You are on page 1of 9

Proceedings of the ASME Summer Heat Transfer Conference

HT2012
July 8-12, 2012 Puerto Rico, USA

HT2012-58443
APPLICATION OF KOVALEV TYPE MODELING TO EVAPORATION IN BIPOROUS
MEDIA
Sean W. Reilly
University of California, Los Angeles Mechanical
and Aerospace Engineering Department
Los Angeles, CA, U.S.A.

ABSTRACT
Sintered copper porous media has found many uses in
the electronics cooling industry as it effectively transfers energy
while maintaining low heater side temperatures. Evaporator
wicks of this type transfer heat through sensible and latent heat
as the liquid evaporates. The evaporation process is a function
of the input heat flux and the capillary pressure available to
pump condensed liquid to the surface of the heater from the
condenser. As vapor is produced, it must be shed from the
wick. A biporous wick is particularly effective for this
application as there are two distinct size distributions of pores;
small pores to provide ample capillary pressure in order to
drive flow through the wick and large pores to provide high
permeability for escaping vapor. Modeling of this process
through the complex geometry of a biporous wick has been
attempted many different ways for boiling in the wick;
however, this work will focus strictly on evaporation. The
modeling proposed in this work is based on the work by
Kovalev (1987), which used a pore size distribution in order to
determine the most probably pore size at a given position. The
model distinguishes phases by choosing a cutoff pore size,
above which all pores were assumed to be filled with vapor and
below which filled with liquid. For a given thickness and
thermophysical properties of the liquid, this 1-D model predicts
a temperature difference across the wick for a given input heat
flux. This model is especially appropriate for Biporous wicks
where the size distribution of pores is fairly well known and the
large and small pores differ enough in size for the assumption
of fluid phases to be reasonably acceptable. The modeling
proposed in this work was compared to experimental data
collected on biporous evaporators at UCLA for validation. It is
hoped that this modeling is the starting point for more extensive
modeling and optimization of biporous evaporators for phase
change heat transfer devices.

Ivan Catton
University of California, Los Angeles Mechanical
and Aerospace Engineering Department
Los Angeles, CA, U.S.A.

INTRODUCTION
Heat pipes are incredibly useful phase change heat transfer
devices for use in many applications, such as electronics
cooling and space based devices. A heat pipe functions by
adding heat at one end and removing it from the other. Where
the heat is added, liquid is evaporated, raising the pressure
inducing flow of vapor to the condenser. Once the liquid has
condensed it travels down the walls of the tube back to the
evaporator. The simplest version of a heat pipe, a thermosiphon or Perkins tube, the walls of the tube are smooth and the
device is oriented perpendicular to gravity with heat added at
the bottom. The fluid is inserted and is put under a vacuum of a
given strength in order to manipulate the saturation
temperature. The saturation temperature is generally chosen to
be application specific, so that operation occurs in a prescribed
range.
In order to expand the performance envelope of these type
of devices, the surface morphology of the three main regions of
the heat pipe (evaporator, condenser, and the adiabatic region
separating the latter and former) can be changed to enhance
performance. A common modification is to add structures to
the interior of the device in order to provide more surface area
for evaporation and provide capillary pressure to aid in liquid
supply to the evaporator. There are many examples of these
structures; grooves, screens, and sintered copper to name a few.
The added capillary pressure for liquid supply allows the device
to be operated in orientations opposing the force of gravity as
well as altered shapes.
This work will be limited to the investigation of sintered
copper wicks as liquid supply and evaporator. The particular
wick used for this investigation was a biporous wick, in
which there are two distinct pore size values within the wick.
The logic being that the small pores provide capillary pressure
and the larger pores provide the vapor permeability required to
effectively transfer heat to the condensation region.

Copyright 20xx by ASME

In this work, the authors are taking the first step in a


comprehensive effort to develop a numerical model to optimize
the performance of biporous wicks in different devices. The
authors used extensive experimental work conducted by
Seminic (2007) and Reilly (2009) as the starting place for
investigation of the various thermophysical properties of a
biporous wick. It is desired to develop a way to predict the
most efficient geometry for a wick without testing, in order to
streamline the process of optimization. Also, it is believed that
a simple model of performance will aid in developing the
correlations needed for closure in more complex investigations
of biporous wicks, needed for higher accuracy.

A schematic of a boiling curve for biporous media can be


seen below in Figure 1. This figure shows how the wall heat
flux is a function of the wall superheat temperature. Note that,
in the schematics below, no monoporous layer of particles is
shown.

NOMENCLATURE
List of symbols
keff effective thermal conductivity of a biporous
Figure 1: Boiling regime map for Biporous Media

medium (W/m/oK)
qchf critical wall heat flux at which dry out of larger
2

pores in the biporous medium occurs (W/cm )


Tw- Temperature at wick/sample interface (oK)

Figure 2 is the schematic showing how vapor and liquid


interact in the various layers of the bi-layer (or double layer)
wick.

Tv Vapor temperature in test chamber (oK)


T Tw - Tv (oK)

Biporous Layer

qw wall heat flux (W/cm2)


Thickness of a particular medium layer (m)
D Cluster size dimension (m)

Monolayer

d Particle size dimension (m)


Tc Temperature measurement around sample neck

Figure 2: Schematic of double layer wick

(oK)
Ts- Temperature measurement around sample collar
perimeter (oK)

WICK CHARACTERISTICS
The wick chosen for use in this investigation was selected
because of the large amounts of experimental data collected at
UCLA that could be used to compare against. This wick had a
particle size of 60 m, an 800 m thick biporous layer and 300
m thick clusters. The overall thickness of the wick was 800
m. The nomenclature I use to describe these parameters is
Biporous Layer dimensions. The wick mentioned above will
appear as 60_300_800.
From previous work by Seminic [3] we know that the
average porosities for the monoporous layer is approximately
0.277 and for the Biporous layer is 0.642. The pore size
distributions range from 3-9 m for the monoporous layer and
30-100 m for the biporous layer.

In general, within the wick, the small pores in the clusters


provide the necessary capillary pressure needed to pump fluid
through the medium. The larger pores provide nucleation sites
within the wick.
MODELING BACKGROUND
There are many modeling techniques for use in porous
media. As phase change in porous media is an extremely
complex process, direct modeling of all the solid/liquid/fluid
interactions is not generally feasible. This is because it is
exceedingly difficult to know the exact structure of the wick,
which makes dealing with the morphology difficult on its own.
Furthermore, accounting for the disjoining pressures between
the three phases as well as the temperature gradients present in
each phase, makes it very difficult to ascertain the location of
phase boundaries and an accurate temperature field of the
region.
Some simple models of porous evaporation emphasize
the relatively high heat transfer rates provided by thin film
evaporation, such as by Hanlon and Ma (2003). In that

Copyright 20xx by ASME

particular work, a monoporous wick is saturated with liquid and


evaporation only takes place at the upper surface. The porous
media drastically increases the surface available to this method
of heat transfer. The evaporation from these thin films is
averaged spatially over the upper surface of the wick and used
as the boundary condition at the upper surface of the wick.
However, in this model, failure is indicated when nucleation
begins to take place in the wick. This is explained to be
because once nucleation begins, a vapor layer starts to form
which eventually prevents liquid from reaching the heater
surface and without it, a dramatic rise in temperature. This
modeling technique is appropriate for a saturated wick, but in
the authors case, failure is not defined by nucleation. Rather,
performance is believed to increase as a result of vapor
formation at the base of the wick.
An example of an extension of this type of modeling
was done by Kaya (2006). In that particular work, as
nucleation was intiated, a vapor region was assumed to form,
similar to what was mentioned previously. However, this was
not indicated as a failure mode, instead, it was assumed that this
was simply an approximation of the physical case where the
vapor region was actually part of a larger two phase region.
Again, for the authors case, it was desired to be able to account
for both liquid and vapor occupying the same region in varying
fractions. A more detailed approach was required.
In order to account for a combined liquid and vapor
flow through the wick with evaporation, there must be some
way of describing the location of liquid and vapor within the
wick based on the temperature and flow fields.
In a Kovalev model, certain assumptions are used in
order to circumnavigate the more complex aspects of flow and
evaporation in porous media. The central assumption made by
Kovalev is that within the wick, there is a pore size R* above
which the pores are vapor filled and below which, are liquid
filled. R* can vary in any direction through the wick, as there is
a pressure and temperature gradient within the wick. In
general, it is assumed that there is thermal equilibrium between
all phases, to make the calculations reasonable. It is also
assumed generally that there is a pressure gradient seen by the
liquid which causes liquid to flow into the wick and a separate
pressure gradient which allows vapor to be drawn out of the
wick, in the opposite direction to liquid.
This particular type of modeling was used extensively
by Uhle 1991, for her dissertation thesis. In order to derive the
equations necessary for solution, Uhle used the equations
originally presented by Kovalev, which is a hybrid of NavierStokes and Darcy flow.
In this equation, an inertial term is present along with
a formulation of the permeability of the medium. The authors
believe that this term is redundant, since inertial effects are
included in the determination of the permeability. Therefore,
the authors decided that it may be possible to simplify the
analysis to use only Darcy flow to represent the pressure
distribution in the numerical field.
This simplification is used in Kovalev based models
such as those by Chernysheva (2009). In this case, a model is

presented where a pore size distribution and R* are used to


interpret the location of liquid and vapor in the wick. However,
different from previous models, the model uses the size and
orientation of a supposed two-phase region in order to describe
phase change. When nucleation begins, a computational two
phase region is formed, where inside this region; a heat sink
term appears in the energy equation in order to account for
evaporation. The determination of R* was made by comparing
the minimum pore radius required for continued nucleation and
maintenance of the capillary pressure necessary to pump liquid
to the heater interface.
In this work, the authors used a modified version of
the modeling technique used by Vadnjal. It was reduced to a 1D model and the boundary conditions would need to be
modified to accommodate the fact that vapor would have to
travel through a liquid saturated region to escape, a requirement
not present in the original model.
DERIVATION AND SOLUTION TECHNIQUE
In order to determine wick behavior, the performance
of the wick is divided into three regions; saturated liquid, two
phase, and dryout. In the saturated liquid phase, liquid is
evaporated from the upper surface of the wick only. Liquid
flows vertically to the upper interface, but the Reynolds
numbers associated with this flow is considered too low to take
into account convective heat transfer.
Once the temperature difference reaches a certain
point over the nucleation threshold, vapor bubbles will start to
be produced at the interface between the substrate and the wick.
Once nucleation begins, energy is transferred through
evaporation in parallel with conduction through the wick. As
more heat flux is added, more evaporation takes place,
expanding the evaporation area. As more vapor is produced,
the pressure drop from the evaporation region to the ambient
becomes larger.
At a certain point, the pressure drop will exceed the
maximum available capillary pressure, and liquid will no longer
be able to reach the heater surface. This is considered to be the
critical heat flux after which, performance will decline due to
liquid receding into the clusters. When the liquid is fully
depleted from the clusters, the biporous region is considered to
only have vapor/solid interfaces, and yields a failure mode.
This is because once there is no more liquid, the vapor will
become superheated and the heater temperature will increase
rapidly as a function of heat flux and will not decline until the
wick is re-wetted. A schematic of these regimes can be shown
schematically in Figure 1.
As stated previously, the equation of motion used in
this work will be that of Darcy flow. Flow was assumed to be
at steady state and incompressible.

P =

(1)

r
v = 0

(2)

Copyright 20xx by ASME

The above equations were combined into the


continuity equation to obtain Equation (3)), yielding the
Laplacian of Pressure.

2 P = 0

(3)

The pressure distribution was solved along the


domain, subject to the boundary conditions, in order to
determine the pressure distribution.
For the energy equation, convection was ignored as
the flow velocities led to convective heat transfer coefficients
that were negligible compared to the evaporation heat transfer
in the flow regimes this work was interested in.. The Reynolds
numbers for the majority of the heat fluxes investigated were
less than 1 and so it was assumed that convection was not a
significant factor.
2

k wick T = qs

where the meniscus between the liquid and vapor is contained,


the authors replaced the r in the Young Laplace equation
(Equation (9))) with R*. Further, in order to combine the
Young-Laplace
Laplace and Darcy equations, Equation (9) was
differentiated by z so that the pressure gradients could be
converted to terms that are in terms of the vapor and liquid
pressure.

Pc Pv Pl
22** *cos
* cos R*
=

=
x
x x
R*2 ( x)
x

The expressions used for the pressure gradients can be


seen below.

(4)

These equations were subject to boundary conditions


at the base of the wick (y = 0) and the top of the wick (y = ).
At y = 0

P
=0
y
k wall

T
= qin
y

(6)

P = Psat

(7)

T = Tsat ( Psat )

(8)

Dryout is defined as being when the pressure drop


between the top of the wick and the bottom exceeds the
capillary pressure limit provided by the wick.

2*
rmin

Pl r l r
= l g
Jl
x
Kl

(11)

Pl r l r
= l g
Jl
x
Kl

(12)

(5)
The above three equations are combined and
rearranged to form a single equation that relates the liquid and
vapor momentum to the cutoff radius R*. This equation is seen
below.

r l r
dR *
R *2 r v r
=

g
+ J l (13)

v
l
dz
cos( ) v
kv
kl

At y =

P =

(10)

In order to develop the form of the energy equations in


the model, a 1-D
D control volume is taken from the wick that is
considered to have rotational symmetry about its center. In the
figure below, it can be seen how the energy equation is
formulated.

(9)

In Equation (9), rmin is the prescribed minimum pore


size in the wick. The fluid is assumed to be perfectly wetting,
so no contact angle is accounted for.
INTEGRATION OF KOVALEV EQUATIONS
Kovalev methodology works on the principle that an
imaginary cutoff radius exists, R*, where pore sizes larger tha
than
R* are filled with vapor and pore sizes smaller than R* are
filled with liquid. Since this imaginary radius represents a pore

Figure 3:: Diagram of control volume

Energy is removed away from the sides and conducted


up through the center. Further, a source term is added to
represent the evaporation taking place within the wick. This

Copyright 20xx by ASME

type of analysis was originally used by Vadnjal (2007) to look


at individual stacks of cluster, but is being extended to looking
at whole sections of wick in this case. Rearranging the
equations in the control volume in order to define an energy
balance yields the equation seen below.

Q
z + Qc = 0
z

(14)


2 ( z )
R * (z)
+
z
z
2hc R *( z ) = 0

(15)

keff

in the above equation is the temperature of the wick


subtracted from the saturation temperature at the given
pressure. In the above equations, R* is used as the area over
which the energy is transferred since it is assumed that energy
is conducted around the periphery of the cutoff pore and is
removed from the surface and boundaries of the meniscus.
Further the equation for Jv can be related to the
temperature by way of the enthalpy of vaporization. The
equations leading to this substitution can be seen below.

hc
Jv
=
dz h fg v z

Equation (18) was reduced to two first order differential


equations, seen below.

G=

G 2 ( z )
2 R * ( z ) ( z )
=
=
2
z
z
R * ( z ) z
z
2hc
+
keff R *( z )

v J v

(16)

(17)

The above equations can be combined with the


original energy and momentum equations to yield Equation
(18) and (19).

R *

Y =
G

Jv

(22)

R *


Y =
z G

Jv

(23)

dR *
R *2
=
*
dz
cos( )

r
r v r
r l v J v
J v l g +
v g

kv
kl l

(18)

=G
z

dR *
R *2
=
*
dz
cos( )

r
r v r
r l v J v
J v l g +
v g

kv
kl l

(21)

The equations used for the Euler solution are listed in order
below.

( z)
2 R * ( z ) ( z )
=
2
z
R *( z ) z
z
2hc
+
keff R *( z )

(20)

SOLUTION METHOD
A modified Euler method was used in order to solve
the above equations. The variables to be solved for in this
model are R*, Jv, G, and .

Also, by conservation of mass, the velocity of the


vapor and liquid can be related.

Jl =

(24)

(25)

(19)

Copyright 20xx by ASME

G 2 ( z )
2
R *( z ) ( z )
=
=
+
2
z
z
keff R * ( z ) z
z
2hc
keff R *( z )

(26)

model to validate the model. The first wick shown is for a wick
that is 69_462_800, examined experimentally by Seminic. The
results of this comparison are seen below in Figure 4.
dT vs q
90
80
70

(27)

60
dT (T-Te)

hc
Jv
=
dz h fg v z
The boundary conditions used are seen below.

75e 6(m,@ z = 0)

0@ z =

Y (0) = q

( keff ,@ z = 0)

0(@ z = 0)

50
40
30
20

(28)

The value used for R* @ z = 0 is meant to represent a


completely liquid saturated boundary, as this is taken to be the
maximum pore size, meaning all pores are filled with liquid.
This also means that vapor velocity at the heater interface (z =
0) is zero. The third boundary condition, for G is representing
the input heat flux. is assumed to be 0 at the uppermost
boundary as all energy is assumed to have been used up in
evaporation by the time it reaches the upper surface.
The value for q was varied from zero up to 600
(W/cm^2) in order to develop the performance curves.

10
0

100

200

300

400

500

600

700

q (W/cm2)

Figure 4: Comparison of model with experimental data for


69_462_800 wick

The above figure shows that the model predicts fairly well
the performance of the wick, at least prior to dryout. However,
this is mainly because the model is not equipped to handle
multiple boiling regions of performance. Once the wick begins
to dryout and the liquid recedes into the pores of the clusters,
the proper cutoff radius cannot be properly estimated by the
model with constant thermophysical properties.
dT vs q
60

EXPERIMENTAL RESULTS
Comparisons of model results and experimental data will
presented in this section. This work compares two wicks with

50

40
dT (T-Te)

PHYSICAL PARAMETERS
This model makes use of approximations for the
thermophysical properties of the wick. A value of 18 W/(m*K)
was used to describe the dry thermal conductivity of the wick
as this was the experimental value collected by Seminic (2007).
Furthermore, values used for the vapor and liquid permeability
were relations pulled from work by Seminic. A constant
evaporation heat transfer coefficient of (he = 2e4 W/(m^2*K)),
which is taken from a similar work by Vadnjal.
The values for density and viscosity were taken for
liquid water at an assumed temperature of 40 degrees Celsius as
this is a typical saturation temperature observed in boiling
chamber testing. These were also taken from experimental
study. Due to the complexity of modeling flow and phase
change in porous media, these constant approximations were
used in place of more complex treatments of the parameters in
order to develop a stable solution. In the future, different
methods of determining these parameters will be used and this
will be discussed later in this work.

30

20

10

50

100

150

200

250

300

350

400

450

q (W/cm2)

Figure 5Comparison of model with experimental data for


82_328_1082

The graph above shows a comparison with another


wick, with dimensions of 82_328_1082. The prediction and
experimental above do not match the experimental data as well,
while it is still reasonably close before the wick begins to dry
out.

Copyright 20xx by ASME

After some investigation, it is believed that the cluster


size has an impact on the internal heat transfer coefficient. It
would appear that with smaller cluster sizes, the internal heat
transfer coefficients decrease.
As a general note, none of the thermophysical
properties of the model presented here are dependent on the
heat flux, which the researchers believe is the reason why the
behavior of the model is mostly linear over the range. As the
model is refined, it will be important to add parameters that are
adjustable with heat flux.
FUTURE WORK
As mentioned throughout the work, there are many
opportunities for improvement with this model, and as this is
merely a starting point to support UCLAs ongoing efforts with
porous media, many additions and edits will be made.
Many of the parameters used in order to solve the model
are taken from empirical relations acquired from colleagues
work. As is the case with any work with porous media, there
are many complex physics involved. Work is underway to
refine the empirical relations used and compare them to similar
mathematic methods used by Kovalev in his original work.
This is the case for the permeabilities, where the original
Kovalev relations compute the permeabilites virtually.
Further, more effort will have to be expended in refining
the implementation of the Kovalev equations since they are
describing a virtual situation in the wick. The cutoff radius R*
does not necessarily exist in the wick so more work will have to
be done with its implementation, which will help with
increasing accuracy over the range of performance curves for
various wicks.
Another parameter which has a significant impact on the
performance of the model versus experimental data is the heat
transfer coefficient. As would be expected, a constant heat
transfer coefficient over the entire domain of heat fluxes is not
adequate for predicting heat transfer performance because it
tends to be a strong function of both geometry and heat flux.
The researchers are currently working on increasing the
accuracy of the heat transfer coefficient by applying work by
Ghiaasiaan (1983), where the heat transfer coefficient was
made dependant on the void fraction .
CONCLUSION
In conclusion, a model is presented which aims to
model the heat transfer performance of biporous wicks. Two
wicks are compared against the model for validation. While the
model performed well, it performed better against a wick with
larger cluster sizes. The authors speculate that this might be a
result of using a constant figure for the internal heat transfer
coefficient. The authors believe that there is a dependence on
cluster size for the internal heat transfer coefficient which is
currently not accounted for. Furthermore, it is believed that
accuracy could be increased with using thermophysical
properties which are more dependent upon the heat flux. These
improvements are currently being implemented and will be
presented in future work.

ACKNOWLEDGMENTS
We would like to acknowledge the support for this work
under DARPA BAA07-36 Thermal Ground Plane (TGP). In
particular, we would like to acknowledge Dr. Tadej Seminic for
suggesting this research in his thesis (Seminic, 2007) and our
program manager, Dr. Avram Bar-Cohen. The views, opinions,
and/or findings contained in this article/presentation are those
of the author/presenter and should not be interpreted as
representing the official views or policies, either expressed or
implied, of the Defense Advanced Research Projects Agency or
the Department of Defense.

REFERENCES.
1.

Semenic T., et al., 2005. Use of biporous wicks to remove


high heat fluxes. Applied Thermal Engineering 28 (2008)
278283
2. Reilly, Sean W, and Ivan Catton. 2009. Improving
Biporous Heat Transfer by Addition of Monoporous
Interface Layer. In ASME Summer Heat Transfer
Conference 2009, 1-8. San Francisco, CA.
3. Semenic, T., Lin, Y. and Catton, I.. 2008.
Thermophysical Properties of Biporous Heat Pipe
Evaporators. Journal of Heat Transfer 130, no. 2: 22602.
4. Wang J. and Catton I., 2001. Biporous heat pipes for high
power electronic device cooling Seventeenth IEEE SEMITHERM Symposium
5. Wang J. and Catton I., 2001. Evaporation heat transfer in
thin biporous media. Heat and Mass Transfer 37 (2001)
275-281
6. Kovalev S.A. 1987. Liquid boiling on porous surfaces.
HEAT TRANSFER Soviet Research, Vol. 19, No. 3 MayJune 1987
7. Liter S. and Kaviany M. 2001. Pool boiling CHF
enhancement by modulated porous layer coating: theory
and experiment. International Journal of Heat and Mass
Transfer 44(2001) 4287-4311
8. Duval F., Fichot F., Quintard M. 2004, A local thermal
non-equilibrium model for two-phase flow with phasechange in porous media. International Journal of Heat
and Mass Transfer 47 (2006) 613-639
9. Miscevic, M, O Rahli, L Tadrist, and F Topin. 2006.
Experiments on flows, boiling and heat transfer in porous
media: Emphasis on bottom injection. Nuclear
Engineering and Design 236, no. 19-21 (October): 20842103.
10. Hartenstine, J.R., R.W. Bonner III, J.R. Montgomery, and
T. Semenic. 2007. Loop Thermosyphon Design for
Cooling of Large Area, High Heat Flux Sources.
Proceedings of IPACK2007
11. Vadnjal, A. High Heat flux Evaporator. PhD
dissertation, UCLA, 2007
12. Ghiaasiaan, S. On Multi-Dimensional Thermal-Hydraulic
and Two-Phase Phenomena During Reflooding of Nuclear
Reactor Cores. PhD dissertation, UCLA 1983

Copyright 20xx by ASME

13. Hanlon, M. a., and H. B. Ma. 2003. Evaporation Heat


Transfer in Sintered Porous Media. Journal of Heat
Transfer 125 (4)
14. Kaya, T, and J Goldak. 2006. Numerical analysis of heat
and mass transfer in the capillary structure of a loop heat
pipe. International Journal of Heat and Mass Transfer 49
(17-18) (August): 3211-3220.
15. Uhle, Jennifer. 1991. Boiling Heat Transfer
Characteristics of Steam Generator U-tube Fouling
Deposits. PhD Dissertation, MIT
16. Chernysheva, M. a., and Y. F. Maydanik. 2009. Heat and
Mass Transfer in Evaporator of Loop Heat Pipe. Journal
of Thermophysics and Heat Transfer 23 (4)

Copyright 20xx by ASME

ANNEX A
PUT ANNEX TITLE HERE

Put text of Annex here

Copyright 20xx by ASME

You might also like