of Heat and Mass Transfer 53 (2010) 59885995

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


journal homepage: www.elsevier.com/locate/ijhmt

Technical Note

Thermal optimization of plate-n heat sinks with variable n thickness


Dong-Kwon Kim a,1, Jaehoon Jung b,1, Sung Jin Kim b,*
a
b

Department of Mechanical Engineering, Ajou University, Suwon 443-749, Republic of Korea


Department of Mechanical Engineering, Korea Advanced Institute of Science and Technology, Daejeon 305-701, Republic of Korea

a r t i c l e

i n f o

Article history:
Received 14 April 2010
Received in revised form 10 July 2010
Accepted 10 July 2010
Available online 20 August 2010
Keywords:
Heat sink
Thermal optimization
Variable n thickness

a b s t r a c t
In the present paper, thermal optimization of a plate-n heat sink with the n thickness varying in the
direction normal to the uid ow was conducted. The model used for this optimization was based on
the volume averaging theory (VAT). It was shown that the thermal resistance of the plate-n heat sink
can be reduced by allowing the n thickness to increase in the direction normal to the uid ow. In
the case of a water-cooled heat sink, the thermal resistance decreases by as much as 15%. The amount
of the reduction increases as the pumping power increases or as the length of the heat sink decreases.
2010 Elsevier Ltd. All rights reserved.

1. Introduction
Recent advances in semiconductor technology have led to a signicant increase in the power densities encountered in microelectronic equipment [1]. The possibility of a failure of an electronic
devices increases as the power density increases. Accordingly, a
higher level of performance in cooling technology is essential for
the reliable operation of electronic components [2]. Many ideas
pertaining to cooling methods have been proposed. Among the
various types of developed cooling systems, the plate-n heat sink
is the most widely used due to the benets of its simple design and
easy fabrication.
Several research works have concentrated on sizing-optimization of plate-n heat sinks. Several optimization methods have
been proposed based on the n model [3], the three-dimensional
numerical model [4], and on the volume averaging theory [5]. In
these studies, the layout of the heat sink was prescribed as shown
in Fig. 1a. The heat sink was optimized by determining the n
thickness and channel width which minimize the thermal resistance for a given heat sink size. This procedure was based on the
assumption that n thickness and channel width are constant
along the directions parallel and normal to the uid ow. However,
there is no guarantee that a rectangular cross-section n is the
most thermally efcient. It is likely that the thermal resistance of
a plate-n heat sink may be further reduced by allowing the n
thickness to vary in the directions parallel and normal to the uid
ow. Morega and Bejan demonstrated that the thermal resistance

* Corresponding author. Tel.: +82 42 350 3043; fax: +82 42 350 8207.
E-mail address: sungjinkim@kaist.ac.kr (S.J. Kim).
1
These authors contributed equally to this work.
0017-9310/$ - see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijheatmasstransfer.2010.07.052

of an air-cooled heat sink can be reduced by about 15% by increasing the n thickness in the ow direction [6]. However, there has
been no study focusing on heat sinks with n thickness varying
in the direction normal to the uid ow, to the authors knowledge.
Which type of heat sink performs better among the three heat
sinks shown in Fig. 1? The present paper is devoted to seeking a
proper answer to this question. Thermal optimization of the
plate-n heat sinks was conducted with n thicknesses varying
in the direction normal to the uid ow using a model based on
the volume averaging theory (VAT). The results demonstrated that
the thermal resistance can be reduced by allowing the n thickness
to increase in the direction normal to the uid ow, i.e., the heat
sink shown in Fig. 1c is the best choice. It will be shown later in this
paper that the reduction of the thermal resistance when using variable-thickness-ns increases as either the pumping power increases or the length of the heat sink decreases.
2. Mathematical formulation
The problem under consideration in the present paper concerns
forced convection through a plate-n heat sink as depicted in
Fig. 2a and b. The direction of the uid ow is parallel to the x
direction. The top surface is insulated and the bottom surface is
uniformly heated. Coolant passes through a number of channels
and takes heat away from a heat-dissipating electronic component
attached below. In analyzing the problem, for simplicity, the ow is
assumed to be laminar, incompressible, and both hydrodynamically and thermally fully-developed. All thermophysical properties
are assumed to be constant. In addition, the pumping power is
assumed to be constant. This condition implies that the power
required to drive the uid through the plate-n heat sink is given.

5989

D.-K. Kim et al. / International Journal of Heat and Mass Transfer 53 (2010) 59885995

Nomenclature
a
c
Dh,fr
f
h
H
k
kse
K
L
_
m
n
N
Nu
p
P
Ppump
Pr
q
q00
q00sf
R
Re
T
u
W

wetted area per volume


heat capacity of a uid
aspect ratio of the frontal area (2HW/(H+W))
friction factor
interstitial heat transfer coefcient based on the onedimensional bulk mean temperature
channel height
thermal conductivity
effective thermal conductivity of the solid
permeability
length of the heat sink
mass ow rate
number of channels
node number
Nusselt number (hDh/kf)
n pitch (wc+ww)
pressure
pumping power
Prandtl number
heat transfer rate
heat ux
heat ux from the ns to the uid
thermal resistance
Reynolds number
temperature
velocity
width of the heat sink

The channel width and the n thickness vary in the direction normal to the uid ow (wc = wc(y), ww = ww(y)), but the n pitch i.e.
the sum of the channel width and the n thickness is constant.
The momentum and energy equations for the fully developed
ow are given as follows:

!
@p
@2u @2u
l

;
@x
@y2 @z2

1
!

qc

@uT
@2T @2T
:
k

@x
@y2 @z2

In the model based on the volume averaging theory, the governing


equations for the averaged velocity and temperature are established
by averaging the momentum and energy equations in the z-direction over the analysis domain shown in Fig. 2c. When the solid conductivity is higher than the uid conductivity and the aspect ratio of
the channel is much higher than 1, the governing equations for the
averaged quantities and the appropriate boundary conditions are
given as follows [5]:

dp lf
ehuif ;
dx
K
f ;bm
f @hTi
eqf cf hui
hahTis  hTif ;bm ;
@x


@
@hTis
hahTis  hTif ;bm ;
kse
@y
@y


huif 0 at y 0; H;
s

hTi hTi
s

f ;bm

Tw

at y 0;

0 at y H;

channel width
n thickness
Cartesian coordinate system
averaged value
one-dimensional bulk mean temperature for the uid
phase

Greek symbols
a
aspect ratio of the channel (H/wc)
afr
aspect ratio of the frontal area (H/W)
e
porosity (wc/(ww+wc))
l
viscosity
q
density
Subscripts and superscripts
bm
bulk mean
cap
capacitive
conv
convective
uni
uniform-thickness-n heat sink
f
uid
fr
frontal area
i
node number
opt
optimized
s
solid
tot
total
var
variable-thickness-n heat sink

2
hui
wc
f

hTis

R ww wc =2

ww wc =2

udz;

f ;bm

hTi

ww =2

2
ww

Tudz
ww =2
;
R w
w wc =2
udz
ww =2

ww =2

T dz:

In Eq. (9), for the average uid temperature, the one-dimensional


bulk mean is employed in order to calculate the bulk mean temperature of the uid easily. e, a, kse, K, and h are the porosity, wetted
area per volume, effective thermal conductivity of the solid, permeability, and interstitial heat transfer coefcient. These can be represented as

wc
;
wc ww

K

ewc huif @u

2
;
wc ww
!1

h

@z zww =2

q00sf
s

hTi  hTif ;bm

70kf
:
17wc

kse

ks ww
;
wc ww

w3c
;
12wc ww
10

4
5
6
7

f ;bm

@hTi
@hTi

@y
@y

wc
ww
x,y,z
<>
<>b,f

where the average uid velocity, average uid temperature, and


average solid temperature are dened as shown below,
respectively:

where e, a, kse, K, and h are functions of y, because they are the functions of wc and ww.
The thermal performance of the heat sink was evaluated by
introducing the concept of thermal resistance. Thermal resistance
is dened here as the difference between the base temperature
of the heat sink at the outlet and the uid bulk mean temperature
at the inlet per unit of heat ow rate. The thermal resistance can be
decomposed into the capacitive resistance, which is responsible for
the temperature rise of the coolant from the inlet to the exit, and
the convective resistance, which is related to the heat transfer from
the ns to the coolant:

5990

D.-K. Kim et al. / International Journal of Heat and Mass Transfer 53 (2010) 59885995

In the present study, the thermal resistance is calculated from the


averaged velocity and temperature obtained by solving Eqs. (3)
(8) numerically. The n and the channel are discretized as shown
in Fig. 2c. Obviously, the averaged velocity and temperature depend
on the discretized channel width wc,i and n thickness ww,i where
i = 0, 1, . . . , N. Therefore the thermal resistance is a function of wc,i
and ww,i.
3. Results and discussion
In order to validate the solution presented in the previous section, thermal resistances obtained from the proposed model were
compared with results of a two-dimensional direct numerical simulation, the solutions of which were obtained by solving the
momentum equation and energy equation (Eqs. (1) and (2)) using
the control-volume-base nite difference method. Fig. 3a depicts
the thermal resistances for the uniform-thickness-n heat sink
and variable-thickness-n heat sink, whose porosities vary linearly
from 0.4 at the base to 0.6 at the top. Fig. 3b depicts the thermal
resistances for variable-thickness-n heat sinks whose porosities
vary as shown in Table 1. The thermal resistances obtained from
the proposed model are in good agreement with the numerical results with a relative error of less than 10%. However, it is shown
that the thermal resistance obtained from the proposed model
deviates from numerical results as the channel height decreases.
It is because the proposed model is based on the assumption that
the aspect ratio of the channel is much higher than 1. In addition,
the present results were also compared with Nusselt numbers results for rectangular and triangular channels given by Shah [7], as
shown in Fig. 4. The present results and the data from Shah match
with a maximum error of 10%.
To design an optimized heat sink, the n thickness and channel
width which minimize the thermal resistance should be determined. In the present study, the optimal values of wc,i and ww,,i
for which the thermal resistance is minimized for a given height,
length, width, and pumping power were numerically obtained
with the gradient descent method. The sequence of numerical simulation can be stated as

(a) Rectangular-fin heat sink

(b) Triangular-fin heat sink

(1) Start with the guessed channel width wc;i and n pitch p*.
(2) Calculate partial derivatives of Rtot with respect to wc,i and p.

@Rtot Rtot wc;i wc;i c  Rtot wc;i wc;i  c

;
@wc;i
2c

14

@Rtot Rtot p p c  Rtot p p  c


:

2c
@p

15

Here, c is a positive small number.

(c) Reverse-trapezoidal-fin heat sink

(3) Calculate wc,i and ww,i from the following equations.

Fig. 1. Plate-n heat sinks.

wc;i wc;i  c
Rtot Rcap Rconv ;
1
T w  T bm
; Rconv
:
Rcap
_ f
q
mc

Z
0

huif wc dy;

RH
T bm

p p  c

@Rtot
:
@p

16

11
12

_ and Tbm are the mass ow rate and the bulk mean
In Eq. (12), m
temperature of the uid, respectively. These values can be calculated by integrating the average velocity and temperature, as shown
below.

_ qn
m

@Rtot
;
@wc;i

hTif ;bm huif wc dy


:
RH f
hui wc dy
0

13

(4) Treat wc,i and p as newly guessed wc;i and p*, return to step 2.
Repeat step 24 until converged values for wc,i and p are
obtained.
(5) Calculate ww,i from the following equations.

ww;i p  wc;i :

17

Subsequently, second degree polynomial equations for the optimal


channel width wc(y) and n thickness wc(y) were calculated from
the wc,i and ww,i values using a least square tting.

D.-K. Kim et al. / International Journal of Heat and Mass Transfer 53 (2010) 59885995

5991

(a) Schematic diagram

(b) Boundary conditions

(c) Analysis domain


Fig. 2. Heat sink with variable n thickness.

In Table 2, the channel width, n thickness, n pitch, n number, n surface area, friction factor, capacitive thermal resistance,
convective thermal resistance, and total thermal resistance for
the optimized variable-thickness-n heat sink are presented and
compared with those of the optimized uniform-thickness-n heat
sink under the same constraints. As indicated in Table 2, the friction factor is lower for the variable-thickness-n heat sink. As a

result, a smaller n pitch and a larger number of ns are allowed


for the variable-thickness-n heat sink without compromising
the capacitive thermal resistance. The larger number of ns lowers
the convective resistance, as the convective resistance generally
decreases as the n surface area, which is proportional to the number of ns, increases. Consequently, the variable-thickness-n heat
sink can have lower convective resistance than the uniform-thick-

5992

D.-K. Kim et al. / International Journal of Heat and Mass Transfer 53 (2010) 59885995

(a) Thermal resistances for rectangular and reverse-trapezoidal heat sink

(b) Thermal resistances for heat sink with variable fin thickness
Fig. 3. Thermal resistances for plate-n heat sinks. (L = W = 1 cm, lf = 0.000855 kg/m s, cf = 4179 J/kg K, qf = 997 kg/m3, p = 1 mm, ks = 148 W/m K, kf = 0.613 W/m K,
Ppump = 2.56 W).

Table 1
Geometrical description of typical variable-thickness-n heat sinks.
Length (L)  width (W)

Fin pitch (mm)

Case A

1 cm  1 cm

0.1

Schematic of heat sinks (not to scale)

0.16(y/H)2  0.42(y/H) + 0.57

Porosity

Case B

1 cm  1 cm

0.1

0.16(y/H)2 + 0.42(y/H) + 0.43

Case C

1 cm  1 cm

0.1

0.21(y/H)2  0.59(y/H) + 0.55

Case D

1 cm  1 cm

0.1

0.21(y/H)2 + 0.59(y/H) + 0.45

5993

D.-K. Kim et al. / International Journal of Heat and Mass Transfer 53 (2010) 59885995

Fig. 4. Nusselt numbers for rectangular and triangular channels.

Table 2
Comparison of the optimized heat sinks with variable-thickness-ns and uniform-thickness-ns.
Variable-thickness-ns
Present model
Constraints
Length (L)  width (W)
Height (H)
Pumping power
Solid
Fluid
Results
Fin pitch (mm)
Fin number
Channel width (mm)
Fin thickness (mm)
Surface area (m2)
f Re
Rcap(C/W)
Rconv(C/W)
Rtot(C/W)
Schematic of channel (not to scale)

Uniform-thickness-ns
2D numerical results

Present model

2D numerical results

0.427
93
0.128
0.299
0.300
23.8
0.0034
0.0098
0.0132

0.0034
0.0098
0.0132

4 cm  4 cm
0.4 cm
2.56W
Aluminum
Water
0.276
144
0.0552((y/H)2  3(y/H) + 2.75)
0.0552((y/H)2 + 3(y/H) + 2.25)
0.501
16.4
0.0040
0.0042
0.0065
0.0066
0.0105
0.0108

ness-n heat sink while their capacitive resistances are similar. Finally, the total resistance of the optimized variable-thickness-n
heat sink is lower than that of the optimized uniform-thicknessn heat sink. In the case presented in Table 2, the total thermal
resistance shows a decrease of about 15% after allowing the n
thickness to vary in the direction normal to the uid ow.
The optimized geometries and thermal resistances of the variable-thickness-n and uniform-thickness-n heat sinks for various
pumping powers and various lengths are listed in Table 3. Through

a comparison of the thermal resistances of the optimized variablethickness-n and uniform-thickness-n heat sinks, a contour map
was drawn, as shown in Fig 5. Fig. 5 depicts the ratio of the thermal
resistances of the variable-thickness-n and uniform-thickness-n
heat sinks (Ropt,var/Ropt,uni) for water-cooled systems. In Fig. 5, in the
region where the ratio is less than 1, the optimized variable-thickness-n heat sink performs better than the optimized uniformthickness-n heat sink. Additionally, the opposite is true when
the ratio is greater than 1. Therefore, the contour map indicates

5994

D.-K. Kim et al. / International Journal of Heat and Mass Transfer 53 (2010) 59885995

Table 3
Comparison of geometries and thermal resistances for optimized heat sinks.
Dimensionless
pumping power
Ppump

l3 q2 D1
h;fr

1011

1011

1012

1012

1013

1013

Dimensionless
length

Type of ns

Dimensionless
channel width

Dimensionless
n thickness

wc
Dh;fr

ww
Dh;fr

Uniform thickness

0.0024

0.0057

0.289

Variable thickness

0.0011((y/H)2  2.8(y/H) + 2.62)

0.0011((y/H)2 + 2.8(y/H) + 2.14)

0.239

Uniform thickness

0.0089

0.0057

0.105

Variable thickness

0.0101((y/H)2  2.43(y/H) + 9.43)

0.0101((y/H)2 + 2.43(y/H) + 4.86)

0.103

Uniform thickness

0.0016

0.0056

0.222

Variable thickness

0.0012((y/H)2  2.41(y/H) + 1.56)

0.0012((y/H)2 + 2.41(y/H) + 1.38)

0.153

Uniform thickness

0.0059

0.0056

0.063

Variable thickness

0.0120((y/H)2  2.55(y/H) + 5.55)

0.0120((y/H)2 + 2.55(y/H) + 3.55)

0.061

Uniform thickness

0.001

0.0057

0.182

Variable thickness

0.0014((y/H)2  1.84(y/H) + 0.86)

0.0014((y/H)2 + 1.84(y/H) + 0.86)

0.099

Uniform thickness

0.0040

0.0055

0.041

Variable thickness

0.0013((y/H)2  2.63(y/H) + 3.56)

0.0013((y/H)2 + 2.63(y/H) + 2.69)

0.037

L
Dh;fr

10

10

10

Thermal
resistance (C/W)
R

Schematic of
channel
(not to scale)

(Dh,fr = 0.01 m, lf = 0.000855 kg/m s, cf = 4179 J/kg K, qf = 997 kg/m3, ks = 148 W/m K, kf = 0.613 W/m K).

that optimized variable-thickness-n heat sinks have lower thermal resistances than optimized uniform-thickness-n heat sinks
in practical situations. The thermal resistance is reduced by as
much as nearly 15% by employing variable-thickness-ns. The difference between the thermal resistances increases as the length

decreases and as the pumping power increases because the convective thermal resistance becomes dominant over the capacitive
thermal resistance as either the pumping power increases or the
length decrease. Moreover the variable-thickness-n heat sinks
can reduce the convective thermal resistance effectively without

D.-K. Kim et al. / International Journal of Heat and Mass Transfer 53 (2010) 59885995

5995

Fig. 5. Contour plots of Ropt,var/Ropt,uni (ks/kf = 2.41  103, Pr = 0.707, afr = H/W, Dh,fr = (2HW/(H + W))).

compromising the capacitive thermal resistance, as mentioned


above. The ratio of the thermal resistances decreases as the aspect
ratio of the frontal area increases as well.
4. Conclusion
In the present paper, thermal optimization of a plate-n heat
sink was conducted with the n thickness varying in the direction
normal to the uid ow. A model based on the volume averaging
theory (VAT) was used for this optimization. The thermal resistance was reduced by as much as 15% compared to uniform-thickness-n heat sinks by employing variable-thickness-ns in the
case of water-cooled heat sink. The amount of the reduction increases as either the pumping power increases or the length of
the heat sink decreases. Due to its high thermal performance, the
variable-thickness-n heat sink is expected to be suitable for as a
next generation cooling solution.

References
[1] A. Bar-Cohen, Thermal management of electric components with dielectric
liquids, in: J.R. Lloyd, Y. Kurosaki (Eds.), Proceedings of ASME/JSME Thermal
Engineering Joint Conference, vol. 2, 1996, pp. 1539.
[2] W. Nakayama, Thermal management of electronic equipment: a review of
technology and research topics, Appl. Mech. Rev. 39 (12) (1986) 18471868.
[3] R.W. Knight, D.J. Hall, J.S. Goodling, R.C. Jaeger, Heat sink optimization with
application to microchannels, IEEE Trans. Compon. Hybrids Manuf. Technol. 15
(1992) 832842.
[4] J.H. Ryu, S.J. Kim, D.H. Choi, Numerical optimization of the thermal
performance of a microchannel heat sink, Int. J. Heat Mass Transfer 45 (2002)
28232827.
[5] S.J. Kim, D. Kim, Closed-form correlations for thermal optimization of
microchannels, Int. J. Heat Mass Transfer 50 (2007) 53185322.
[6] M. Morega, A. Bejan, Plate ns with variable thickness and height for air-cooled
electronic modules, Int. J. Heat Mass Transfer 37 (1994) 433445.
[7] R.K. Shah, Laminar ow friction and forced convection heat transfer in ducts of
arbitrary geometry, Int. J. Heat Mass Transfer 18 (1975) 849862.

Anda mungkin juga menyukai