You are on page 1of 42

ACCEPTED MANUSCRIPT

Boiling heat transfer performance in a spiraling radial inflow microchannel cold plate

Maritza Ruiz1, Claire M. Kunkle2, Jorge Padilla3, Van P. Carey2


1
SunPower Corporation, San Jose, CA, USA
2
University of California, Berkeley, CA, USA
3
Google, Inc., Mountain View, CA, USA

Address correspondence to Professor Van P. Carey, Department of Mechanical Engineering,

University of California, Berkeley, CA, 94720, USA. E-mail: vpcarey@berkeley.edu

Maritza Ruiz received her Ph.D. from UC Berkeley in 2015 where she studied

two-phase high heat flux cooling systems, and was the recipient of a National

Science Foundation graduate research fellowship. She is currently a mechanical

design engineer at SunPower Corporation where she develops and studies tools

for producing the next generation of solar photovoltaic modules. She is also a part time adjunct

lecturer at Santa Clara University.

Claire Kunkle received her B.S. degree in Mechanical Engineering from

Santa Clara University in Santa Clara, California, USA. She received her M.S.

degree and is currently pursuing her Ph.D. degree at University of California,

Berkeley, in Berkeley, California, USA. She is a National Science Foundation

Graduate Research Fellow and her research interests include phase change for electronics

cooling, analysis and development of microstructured surfaces, and engineering education.

Jorge Padilla received the B.S. degree in mechanical engineering from the

Massachusetts Institute of Technology in 2005. He received the M.S. and

Ph.D. degrees in mechanical engineering from the University of California,

1
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Berkeley in 2007 and 2014, respectively. His research interests include liquid-vapor phase

change phenomena on micro- and nanostructured surfaces, liquid-solid interfacial energy

exchange, nucleate boiling processes and renewable energy applications. He previously worked

at the solar thermal company, eSolar. He is currently working for Google on thermal

management technologies of data centers.

Van P. Carey is widely recognized for his research on near-interface micro-

scale phenomena, thermophysics and transport in liquid-vapor systems, and

computational modeling and simulation of energy conversion and transport

processes. Since joining the Berkeley faculty in 1982, Professor Carey’s

research has spanned a variety of applications areas, including fuel cells, solar power systems,

building and vehicle air conditioning, forging and casting of aluminum, phase change thermal

energy storage, Rankine cycle power for manned space missions, heat pipes for aerospace

applications, high heat flux cooling of electronics, heat transfer in porous burners, data center

energy efficiency, energy sustainability of information processing, and advanced solar absorber

and turbomachinery technologies for Rankine cycle power generation.

ABSTRACT

This study presents an experimental exploration of flow boiling heat transfer in a spiraling radial

inflow microchannel heat sink. The effect of surface wettability, fluid subcooling, and mass fluxes

are considered. The design of the heat sink provides an inward radial swirl flow between parallel,

coaxial disks that form a microchannel of 300 microns. The channel is heated on one side, while the

opposite side is essentially adiabatic to simulate a heat sink scenario for electronics cooling. To

explore the effects of varying surface wetting, experiments were conducted with two different

2
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

heated surfaces. One was a clean, machined copper surface and the other was a surface coated with

zinc oxide nanostructures that are superhydrophilic. During boiling, increased wettability resulted

in quicker rewetting and smaller bubble departure diameter, as indicated by reduced temperature

oscillations during boiling, and achieving higher maximum heat flux without dryout. The highest

heat transfer coefficients were seen in fully developed boiling with low subcooling levels as a result

of heat transfer being dominated by nucleate boiling. The highest heat fluxes achieved were during

partial subcooled flow boiling at 300 W/cm2 with an average surface temperature of 134 degrees

Celsius. Recommendations for electronics cooling applications are also discussed.

3
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

INTRODUCTION AND BACKGROUND

Exploring innovative technologies for improved cooling capability for the thermal management

of high powered electronic chips or processors, as well as other high heat flux applications, is

crucial for the safety, performance and lifetime of devices. These technologies have the

opportunity to increase performance of devices with heat fluxes on the order of 100 to 1000 of

W/cm2. High heat flux cooling systems must be able to provide high heat dissipation rates at

high conductance levels while maintaining low and uniform wall temperatures. For active

systems, minimal pumping power, as well as simple, reliable and robust designs are desired. For

electronics cooling, Agostini et al. [1] listed a cooling benchmark for modern microelectronics

and power electronics of heat fluxes beyond 300 W/cm2 at chip temperatures of 85°C. The

potential for integrating cooling systems into a 3D chip architecture also provides an attractive

solution to meet processor cooling needs and is an active area of research [2, 3].

Two-phase microchannels are researched for their ability to utilize the large latent

heat of a working fluid during boiling to achieve high heat dissipation rates. These systems have

the ability to achieve desired cooling rates at lower pumping costs and more uniform surface

temperatures when compared to single phase microchannels. However, heat dissipation through

two-phase microchannel heat sinks are limited by the critical heat flux which occurs when the

channel wall dries out due to vapor crowding. Compressibility, instabilities and oscillations in

the flow may lead to early onset of critical heat flux (CHF) and is a major drawback as discussed

by Kim and Mudawar [4]. Additionally, despite the enhanced heat transfer performance, surface

temperatures are dependent on the saturation temperature of the working fluid. Innovative

enhancements and fundamental understanding of the vaporization process in microchannels are

4
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

still necessary to realize the full benefits of microchannel boiling. Kandlikar et al. [2] suggest

future work should include enhancing fundamentals, providing accurate mapping of flow

regimes, using nanoscale surface features for heat transfer enhancement, and developing robust,

inexpensive suppression techniques for mitigating flow boiling instabilities. Ebadian and Lin

also stress the need for increasing reliability and reducing cost by using innovative designs and

optimizing geometries [5].

Two design aspects that have potential for further research in microchannels, is the

use of enhancements in flow boiling with streamwise curvature and the effects of surface

wettability on two-phase flow in microchannels. Studies with flow boiling in curved channels

subject to concave heating have shown enhanced critical heat flux when compared to straight

channels [6-9]. This enhancement is due to increased buoyant forces on vapor due to centripetal

acceleration of the flow as well as increased subcooling at the heated surface due to radial

pressure gradients. This effect was most significant for low subcooling levels where vapor

proportions were higher, with enhancements up to 50% in flow using FC-72 and subcooling of

5°C in Leland and Chow [6]. Increased surface wettability is also expected to have significant

effects on boiling heat transfer mechanisms and prevent dryout of the channel wall. Liu et al.

[10] found a superhydrophilic Si-nanowire coated surface could reduce temperature oscillations

of the channel wall, likely as a result of smaller bubbles. However, they also found it was more

difficult for nucleation to occur at low superheat and high mass flux rates. This follows trends

predicted by Qu and Mudawar [11], which suggest bubble departure diameter decreases with

equilibrium contact angles. There are many additional studies on further advancements to

5
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

microchannels, including sidewall electro-less etching which showed a 150% improvement in

max heat flux over a plain microchannel [12], small-scale (piranha) pin-fins,

These most recent studies fall under a similar surface enhancement model of

improvement as hydrophilic surfaces on microchannels. Hydrophilic surfaces have shown many

desirable characteristics in heat transfer applications and the choice of substrate material, nano-

particle deposition technique, and nano-material is diverse. Liu et al. [10] used silicon nano-

wires on an aluminum substrate while Kim and Mudawar [4] deposited SiO2 nanoparticles

(600nm thick) on a silicon substrate. The hydrothermal method of nanoparticle growth, first

demonstrated by Yang et al. [13], showed how nano-particles could be grown in a liquid growth

solution. This technique of hydrothermal synthesis on copper was subsequently used by Padilla

and Carey [14]. They demonstrated that ZnO nano-particles on a copper substrate could increase

the Leidenfrost point. Increased vaporization heat transfer coefficients were also demonstrated,

which is beneficial for flow boiling in microchannels. A copper substrate was used due to its

favorability as a thermally conductive metal for heat transfer equipment.

Motivated by the need to solve high flux cooling issues, a novel, spiraling radial

inflow microchannel heat sink was developed. This system was designed to study the

enhancements possible in both single phase flow and two-phase flow boiling. Single phase flow

enhancements in this type of heat sink due to unsteady and secondary motions, resulting in high

forced convective heat transfer coefficients, have been experimentally demonstrated in Ruiz and

Carey [15]. For both single phase and two phase flow, the microchannel design was

hypothesized to aid in providing uniform surface temperatures while achieving a high heat

dissipation rate and preventing the maldistribution of flow which can be an issue in rectangular

6
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

microchannel arrays. For two-phase flow, enhancements to boiling similar to those seen in

curved channels with concave heating were expected. A radial inflow heat sink for two-phase

flow can exploit the buoyant force on vapor bubbles due to centripetal acceleration, helping to

draw vapor off the surface towards the outlet and preventing dryout. This type of microchannel

heat sink, first described in Ruiz and Carey [15], consists of a spiraling inflow of fluid sweeping

the surface of two coaxial disks with a constant gap height and exiting through a port in the

center. A diagram of the device is shown in Fig. 1. A single entrance port on the edge of the

device leads into a larger ``feeder" channel region to allow for the development of a more

axisymmetric flow distribution in the microchannel region. The tangential flow helps to provide

a uniform flow distribution over the entire heated surface which aids in temperature uniformity

and prevents the development of hot spots. The simple design with a single inlet and outlet

allows for ease in manufacturing since it does not involve microstructured surfaces or complex

manifolding. The use of microchannels on the order of 100μm allows for high heat transfer

coefficients and a compact footprint. Increased surface wettability is expected to prevent dryout

and increase the performance of boiling heat transfer. This study entails a description of this

type of heat sink and experimental characterization of two-phase boiling flow in a fabricated

device with a constant gap height of 300μm subject to high heat dissipation rates up to 300

W/cm2 with both a plain copper and a ZnO superhydrophilic surface.

METHODS

The fabrication of the device for this study was done using standard machining operations. A

copper alloy 145 heat sink was paired with an insulating cover made from Ultem TM

thermoplastic. The top view of the device is shown in Fig. 2 where the thermoplastic is placed on

7
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

top of the heat sink surface. The body of the heat sink has bored holes on the bottom for three

Watlow cartridge heaters which output 400W at 120 VAC. These were inserted into the bottom

of the heat sink to provide the desired heat levels. Located along the edge of the device are

twelve holes for thermocouple probes, which measure temperature within the copper during

testing.

The design parameters for this copper device are shown in Table 1. A channel height, b,

of 300 µm and a radius of 1 cm was chosen based on typical size of systems desired for

electronics cooling applications. The inlet feeder region is located in the plastic cover and

consists of a uniform channel of cross section, Ac, 2.5mm deep x 1.25 mm wide, which was

designed to provide an inlet tangential to radial flow ratio of ̅ ̅ .

Two different heat transfer surfaces were used for this study; one was a clean, machined,

copper surface, while the other was coated with ZnO nanostructures. The growth of

nanostructures on the surface of this coated test device was done using a technique outlined in

Padilla and Carey [14]. The technique, known as hydrothermal synthesis, involves deposition of

a nanoparticle solution onto a polished copper surface. The surface is then submerged in a

growth solution and baked for ten hours. The resulting nanocoated surface has extremely high

wetting characteristics and contact angles of less than 10 degrees. The coated and uncoated test

pieces were used separately to run identical experiments, allowing for direct comparison of the

coated and non-coated surface in this high heat flux removal study.

Figure 3 shows experimental setup of the test system. Flow was maintained throughout

the system at a constant pressure and flow rate using a gear pump. Flow rate was measured by a

rotameter upstream of the device. Two preheaters were set up downstream of the rotameter to

8
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

provide elevated inlet temperatures. The inlet fluid temperature was measured just upstream of

heat sink. After the fluid passed over the heated surface of the device, the water was condensed

and cooled in a heat exchanger submerged in a cold water or ice water bath. The water then

flowed into a reservoir for pumping back into the system. To prevent losses, insulation was

wrapped around the bottom of the heat sink and around the top exposed copper regions. The

pressure drop in the system was measured through ports in the plastic cover (Fig. 2) using two

Cole-Parmer pressure transducers. The inlet and outlet fluid temperatures as well as the

temperature within the copper device was measured using T-type thermocouple probes from

Omega. The mass flow rate was measured using a 65mm variable area rotameter from Omega

Engineering.

The heat sink was tested over a range of flow rates and heat inputs as shown in Table 2,

using distilled water as a working fluid. For each flow rate, the input power was increased over a

range of heat rates and tested over this range until either (a) the maximum heat rate, based on

experimental design limits, was reached or (b) the temperature of the system began to unstably

increase, indicating dryout conditions on the surface. Once steady state was reached for each

heat rate, temperature and pressure data was recorded for approximately 1.5 minutes and

averaged over this time.

For each flow rate and heat rate pair tested, heat transfer characteristics were found by

averaging over 1.5 minutes of steady state temperature data. For boiling tests, the total heat flux

into the water from the copper was calculated from a discretized integrated total of the heat flux

from two thermocouples at each radial location divided by the total heated copper

area.

9
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

̅ ∫ (1)

Here, the value for thermal conductivity of Copper Alloy 145, = 355 W/mK and the

heated copper surface area, = 4.18 cm2 was used. This average was taken because the

asymmetry in the heaters did not provide uniform heat flux through the thermocouple locations.

However, this method agreed well with the total heat rate into the water for single phase heat

transfer measured from the bulk fluid temperature difference. This data revealed a total average

system loss of 12% of the power input, ranging from 4 to 22% over all tests.

The surface temperature, was calculated by using the average heat flux and the

thermocouple measurement closest to the top at each radial location as:

̅
(2)

The distance from the surface to the top thermocouple is 1.1 mm. Because the

thermocouple distance to the top is small, using an average heat flux rather than a local heat flux

won't result in significant errors in the surface temperature. To calculate the heat transfer

coefficient a predicted bulk fluid temperature at each location was calculated from an

energy balance as the total sensible heat into the water or the saturation temperature.

The heat transfer coefficient can then be found as:

̅
(3)

Because three radial thermocouple locations were recorded, the weighted average of the three

locations was taken using the corresponding radii surrounding each measurement location.

CRITICAL HEAT FLUX MODELS

10
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Critical heat flux values for flow boiling heat transfer subject to high flux levels is typically a

result of vapor crowding on the surface. Bergles and Kandlikar [16] have discussed the distinct

nature of critical heat flux in microchannels, stating that most of the available data for critical

heat flux levels in the literature are likely a result of instabilities. Thus, models based on this data

do not predict the true level of critical heat flux in microchannels. The spiral radial inflow

microchannel device tested in this study is also subject to instabilities during flow boiling.

In order to understand the general trends of critical heat flux during flow boiling,

however, the peak heat flux levels measured were compared to three models for critical heat flux

in macroscale channels or microchannels. The first model considered is based on a study by

Zuber [17] of critical heat flux values for boiling in an extensive pool. This model is based on

vapor coalescence over a horizontal heated surface and does not incorporate any additional

forced convective effects. Thus, this model will likely be a lower limit for the expected critical

heat flux levels in flow boiling. The critical heat flux for this model is found from:

( ) (4)

Lienhard and Dhir [18] previously demonstrated that for water at atmospheric pressure

and for heater sizes on the same order of magnitude to our swirl flow experimental setup, this

Zuber model was within 14% of experimental values of CHF. The second model comparison is

for subcooled flow boiling in macroscale tubes from Celata et al. [19] and based on empirical fits

to CHF data. The critical heat flux for this model is found from:

(5)

Where:

11
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

( ) (6)

{ (7)

Where P is in bar and xe is the level of subcooling at the flow exit. The last correlation

considered in this study, is the saturated flow boiling correlation proposed by Qu and Mudawar

[20] for flow boiling of water and R113 in microchannels, originally derived from correlations of

Katto and Ohno [21].

( ) ( ) (8)

This correlation is also for water at atmospheric pressure and was tested over a range of

subcooling levels. The length to diameter ratio used in Eq. (8) was tested for ratios greater than

150 and our system. Comparing channel diameter to microchannel height, our system had a flow

path to diameter ratio of roughly 160.

RESULTS AND DISCUSSION

Boiling heat transfer experiments enabled two-phase flow characterization of the spiraling radial

inflow microchannel heat sink for heat rates up to 300 W/cm2 using distilled water as a working

fluid. Pressure drop and heat transfer data were acquired for a range of mass flow rates, heat

rates and subcooling levels. Values of the mean mass flux rates, inlet temperatures and

maximum wall heat flux levels tested on the plain copper surface are shown in Table 2. The

same values are shown for the ZnO coated surfaces in Table 3. The mean mass flux rate value

used is found from a flow length averaged mass flux rate through the channel:

∫ (9)

12
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Where G(r) is derived from a mass balance through the device, the inlet mass flow rate is

experimentally measured through the rotameter upstream.

Flow images were acquired for a variety of flow conditions in order to help characterize

the modes of heat transfer and identify distinct regimes of boiling heat transfer. Boiling

experiments in the device revealed that, beyond the single phase flow regime, the flow

throughout the entirety of the device could be characterized into one of three regimes: partial

subcooled flow boiling, oscillatory flow boiling, and fully developed flow boiling. Pressure drop,

heat transfer and flow characteristics in each of these regimes are distinct. Images of the three

flow regimes on the plain copper surface are shown in Fig. 4, while bubble growth on the ZnO

coated surface is shown in Fig. 5

During partial subcooled flow boiling, both forced convective and nucleate boiling

effects are significant. The onset of nucleate boiling in this regime leads to increased heat

transfer coefficients relative to the single phase. Sparse, bubbly flow can be seen with some

bubbles growing to the height of the channel, but major constriction of the flow due to these

bubbles does not occur. At the outlet, the bulk flow is subcooled, although vapor occasionally

gets drawn through the outlet. Pressure drop in this regime does not increase rapidly relative to

the single phase flow. For the highest flow rates tested, the flow throughout the device never

transitioned beyond the partial subcooled flow boiling regime due to high levels of subcooling

and low levels of vaporization on the surface. For the hydrophilic, coated surfaces, less bubbles

are seen to occur during subcooled boiling, indicating either smaller bubble departure size below

visible ranges which then condense back into the liquid or less active nucleation sites. This

13
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

results in the less frequent growth of larger bubbles which tend to sweep through the channel

resulting in effective evaporative heat transfer in the region below the bubble.

During oscillatory flow boiling, the system fluctuates between partial subcooled flow

boiling and fully developed nucleate boiling. Temperature and pressure oscillations were

significant in this regime due to bubble constriction of flow in the microchannel. The percentage

increase in pumping power for this regime was significantly larger than in the subcooled flow

boiling regime. This flow regime is generally undesirable due to the large oscillations in

temperatures and pressure, and thus, design constraints should be established to avoid the

oscillations from occurring. The oscillations are a result of pressure build up behind bubbles

constricting the channel, followed by subcooled liquid flooding the surface and suppressing

nucleation. The surface will then heat back up and continue this sequence in an oscillatory way,

alternating between forced convective dominated heat transfer and nucleate boiling dominated

heat transfer. A decrease in inlet subcooling levels tends to decrease the level of oscillation

because of less aggressive nucleation suppression. The coated surfaces also tended to decrease

the level of oscillations, likely as a result of sparser nucleation or smaller bubble entrainment.

The magnitude of these temperature oscillations, as well as the clear decrease in magnitude for

oscillations on the coated surface is shown in Fig. 6. As shown in Fig. 6, the onset of

temperature oscillations occurred at increasingly higher heat fluxes as the mass flux rate

increased.

During fully developed flow boiling, water vapor rapidly leaves the surface and the flow

does not sustain large oscillations, but small and rapid fluctuations of temperature and pressure

occur as vapor is continuously driven through the exit. Bubbly flow is seen throughout the

14
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

channel with coalescing of bubbles into larger slugs of vapor. This regime was always

characterized by a nonzero average exit quality and was tested up to an exit quality of 0.15.

In general, regardless of the boiling regime, the exit quality was non-equilibrium, in that

portions of the liquid were subcooled, but there was visible vapor departing from the outlet. This

was a result of the generation of large vapor slugs during all of the boiling regimes, and may

result in annular flow through the channel. In this case, one large bubble fills the majority of the

cross-section of the channel and a thin film of liquid evaporates on the surface. This mechanism

results in high heat transfer coefficients due to evaporative rather than nucleate boiling heat

transfer driving the cooling of the surface. Evidence for this is seen in the fact that large regions

of vapor are present in images of boiling yet the surface temperatures indicate dryout cannot be

occurring over these large regions. However, these regions of annular flow may be the

mechanism for initial dryout in the channel, where the wetted surface evaporates into a bubble

and dries out the surface. This mechanism of heat transfer is expected especially during boiling

on the hydrophilic coated surfaces because smaller and sparser nucleation sites are expected as

well as increased wicking of the fluid on the surface underneath large vapor slugs, allowing the

surface to maintain wetted.

Pressure Drop Data

Pressure drop data was tested for all flow rates and heat fluxes. During partial subcooled flow

boiling, pressure drop was not significantly higher than the single phase pressure drop data

predicted in Ruiz [15] Fully developed and oscillatory flow boiling tend to result in increased

pressure drop across the channel. An example of a typical pressure drop curve is shown in Fig.

7. Here we compare the pressure drop across both the coated and uncoated surfaces for a mass

15
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

flux rate of 320 kg/m2s and inlet subcooling level of 80ºC. There is very little difference

between these two data sets, suggesting that the addition of nanostructures on the surface does

not negatively affect pressure drop. However, we see that roughly below 150 W/cm2, the

pressure drop remains low, but increases as two-phase oscillatory flow develops. More detailed

pressure results can be found in Ruiz [22]

Heat Transfer Coefficient Data

Heat transfer coefficients for the lowest mass flux rate tested and four different inlet subcooling

levels is shown in Fig. 8 indicating the dominant boiling regimes for each data point. This

Figure reveals that lower levels of subcooling correspond to higher heat transfer coefficients as

nucleate boiling or evaporation become the dominant method of heat transport over forced

convective effects. Additionally, the heat transfer coefficients level out as nucleate boiling

contributions reach a maximum. This same trend of increasing heat transfer coefficient values for

decreased subcooling is seen for all mass flux rates tested at varying subcooling levels. For

higher mass flux rates, the main difference in the trends is that the heat transfer coefficients did

not increase as rapidly, although higher values of heat transfer coefficients and heat fluxes were

seen, with averaged heat transfer coefficient values up to 6.5 W/cm2K at Gm = 505 kg/m2s and To

= 45°C. Heat transfer coefficients were highest when boiling or vaporization contributions were

largest during fully developed flow boiling.

Figure 9 shows a comparison of the heat transfer coefficients for the lowest mass flux

rate tests on the two ZnO coated surfaces and the plain Cu surface. Both coated surfaces

demonstrated enhancements in heat transfer coefficients relative to the plain copper surface. The

first coated surface tested had the most significant enhancements in heat transfer coefficients, up

16
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

to 25% higher. This enhancement occurs mostly in the oscillatory boiling regime. Studies

indicate that the temperature oscillations in the coated surfaces are decreased likely due to

reduced bubble generating regions and increased evaporative effects as vapor slugs grow in the

channel. For higher mass flux rates, the enhancements in heat transfer coefficients are less

significant as forced convective heat transfer, which is not affected by the surface chemistry,

dominates.

Maximum Heat Flux Results

The maximum heat flux levels recorded during experimental tests for all three surfaces discussed

in this study are listed in Tables 2-3. The maximum value recorded is the peak heat flux levels

reached for at least two of four tests on the system for each corresponding flow rate and inlet

temperature combination. The peak heat were not recorded if either the surfaces temperature

steadily and rapidly increased, indicating thermal runaway or dryout on the surface, or if the

surface temperatures reached higher than about 175°C and were considered unstable for the test

set up. Otherwise the heat was increased at increments of about 21 W/cm2 until the maximum

experimental limit was reached, around 300 W/cm2. At the three highest mass flux rates, the

peak heat flux values were stable and had relatively low, stable surface temperatures, so it is

likely that the critical heat flux at these rates are higher than the experimental maximum tested.

From Table 2 it is evident that lower levels of inlet subcooling were associated with lower levels

of maximum heat flux values. This is likely due to the increased quality of the flow near the

outlet region resulting in higher vapor density which can result in earlier dryout of the passage.

Figure 10 shows a comparison of the peak flux values for the plain copper surface to

three models discussed in the preceding theory section. The experimental limit of about 300

17
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

W/cm2 is also plotted, and the maximum heat flux is likely higher at the highest mass flux rates.

The critical heat flux is represented in dimensionless form as the maximum boiling number of

the flow, defined as:

(10)

For the Zuber model critical heat flux values, the general trend holds but under predicts

the experimental two-phase flow data by as much as 55% (with a mean absolute error of 36%).

For the Celata et al. [19] subcooled flow boiling model, the general trend holds for the

experimental data even though the diameter and subcooling levels experimentally tested are

outside the ``range of applicability" for this model. However, the model always predicts much

lower values than the experimental data suggests. The mean absolute error for this model is

42%, under predicting the critical heat flux by as much as 48%. The saturated flow

microchannel boiling model by Qu and Mudawar [20] generally under predicts the critical heat

flux by up to 28%, yet it provides the most accurate prediction with a mean absolute error of

11%. This model is the most promising predictor of critical heat flux, thus, because it

incorporates microchannel boiling data. For the last two models, the exit mass flux rate of the

experimental device is used as the G value in order to compare to the experimental data. This

would result in the highest possible prediction, but is indicative of the mass flux rate where the

initial dryout may occur because it is the location of the highest quality or lowest subcooling

levels.

Enhancements in critical heat flux due to increased wettability effects were also assessed.

Figure 11 shows the comparison between the peak heat flux values for the plain and coated

surfaces. The hydrophilic coated surfaces were able to reach higher values more consistently,

18
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

indicating slight increases in critical heat flux values for the more wetting surfaces at the lower

mass flux rates. At the highest flux rate tested for hydrophilic surfaces, Gm = 505 kg/m2s, the

peak value reached was about the same for all three surfaces. Two key contributing factors for

the observed increase in critical heat flux rates are postulated. First, studies indicated that,

especially at lower mass flux rates, the level of temperature oscillations during nucleate flow

boiling in the system are decreased. This will tend to reduce the possibility of a temperature

oscillation causing part of the surface to dryout and triggering thermal runaway. Second, the

more highly wetting surface tends to have decreased tendency to dryout due to both wicking

effects and lower contact angles, as can be seen in Fig. 12. Thus, as large vapor regions grow,

wicking of liquid underneath the vapor slugs will tend to keep the surface wetted and prevent

dryout. This is especially important at higher exit qualities which are seen at the lower mass flux

rates, where swashes of vapor may occupy a significant portion of the channel. Surface

wettability can thus allow for higher critical heat flux levels, however, these increases are only

moderate, up to about 16% at the lowest mass flux rate.

It is of interest to explore what other design aspects could be adjusted to increase the

CHF levels further. From the data, it is clear that increasing the flow rate would result in higher

possible flux levels due to larger forced convection heat transfer contributions and higher exit

subcooling levels. This may be undesirable because it could lead to higher pumping penalties

and reduced heat transfer coefficients. Increasing the pressure in the channel is also expected to

increase critical heat flux levels. The increased density of vapor at higher pressures can lead to

more stable vaporization. However, one key disadvantage is that the higher pressure fluid will

have a higher saturation temperature. Both critical heat flux models for flow boiling predict that

19
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

increased channel pressure will result in higher critical heat flux values. Figure 13 shows the

model variation for critical heat flux with pressure plotted relative to the modeled value for

critical heat flux levels at 1 bar. For the correlation by Qu and Mudawar [20] for saturated

microchannel boiling, critical heat flux level enhancements are fairly large at increased pressure,

with values at P = 10 bar at 11 times the value for P = 1 bar. This would indicate the highest

values tested experimentally would be greater than 2500 W/cm2 at an exit pressure of 10 bar.

The Celata et al. model [19] for subcooled flow boiling has less drastic critical heat flux

enhancements at increased pressure. The critical heat flux value is modeled to increase up to 440

W/cm2 at P = 5.5 bar which is the recommended upper limit of validity for the model.

CONCLUDING REMARKS

Flow boiling heat transfer in a spiraling radial inflow microchannel heat sink was tested

and characterized. The device is simple to manufacture and can easily achieve high heat fluxes

with relatively low, uniform surface temperatures. High heat transfer performance was shown

resulting from distinct features of this type of device. Enhanced forced convective coefficients

have been shown in single phase flow due to unsteady and secondary flow motions, with similar

enhancements seen in two-phase flow. Increased drag and buoyant forces on vapor also will tend

to draw bubbles into the flow from all regions of the microchannel. Finally, evaporation on the

surface of vapor slugs during annular flow in the channel leading to enhanced heat transfer

coefficients. The sweeping motion of these vapor regions also entrain other bubbles into the

flow. The annular flow region is especially expected to provide high heat transfer coefficients on

the superhydrophilic surface where fluid wicking underneath bubbles helps to prevent dryout and

promote effective vaporization. On the plain copper surface heat transfer coefficients were seen

20
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

up to 6.5 W/cm2K during fully developed boiling in the channel, revealing promising cooling

performance for electronics cooling or concentrated photovoltaic cooling applications.

Maximum heat flux data is compared to critical heat flux models for flow in

microchannels. Increased surface wettability provides enhanced critical heat flux levels, up to

about 16%. Further potential for enhancements of critical heat flux are proposed including

increased inlet pressure. Heat fluxes on the order of 1000 W/cm2 can be expected with moderate

pressure increases of around 10 bar. For water, this would correspond to an increase in

saturation temperature up to 180°C which would be undesirable for electronics applications, but

could be used for other high flux cooling applications. Alternate working fluids should be

investigated for lower surface temperatures. The ZnO coated surfaces showed durability during

testing, retaining their contact angles throughout the testing process over weeks of use. Using

this type of coating could be beneficial for cooling applications where long term enhancements

resistant to fouling are needed.

NOMENCLATURE

A area, cm2

Bo Biot number

b gap size, μm

CHF critical heat flux, W/cm2

d inner diameter of channel, mm

G mass flux, kg/m2s

g gravity, m/s2

h convective heat transfer coefficient, W/cm2K

21
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

latent heat of vaporization, kJ/kg

k thermal conductivity, W/mK

L heated length of channel, mm

P pressure, bar

̅ mean heat flux, W/cm2

Re Reynolds number

r radial measurement in heat sink, cm

T temperature, ºC

̅ mean angular outer fluid velocity, m/s

̅ mean radial outer fluid velocity, m/s

We Weber number

level of subcooling at flow exit

z axial measurement of channel height, mm

Greek Symbols

ρ density, kg/m3

surface tension, N/m

azimuthal angle of heat sink, degrees

Subscripts

c property of feeder channel

e exit flow property

i property at the inner radius

l liquid property

22
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

m mean flow property

o property at the outer radius

s property at heated surface

sat saturated

v vapor property

w property at heated wall

23
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

REFERENCES

[1] Agostini, B., Fabbri, M., Park, J.E., Wojtan L., Thorne, J.R., and Michel, B., State of the

Art of High Heat Flux Cooling Technologies, Heat Transfer Engineering, vol. 28, no. 4, pp.

258-281, 2007.

[2] Kandlikar, S.G., Colin, S., Peles, Y., Garimella, S., Pease, R.F., Brandner, J.J., and

Tuckerman, D.B., Heat Transfer in Microchannels—2012 Status and Research Needs,

Journal of Heat Transfer, vol. 135, no. 9, pp. 1-18, 2013.

[3] Koo, J.-M., Im, S., Jiang, L., and Goodson, K.E., Integrated Microchannel Cooling for

Three-Dimensional Electronic Circuit Architectures, Journal of Heat Transfer, vol. 127, no.

1, pp. 49-58, 2005.

[4] Kim, S.-M., and Mudawar, I., Review of Databases and Predictive Methods for Heat

Transfer in Condensing and Boiling Mini/Micro-channel Flows, International Journal of

Heat and Mass Transfer, vol. 77, pp. 627-652, 2014.

[5] Ebadian, M., and Lin, C., A Review of High-Heat-Flux Heat Removal Technologies,

Journal of Heat Transfer, vol. 133, no. 11, pp.1-11, 2011.

[6] Leland, J., and Chow, L., Channel Height and Curvature Effects on Flow Boiling from an

Electronic Chip, Journal of Thermophysics and Heat Transfer, vol. 9, no. 2, pp. 292-301,

1995.

[7] Galloway, J.E., and Mudawar, I., Critical Heat Flux Enhancement by Means of Liquid

Subcooling and Centrifugal Force Induced by Flow Curvature, International Journal of

Heat and Mass Transfer, vol. 35, no. 5, pp. 1247-1260, 1992.

24
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

[8] Sturgis, J., and Mudawar, I., Assessment of CHF Enhancement Mechanisms in a Curved,

Rectangular Channel Subjected to Concave Heating, Transactions-ASME Journal of Heat

Transfer, vol. 121, pp. 394-404, 1999.

[9] Sturgice, J.C., and Mudawar, I., Critical heat Flux in a Long, Curved Channel Subjected to

Concave Heating, International Journal of Heat and Mass Transfer, vol. 42, no. 20, pp.

3831-3848, 1999.

[10] Liu, T.Y., Li, P., Liu, C., and Gau, C., Boiling Flow Characteristics in Microchannels with

Very Hydrophobic Surface to Super-Hydrophilic Surface, International Journal of Heat

and Mass Transfer, vol. 54, no. 1, pp. 126-134, 2011.

[11] Qu, W., and Mudawar, I., Prediction and Measurement of Incipient Boiling Heat Flux in

Micro-Channel heat Sinks, International Journal of Heat and Mass Transfer, vol. 45, no.

19. pp. 3933-3945, 2002.

[12] Yao, Z., Lu, Y-W., Kandlikar, S.G., Micro/Nano Hierarchiccal Structure in Microchannel

Heat Sink for Boiling Enhancement, Proc. of IEEE 25th International Conference on Micro

Electro Mechanical Systems, pp. 285-288, 2012.

[13] Yang, Q., Li, Y., Yin, Q., Wang, P., Cheng, Y.-B., Hydrothermal Synthesis of Bismuth

Oxide Needles, Materials Letters, vol. 55, no. 1, pp. 46-49, 2002.

[14] Padilla, J., Carey, V.P., An Experimental Study of the Leidenfrost Transition for Water on

Nanostructured Superhydrophilic Surfaces, Proc. of 15th Int. Heat Trans. Conf., IHTC 15-

9581, 2014.

25
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

[15] Ruiz, M., & Carey, V. P. Experimental study of single phase heat transfer and pressure loss

in a spiraling radial inflow microchannel heat sink. Journal of Heat Transfer, vol. 137 no.7,

pp. 1-8, 2015.

[16] Bergles, A., and Kandlikar, S., On the Nature of Critical heat Flux in Microchannels,

Journal of Heat Transfer, vol. 127, no. 1, pp. 101-107, 2005.

[17] Zuber, N., Hydrodynamic Aspects of Boiling Heat Transfer (thesis), Technical Review,

California University, Los Angeles, and Ramo-Wooldridge Corp., Los Angeles, 1959.

[18] Lienhard, J.H., and Dhir, V.K., and Riherd, D.M., Peak Pool Boiling Heat Flux

Measurements on Finite Horizontal Flat Plates, ASME Journal of Heat Transfer, vol. 95,

pp. 477-482, 1973.

[19] Celata, G., Cumo, M., and Mariani, A., Assessment of Correlations and Models for the

Prediction of CHF in Water Subcooled Flow Boiling, International Journal of Heat

Transfer, vol. 37, no. 2, pp. 237-255, 1994.

[20] Qu, W., and Mudawar, I., Measurement of Correlation of Critical Heat Flux in Two-Phase

Micro-Channel Heat Sinks, International Journal of heat and Mass Transfer, vol. 47, no.

10, pp. 2045-2059, 2004.

[21] Katto, Y., and Ohno, H., An Improved Version of the Generalized Correlation of Critical

heat Flux for the Forced Convective Boiling in Uniformly Heated Vertical Tubes,

International Journal of Heat and Mass Transfer, vol. 27, no. 9, pp. 1641-1648, 1984.

[22] Ruiz, M., Characterization of Single Phase and Two Phase Heat and Momentum Transport

in a Spiraling Radial Inflow Microchannel Heat Sink (Doctoral dissertation), University of

California Berkeley, 2014.

26
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

TABLE 1: Design parameters for experimental device

Parameter Size

b 300 μm

ro 1 cm

ri 0.15 cm

̅ ̅ 6

27
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

TABLE 2: Range of mean mass flux rates, Gm, corresponding inlet temperatures and max heat

flux levels tested for the plain copper device

184 22, 45, 60, 79 169, 150, 150, 150

255 20, 45, 60, 70 213, 195, 193, 193

320 21, 45, 59 236, 240, 236

381 21, 45 265, 235

442 21, 45 277, 259

505 22, 45 295, 291

618 21 301

716 21 293

28
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

TABLE 3: Range of mean mass flux rates, Gm, and corresponding inlet temperatures and max

heat flux levels tested for the ZnO Coated Surfaces

ZnO Coated Surface #1

184 22 196

320 22 271

505 23 293

ZnO Coated Surface #2

184 22 193

255 22 227

320 22 244

505 24 284

29
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Figure 1: Schematic of the heat sink design

30
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Figure 2: Schematic of copper alloy 145 heat sink

31
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Figure 3: Schematic of experimental components used in the laboratory

32
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Figure 4: Flow images in the three flow boiling regimes for the plain copper surface. A: Partial

subcooled flow boiling, B: oscillatory flow boiling, C: fully developed flow boiling

33
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Figure 5: Flow images of bubble growth at 184 kg/m2s and heat flux of 121 W/m2 on the

nanostructure coated surface

34
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Figure 6: Root mean squared of temperature oscillations on the heat sink surface for coated and

uncoated setup at three mass flux rates.

35
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Figure 7: Recorded pressure drop data across the microchannel for the coated and uncoated heat

sink at a mass flux of G = 320 kg/m2s and subcooling of 80ºC

36
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Figure 8: Heat transfer coefficients for a mean mass flux Gm = 184 kg/m2s for four

different subcooling levels

37
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Figure 9: Comparison of the heat transfer coefficients for the ZnO coated surfaces and the plain

copper surfaces for tests and Gm = 184 kg/m2s with subcooling levels of about 80ºC

38
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Figure 10: Experimental maximum heat flux rates tested on the plain copper surface

compared to three critical heat flux models

39
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Figure 11: Comparison of the maximum heat flux for the ZnO coated surfaces to the plain

copper surface for tests with subcooling levels of about 80 ºC

40
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Figure 12: Contact angle comparison of the plain copper surface to the two ZnO

nanostructure coated hydrophilic surfaces. Side view and top view shown. (White spots on

top view are reflected light, not a surface color)

41
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT

Figure 13: Effect of pressure variation on critical heat flux for two models, the value of the

critical heat flux is shown relative to the model estimates at P = 1 bar

42
ACCEPTED MANUSCRIPT

You might also like