You are on page 1of 8

1

NUMERIC SOLVER FOR MULTIMODE HEAT TRANSFER


IN A HIGH TEMPERATURE HEAT SINK
Ben Jones
Purdue University
Mark Kimber
Purdue University
ABSTRACT
A finite volume based numerical scheme is developed to
estimate the temperature distribution within a high temperature
rectangular fin heat sink. A coupled conduction, convection,
and radiation problem is solved with conduction occurring
through the substrate and radiation and convection boundary
conditions on the surfaces of the fins. Those boundaries
involving both radiation and convection to the ambient are
accounted for by linearizing radiation terms and updating the
temperature-based radiosities at each iteration. The convection
coefficient and heat sink thermal conductivity are assumed to
be uniform and independent of temperature. A geometric
multi-grid method is implemented to drastically decrease
computation time. A number of test cases are analyzed for a
heat sink with overall dimensions of 50 mm x 50 mm x 5 mm
while varying the fin thickness and number of channels.
Using the performance metric of thermal resistance based on
differences between base and ambient temperatures, the
optimum design composes of 6 channels with 1 mm thick fins
corresponding to a thermal resistance of 13.6 K/W.
INTRODUCTION
Heat sinks are commonly used in many industries to
increase the rate of heat removal over which forced or free
convection can occur. They can most often be found in various
electronic devices such as microprocessors and high
performance video cards. In many cases, heat sinks are coated
with a high emissivity paint to further increase the heat removal
rate. Radiation then becomes important and determining the
temperature distribution within the heat sink is non-trivial and
must account for radiosities of all exposed surfaces and view
factors from one fin to the next. The current work provides a
numerical approach to solving this problem. For reference, the
code in its entirety is included in Appendix A and is based on a
finite volume approach. In the remainder of this paper, the
problem is first described in more detail including treatment of
all boundary conditions. Discretized equations representing the
heat transfer throughout the entire domain are presented in
forms most easily implemented into line by line tridiagonal
matrix (LBLTDMA) solvers. The exchange of radiation is then
discussed with methods for calculating view factors and
radiosities of all surfaces exposed to ambient conditions. The
overall solution procedure is given as well as a description of
the multi-grid scheme implemented to accelerate convergence.
This is followed by a code validation comparing the present
work with similar results obtain using Fluent. Finally results
from a number of test cases are given showing the ability to
optimize the geometric dimensions of the heat sink.
NOMENCLATURE
A Coefficient for q
b
A
i
Cell surface area
a Cell coefficient
B Coefficient for q
b
b Source term
E
b
Black body emissive power
F
ij
View factor from surface
I
to surface j
g Gravitational constant
H Overall height of heat sink
h Convection coefficient
k Thermal conductivity of substrate
k
f
Thermal conductivity of fluid
L
0
Overall length and width of heat sink
Q
in
Total heat input into the base of the heat sink
q Heat flow
q
0
'' Heat flux into base of heat sink
Ra
S
Raleigh number based on channel width
R
th
Overall thermal resistance of heat sink
S Channel width
T Temperature
Ax Cell width
ox Distance in x-direction
Ay Cell height
oy Distance in y-direction
Greek
o Thermal diffusivity
| Coefficient of thermal expansion
c Emissivity
v Kinematic viscosity
o Stefan-Boltzmann constant
Subscripts/Superscripts
b Boundary
conv Convection
E East neighboring cell
e East face
i,j Surface indices
N North neighboring cell
n North face
nb Near boundary
2
p Particular cell
rad Radiation
S South neighboring cell
s South face
Tb Linearized with respect to T
b
W West neighboring cell
w West face
w
1
Arc length of boundary surface
Ambient
* Value based on prevailing iterate
PROBLEM DESCRIPTION
The heat sink of interest in this paper is shown in Figure 1
where dashed lines reveal the domain considered during
analysis. This domain is shown in Figure 2 where the heat flux
at the base is specified (q
0
''). In addition, due to symmetry,
both the far left and right boundaries can be considered
insulated yielding a total of three pure Neumann boundary
conditions. The discrete equations for these cells are given in
Eq. (1) where the coefficient for any particular cell (a
P
) is
simply the sum of the coefficients for all neighboring cells to
the east (E), west (W), north (N), and south (S). For the case of
a specified non-zero heat flux (as is the case with the heat sink
base), the b term simply becomes q''Ax as shown in Eq. (1), but
in all other cases, this is zero.
Figure 1: Illustration of heat sink with overall dimensions
of L
0
x L
0
x H and heat flux entering through base.
Figure 2: Heat sink subsection used for analysis with
temperature of surroundings equal to ambient conditions.
Surfaces 1-5 experience heat transfer to ambient through
both convection (q
conv
) and radiation (q
rad
)
( ) ( ) ( ) ( )
P P nb nb
nb
E W N S
e w n s
P nb
nb
a T a T b
k y k y k x k x
a , a , a , a
x x y y
a a
q'' x for specified non-zero heat flux
b
0 otherwise
= +
A A A A
= = = =
o o o o
=
A
=

(1)
The treatment of the remaining boundaries (surfaces 1-5 of
Figure 2) however is more difficult because both convection
and radiation contribute to the heat leaving those boundaries.
The radiation in particular requires careful consideration and
due to the enclosure type problem, radiosities must be
computed to accurately describe the total heat transfer
occurring at these surfaces. Moreover, once the domain is
discretized these procedures become increasingly more
complex and are described separately in the next section.
MULTIMODE BOUNDARY CONDITIONS
We begin the analysis of boundaries where both convection and
radiation occur by observing the heat leaving a discretized
surface as shown in Figure 3 where T
P
and T
b
are the cell
centroid and boundary temperatures, respectively. The total
heat leaving (q
b
) through this surface is expressed in Eq. (2)
where h is the convection coefficient, c is the surface
emissivity, and J
b
*
is the radiosity. The notation (*) denotes a
value which is computed using prevailing values, or those from
a previous iteration. Also, it is important to note that this
expression is strictly representative of a cell with a vertical
surface exposed to ambient conditions, hence the use of Ay.
For a cell with a horizontal boundary, one needs to simply
replace this with Ax for this equation and all equations
hereafter. The detailed treatment of J
b
*
will follow in
subsequent sections and is based on view factors and
temperatures of all surfaces within the enclosure. For the time
being, we will simply accept that this quantity is calculated
from prevailing values and treat it as a known parameter. The
expression in Eq. (2) can be rearranged grouping terms
common to T
b
and those based on specified parameters as is
shown in Eq. (3). The radiation term containing T
b
4
must now
be linearized to obtain discrete equations which ultimately form
a linear set of equations. This is done by expressing the non-
linear quantity in terms of prevailing and current iteration
values of T
b
as is shown in Eq. (4) where the non-linear term
involving T
b
in Eq. (3) has been redefined as q
Tb
. It is worth
noting that once convergence is met (T
b
= T
b
*
), the expression
simplifies to the original definition of q
Tb
. Upon substitution of
Eq. (4) into Eq. (3), the overall expression for heat leaving the
boundary takes the form shown in Eq. (5), where again the
terms involving T
b
and those based on known parameters are
grouped together. The T
b
coefficient and the constant are
redefined as B
*
and A
*
, respectively and it is important note
that regardless of the initial guess or the accuracy of the current
Insulated due
to symmetry
L
4
5
q
rad
q
conv
T

Enclosure
2
T

1
3
q
0
''
q
0
''
H
L
0
L
0
3
iteration, both B
*
and A
*
remain positive quantities. This will
prove useful once the final discretized equations are presented.
( )
( )
4 *
b b
b conv rad b
T J y
q q q h y T T
1

o cA
= + = A +
c
(2)
Figure 3: Illustration for calculation of total heat leaving
discretized boundary cell
4 *
b b
b b
T y J y
q h yT h yT
1 1

| | | | o cA cA
= A + A +
| |
c c
\ . \ .
(3)
( ) ( )
( ) ( )
( )
*
4
* * b
Tb Tb Tb b b
b
4 3
* *
b b
*
b b
T y d
q q q T T
1 d(T )
T y 4 T y
T T
1 1
( o cA
= = +
(
c

o cA o cA
= +
c c
(4)
( )
( )
( )
3
*
b
b b
4
* *
b b
* *
b
4 T y
q h y T
1
3 T J y
h yT B T A
1

| |
o cA
|
= A +
|
c
|
\ .
| |
o + cA
|
A + =
|
c
|
\ .
(5)
The next step is to balance the expression for q
b
with the
conduction from the cell centroid to the boundary as is shown
in Eq. (6), which can then be solved for T
b
as shown in Eq. (7).
This result is then substituted back into Eq. (5) and used in
conjunction with the overall energy balance within the cell,
namely all the heat leaving through conduction to neighboring
cells plus the heat leaving through the boundary must equal
zero. For the particular cell configuration shown in Figure 3,
the three neighboring cells are to the west, north, and south.
The conductive fluxes are balanced as previously described and
the result is shown in Eq. (8), which represents the discrete
equation used for all boundary cells where both convection and
radiation occur. A more convenient form is presented in Eq.
(9), which can easily be adapted to cells with horizontal
surfaces exposed to ambient conditions as well. Notice that
because both B
*
and A
*
are positive quantities, a
P
is greater
than the sum the neighboring coefficients (a
W
, a
N
, and a
S
)
thereby satisfying the scarborough criterion in the inequality
enabling the use of iterative techniques for solving the system
of equations.
( )
P b * *
b b
2k T T y
q B T A
x
A
= =
A
(6)
*
b P * *
2k y A x
T T
2k y B x 2k y B x
( ( A A
= +
( (
A + A A + A

(7)
( )
( )
( )
( )
( )
( )
P W P N P S
w n s
*
* *
P * *
k y k x k x
T T T T T T
x y y
2k y A x
B T A 0
2k y B x 2k y B x
A A A
+ + +
o o o
( | | ( ( A A
+ = ( |
( (
|
A + A A + A
( \ .
(8)
( ) ( ) ( )
P P nb nb
nb
W N S
w n s
*
P nb *
nb
*
*
*
a T a T b
k y k x k x
a , a , a
x y y
2kB y
a a
2k y B x
B x
b A 1
2k y B x
= +
A A A
= = =
o o o
A
= +
A + A
| | A
=
|
A + A
\ .

(9)
RADIATION EXCHANGE
As discussed in the previous section, J
b
must be evaluated
to determine the q
rad
portion of the boundary heat flux. The
radiosity from each discretized surface can be evaluated by
solving the system of linear equations [1]:
( )
( )
N
i j
bi i
1
i i i
j 1
i ij
J J
E J
1 A
A F

=

=
c c

(10)
Rewriting Eq. (10) in a form more convenient for programming
yields:
q
conv
q
rad
Ax
Ay
T
b
T
P
4
( )
N
i i ij j i
j 1, j i
N
i i
i ij
i
j 1, j i
ij i ij
4 i i i i
i bi bi
i i
a J a J b
A
a a
1
a A F
A A
b E T
1 1
= =
= =
= +
c
= +
c
=
c c
= = o
c c

(11)
where c
i
is the emissivity, A
i
is the surface area, o is the Stefan-
Boltzmann constant, E
bi
is the black body emissive power, and
T
bi
is the surface temperature. The form of Eq. (11) is similar
to a diffusion equation. However, instead of being linked to
just the near boundary cells, J
i
is dependent on the radiosity
from all surfaces j in the domain that surface i undergoes
radiation exchange with.
In the event that c
i
= 0, the source term, b
i
, goes to 0 and a
i
evaluates to just the sum of the a
ij
coefficients. In the other
extreme, if c
i
= 1, a
i
and b
i
goes to infinity. Clearly this is not
desirable from a computational standpoint. From Eq. (10), it is
apparent that J
i
= E
bi
when c
i
= 1. By preserving the form of
Eq. (11), the coefficients can be evaluated as:
i
ij i
4
i bi
a 1
a 0 for 1
b T

= c =
`

= o

)
(12)
Eq. (11) satisfies the Scarborough Criteria unconditionally
in the equality and conditionally in the inequality as long as c
i
>
0 for at least one surface in the domain. Since c
i
> 0 for
practical engineering surfaces, the Gauss-Seidel method is
guaranteed to converge and is the solution method chosen in
the present work. A TDMA method is not recommended for
solving Eq. (11) since the radiosity of surface i is linked to all
of the visible surfaces j in the enclosure, not just the neighbors.
Eq. (11) also guarantees boundedness of the solution; J
i
is
bounded by min[J
j
for all j i, E
bi
] and max[J
j
for all j i, E
bi
].
The view factors for 2D surfaces, infinite in one extent, can
be conveniently evaluated using the cross-strings method [1].
The view factor between any two surfaces 1 and 2 (see Figure
4) is given by the expression:
( ) ( )
12
1
1
F ac bd ad bc
2w
= + + (13)
where w
1
is the arc length along surface 1. The absolute value
in Eq. (13) is to emphasize that whether ac and bd are cross-
distances or direct-distances (shown as ad and bc in Figure 4),
the result only differs in sign. By implementing the absolute
value in the program, no elaborate algorithms are required to
ensure that the cross-distances are evaluated first.
1
2
w
1
a
b
c
d
1
2
w
1
a
b
c
d
Figure 4: Cross-strings method for calculating view factors
SOLUTION PROCEDURE
The calculation of the cell values, T
P
, requires that the
boundary surface radiosities, J
b
, are known a priori. However,
the surface radiosities are dependent on the boundary surface
temperatures, T
b
, which are unknowns. Therefore, the basic
methodology behind the solution procedure is to first calculate
the J
b
based on a guess of T
b
, calculate T
P
based on the
estimated J
b
, update T
b
based on the new T
P
values, and iterate.
The detailed solution procedure is given below. Again, the
starred (*) terms indicate values based on the previous iterate.
1) Generate grid based on user inputs for the number of
cells spanning the fin width and fin height and the
number of cells spanning the overall domain height
and width.
2) Calculate view factors, F
ij
, using Eq. (13).
3) Calculate the cell coefficients a
W
, a
E
, a
N
, and a
S
using
Eq. (1).
4) Use a guess to initialize the cell centroid temperatures,
T
P
, and boundary temperatures, T
b
.
5) Solve the system of equations in Eq. (11) to find the
surface radiosities, J
b
, based on T
b
*
starting with an
initial guess for J
b
. Iterate until convergence using
Gauss-Seidel iteration.
6) Estimate the convection coefficient, h, using the
Elenbass correlation [1] based on T
b
*
. This step was
deemed necessary since the convection coefficient
changes considerably with the channel dimensions.
The Elanbass correlation is given below.
( )
3 4
* f
S *
0 S 0
k 1 S 35
h Ra 1 exp
S 24 L Ra S L
( | |

=
( ` |
( \ . )
(14)
where k
f
is the thermal conductivity of the fluid, S is
the channel width, and Ra
S
*
is the Raleigh number
based on the channel width given by:
( )
* 3
b
*
S
g T T S
Ra

|
=
vo
(15)
where g is the gravitational constant, | is the
coefficient of thermal expansion with | 1/T

for air,
v is the kinematic viscosity and o is the thermal
diffusivity.
7) Calculate the remaining cell coefficients, a
P
and b,
from Eq. (9).
5
8) Solve for the cell centroid temperatures, T
p
, using T
b
*
and J
b
*
(see Eq. (9)) with a Line-by-Line TDMA
routine. The Line-by-Line TDMA routine performs a
user specified number of sweeps through the domain
so that a reasonably accurate temperature field is
obtained based on the current values of T
b
*
and J
b
*
.
This was found to improve computation efficiency
compared to evaluating J
b
* and T
b
* after every sweep.
9) Calculate the boundary temperatures, T
b
, based on the
newly calculated values of T
P
using Eq. (7).
10) Check for converge. If the convergence criterion is
met, the solution is complete. Otherwise, go back to
step 5) and iterate.
It is worthwhile discussing the data structures used in the
program; most notably, how to store the view factors,
radiosities, and boundary temperatures and link them to their
respective boundary cells. Figure 5 shows the labeling scheme
used in the program. Each of the labeled surfaces 1 through 6
is divided into discrete surfaces based on the grid used in the
domain. Surface 6 is a virtual surface representing the
enclosure for surfaces 3 through 5. The node points of surface
6 are determined by projected the nodes from surface 4 onto 6.
1 2
3
4
5
6
T

1 2
3
4
5
6
T

1 2
3
4
5
6
T

(a)
1 2 3 4 5 6
b
T 1 P 1 P 1 P 1 P 1 P 1 P =
(b)
Figure 5: (a) Labeling of boundary surfaces and (b) data
structure of array
The boundary temperatures, T
b
, and radiosities, J
b
, are
stored into 1xP arrays with P = EP
i
representing the total
number of boundary surfaces and P
i
represents the number of
boundary surfaces on the labeled surface i shown in Figure 5a.
The structure of the array is shown in Figure 5b. The direction
of the arrows on Figure 5a indicates the direction of increasing
index for the array along each labeled surface. The view
factors, F
ij
, are stored in a PxP matrix with each index
indicating the i
th
and j
th
surface. Since the number of cells
spanning the fins and overall domain is known, the cell to
which each surface belongs can be readily determined.
GEOMETRIC MULTIGRID
At the boundary cells, T
p
is linked to T

through the a
P
and
b coefficients of Eq. (9). If these coefficients are small
compared to the a
nb
coefficients, T
p
will be more dependent on
the current values of the near boundary temperatures than T

.
For the heat sink properties and conditions used in the present
study, the convergence rate was quite slow. Therefore, it was
decided to implement a rudimentary multi-grid method to
accelerate convergence.
(a) (b)
(c)
Figure 6: Meshing for heat sink: (a) 2 x 3, (b) 10 x 10, (c)
40 x 40.
An example of the meshing using in the geometric multi-
grid method is shown in Figure 6. The program first iterates on
a coarse mesh such as the one shown in Figure 6a using a user
supplied initial guess. Once the solution has converged on the
coarse mesh, the cell values are interpolated onto a finer mesh.
The process repeats until the finest mesh level is achieved.
Five grid levels were implemented in the present work: 2x3,
5x5, 10x10, 20x20, and 40x40 meshes.
There are some limitations to the multi-grid method as
implemented. For instance, if the program stalls on a fine
mesh, it will not go back to a coarser mesh. The program
simply proceeds from the coarsest to finest mesh. In some
instances, it was found that if the convergence criterion was set
too small, the program would stall on the finer meshes. For the
present study, a relative error between iterates of 10
-6
was
found to allow fast convergence and accurate solutions. The
computational acceleration due to the multi-grid method is
astonishing. Prior to the multigrid, one case took 2 hrs, 45 min,
and 58.3 sec to converge using a 40x40 mesh. The same case
can now be completed in 2.02 sec, almost a 5000x decrease in
computation time.
MODEL VERIFICATION
To ensure that our results are meaningful, our numerical
code must be validated. This was achieved by benchmarking
our code against a commercial CFD package, Fluent. Since the
6
boundary conditions of the current problem (convection and
radiation) are difficult to implement directly in Fluent, it was
decided to benchmark the radiation and convection parts
separately.
Table 1: Properties and dimensions of heat sink used in
model verification
Overall Width 6.25 mm (40 cells)
Overall Height 5 mm (40 cells)
Fin Width 0.6 mm (8 cells)
Channel Height 1.3 mm (16 cells)
Thermal Conductivity 180 W/m-K
Heat Flux 10,000 W/m
2
The dimensions and properties of the heat sink used in the
model validation are given in Table 1. To benchmark the
convection part of the code, the radiation was turned off. The
convection coefficient was set to a constant of 15 W/m
2
-K with
an ambient temperature of 300 K. The results from the Fluent
simulations and our code is shown in Table 2.
Table 2: Comparison of results from numerical code and
Fluent
Convection Only, h = 15 W/m
2
-K, T

= 300 K
T
p
, min (K) T
p
, min (K)
Code 770.36 770.71
Fluent 770.70 771.06
Radiation Only, c = 0.9, T

= 300 K
T
p
, min (K) T
p
, min (K)
Code 661.32 661.66
Fluent 661.51 661.85
The radiation part of the code was verified by turning
convection off and setting the emissivity to 0.9. The
computational domain and boundary conditions used in the
Fluent model is shown in Figure 7. A dummy fluid was
specified to fill the channel region, which allows the surface-to-
surface (s2s) radiation model to be implemented. The fluid
motion was specified to be stagnant and the thermal
conductivity was set to a negligible value to eliminate the
effects of the fluid on the heat transfer. Surfaces 3-6 used the
s2s model. Surfaces 1 and 2 were set to radiation boundary
conditions and surface 6 was set to a constant temperature. The
results of the Fluent simulations and computational model are
shown in table 2. It is apparent that our code is performing as
expected.
1 2
3
4
5
6
T

q"
Dummy
Fluid
1 2
3
4
5
6
T

q"
1 2
3
4
5
6
T

1 2
3
4
5
6
T

q"
Dummy
Fluid
Figure 7: Boundary conditions used in Fluent simulation
The mesh independence of the solution was also explored
using the mesh sizes as stated before: 2x3, 5x5, 10x10, 20x20,
and 40x40. Due to the high thermal conductivity of the heat
sink, the temperature in the heat sink is close to uniform so the
results are relatively insensitive to mesh size. In fact, even the
2x3 mesh yielded temperatures very close to the 40x40 mesh
(within 0.05 K). However it was decided to use the 40x40
mesh both due to the computational efficiency of the multi-grid
algorithm and because the temperature contour plots are more
aesthetically pleasing on the 40x40 mesh than the coarser
meshes.
RESULTS FROM OPTIMIZATION
To further explore the performance of an entire family of
heat sinks, the thermal resistance was analyzed varying the fin
thickness and number of channels keeping all other parameters
fixed at specified values. Figure 8 illustrates the heat sink
orientation considered and the various parameters which are
kept fixed during the optimization routines. The values used
for these parameters are listed in Table 3 along with the
specified channel height, which also remains fixed. The
thermal and radiative properties of the heat sink are based on
aluminum painted black. The range considered for fin
thickness and number of channels can also be found in Table 3
and analysis was performed at 21 and 19 discrete levels over
each respective range. The channel width was constrained to
be greater than 1 mm to prevent unrealistically small or
negative values. For the case of 1 mm thick fins and 10
channels, the converged temperature contours are shown in
Figure 9. For this case and all cases considered during
optimization, the result is very near isothermal (a temperature
difference of 0.24 K for the case shown in Figure 9). This is
due to the geometric sizes considered as well as the high
thermal conductivity of the heat sink.
7
Figure 8: Geometric orientation of heat sink considered
during optimization
Table 3: Fixed and variable parameters used during
optimization routines
Fixed Parameters
L
0
(mm) H (mm) q
0
'' (W/m
2
) k (W/m-K)
50 5.0 5000 180
T

(K) emis. - c Chan.Height (mm)


300 0.95 3.5
Variables for Optimization
Fin Thickness (mm) Number of Channels
[1 5] [2 20]
Figure 9: Temperature distribution for fin thickness of 1
mm and 10 channels
Thermal resistance (R
th
) defined by Eq. (16) is used as the
performance metric, where
bs
T is the average base temperature
and Q
in
is the overall heat input to the heat sink. The thermal
resistance was computed for each unique case and a surface
plot of the results is shown in Figure 10 where a minimum
occurs with a fin thickness of 1 mm and 6 channels yielding an
overall resistance of 13.6 K/W. For heat sinks with 9 channels
or greater, the thermal resistance could not be computed up to
fin thicknesses of 5 mm because the channel width was below
the constraint of 1 mm.
bs
th
in
T T
R
Q

= (16)
Figure 10: Optimization results varying fin thickness and
number of channels
As the number of channels increases for any given fin
thickness, a drop in thermal resistance is observed for a small
number of channels. A minimum thermal resistance is reached
before the resistance increases with increasing number of
channels. The thermal resistance of the heat sink is a balance
between the convective and radiative fluxes and the overall
surface area. The natural convection coefficient (given by Eq.
(14)) typically decreases with increasing number of channels
due to the decreased fluid motion caused by the constrained
channel dimensions. The radiative heat flux also diminishes
due to the narrowing channel widths. However, the surface
area of the heat sink increases with increasing number of
channels. For a small number of channels, the increase in
surface area more than compensates for the decreasing
convective and radiative fluxes. However, after an optimum is
reached, the increase in surface area is no longer able to
compensate for the decreasing boundary fluxes resulting in an
increase in thermal resistance.
The optimum number of channels is also dependent on the
fin thickness. However, the high thermal conductivity of the
heat sink material results in high fin efficiency for the fin
thicknesses under consideration. Therefore, the thermal
resistance of the heat sink increases for increasing fin thickness
(i.e. decreasing channel width) due to the reduced boundary
fluxes.
CONCLUSION
A numerical approach to multimode heat transfer in a high
temperature heat sink has been developed and validated against
a commercial CFD package. This code developed uses
linearization techniques to treat boundaries that experience heat
X (mm)
Y
(
m
m
)
0 1 2 3 4 5
0
1
2
3
4
5
T (K)
498.68
498.64
498.6
498.56
498.52
498.48
498.44
q
0
''
H
g
L
0
L
0
T

c, h
8
transfer through both convection and radiation. Radiosities are
computed using a surface to surface radiation exchange model.
A geometric multi-grid method was found to significantly
improve computational performance (a 5000x speedup for the
cases explored). The model was used to explore the optimal
dimensions of a high temperature heat sink that undergoes
radiation and natural convection. The optimal design of a 50
mm x 50 mm x 5 mm aluminum heat sink with a heat input of
12.5 W consists of 6 channels with 1 mm fin widths. The code
could be easily implemented to study the heat transfer in other
geometries where convection and radiation are important.
However, the numerical model is currently formulated for
rectangular, structured meshes and the formulation may require
modification for other meshing geometries.
REFERENCES
[1] Incropera, F.P., DeWitt, D.P., Fundamentals of Heat
and Mass Transfer, 5
th
ed., John Wiley & Sons, NY.

You might also like