Sei sulla pagina 1di 8

Farzad Houshmand

Impact of Flow Dynamics


on the Heat Transfer of Bubbly
Flow in a Microchannel

e-mail: farzad.houshmand@gmail.com

Yoav Peles1
e-mail: pelesy@rpi.edu
Department of Mechanical,
Aerospace, and Nuclear Engineering,
Rensselaer Polytechnic Institute,
110 8th Street,
Troy, NY 12180

During nucleate flow boiling, the bubble dynamics affect the liquid flow field and the corresponding heat transfer process through several distinct mechanisms. At the microscale,
this effect is different than at the macro scale partly because the bubble dimensions are
comparable to the characteristic length scale of the channel. Since the process involves
several mechanisms, an attempt to isolate and study them independently from one another
is desired in order to extend knowledge. To remove the evaporation effect from the heat
transfer process, noncondensable gas bubbles were introduced upstream of a
1 mm  1 mm heater into a 220 lm deep and a 1.5 mm wide microchannel and the heat
transfer coefficient was measured and compared to single-phase liquid flow. High speed
imaging and micro particle image velocimetry (l-PIV) measurements were used to elucidate the bubble dynamics and the liquid velocity field. This, in turn, revealed mechanisms
controlling the heat transfer process. Acceleration and deceleration of the liquid flow due
to the presence of bubbles were found to be the main parameters controlling the heat
transfer process. [DOI: 10.1115/1.4025435]
Keywords: bubbles, microchannel, convective heat transfer, two-phase flow, micro-PIV

Introduction

In flow boiling, the critical heat flux (CHF) condition declines


with increasing mass quality [1]. Thus, to dissipate very high heat
flux while avoiding CHF, low mass quality is required. Since low
mass quality corresponds to nucleate flow boiling, research pertinent to bubbly flow in micro domains is important to enable high
heat flux applications, such as cooling high performance microprocessors, laser diodes, and high power radars.
Flow nucleate boiling heat transfer is complex and involves
several distinct mechanisms including quenching, evaporation,
and single-phase liquid flow [2,3]. In conventional scale studies,
these three processes are often considered as the primary (if not
the only) mechanisms involving flow boiling heat transfer. However, other mechanismssuch as enhanced mixing away from the
bubble due to rapid bubble growth/collapse, liquid flow modification downstream of an attached or detached bubble, and heating
of a thin liquid layer sandwiched between a large bubble and the
heat transfer surfacemight also be significant. Heat transfer
mechanisms have been discussed before in connection to flow nucleate boiling at the microscale [49]. Due to the large number of
possible mechanisms involved in microchannels (and perhaps in
conventional channels) and the interaction between them, often it
is challenging to identify the dominant mechanism/s.
Ideally, to the extent possible, this can be resolved by studying
each mechanism independently. Eliminating the evaporation process by introducing gas bubbles to the flow at high subcooling
conditions, several mechanisms can be turned off. Betz and
Attinger [10] studied gas segments in liquid flow in microchannels
and demonstrated that heat transfer coefficients can be much
enhanced. Poulikakos et al. [11] studied convective heat transfer
in a segmented immiscible liquidliquid flow in a microchannel
and similar to Betz and Attinger, observed significant enhancement in the heat transfer coefficients. In several other studies, air
1
Corresponding author.
Contributed by the Heat Transfer Division of ASME for publication in the
JOURNAL OF HEAT TRANSFER. Manuscript received October 24, 2012; final manuscript
received August 13, 2013; published online November 7, 2013. Assoc. Editor: Ali
Ebadian.

Journal of Heat Transfer

bubbles [12] and vapor bubbles [13] were used in an attempt to


suppress flow boiling instabilities and enhance heat transfer.
In this study, immiscible gas (air) bubbles were introduced
upstream of a 1 mm  1 mm heater in a 220 lm deep microchannel and the effect of bubbles on the average heat transfer coefficient was investigated. Micro particle image velocimetry (l-PIV)
measurements were performed to elucidate the effect of bubbles
on the carrier liquid flow and to infer the mechanisms affecting
the heat transfer process.

Experimental Setup and Method

2.1 Micro Device. A 220 lm deep, 1.5 mm wide, and


15.5 mm long microchannel was fabricated by bonding two processed Pyrex substrates between an epoxy-covered vinyl layer (Fig.
1). Water entered from an inlet manifold and flowed about 7 mm
before passing over the 1 mm  1 mm heater. Air bubbles were
introduced into the channel through two orifices upstream of the
heater: Orifice I (D 350 lm) and Orifice II (D 250 lm) located
0.5 mm and 4 mm upstream of the heater, respectively.
A 30 nm thick titanium layer along with a 1 lm thick aluminum
layer were deposited on the bottom Pyrex substrate in consecutive
sputtering processes. The heater (1 mm  0.94 mm due to over
etching) and the aluminum vias were formed after removing the
extra material through chemical wet etch processes. A 600 nm
thick SiO2 film was then deposited on the heater and vias for electrical insulation. Inlets and outlet ports were formed by drilling
holes through the wafer. The microchannel was formed by cutting
the pattern of the channel in the vinyl layer and attaching the top
and bottom Pyrex substrates on the opposite side of the vinyl
layer. Before attachment, a couple of holes were drilled on the top
Pyrex substrate as well as the vinyl layer to provide access to the
electrical pads. Finally, individual devices on the attached substrates were separated from each other using a die-saw cutting
machine.
The micro device was seated in a packagebuilt from Delrin
using a computer numerical control (CNC) machinewhich connected the fluidic ports in the microchannel to external fittings and
to the measurement apparatus. The micro device was held in place

C 2014 by ASME
Copyright V

FEBRUARY 2014, Vol. 136 / 022902-1

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/04/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 1

Micro devices schematics

by a cover plate bolted to the package, and sealed through a set of


miniature O-rings (Fig. 2).

2.2 Experimental Setup. Figure 3 schematically depicts the


open-loop experimental setup. Distilled water was pressurized
inside the tank and propelled to the micro device through a control
valve. An Omega FL-3600 series rotameter was used to measure
the flow rate, and a filter was mounted in the flow line to prevent
debris from entering the micro device. To introduce bubbles, a
controllable syringe pump delivered the air stream into the
channel.
A direct current (DC) power supply was connected to the heater
through gold-plated spring-loaded connectors. Voltage across the
heater and the current was measured by two multimeters to calculate the power and the electrical resistance of the heater. The heaters electrical resistance was used to infer the average wall
temperature. All the measurements were recorded in a personal
computer (PC) and processed in National Instrument (NI)
LabVIEW.
A Zeiss inverted microscope (Observer Z1m) was used to visualize the flow. The package and external connections were
mounted on a support structure above the microscope and illuminated by a halogen lamp and a double-pulse 120 mJ Nd:YaG laser

(Solo 120XT from New Wave). A charge coupled device (CCD)


camera (Imager Pro X2M from Lavision) as well as a high speed
complementary metal-oxide semiconductor (CMOS) camera were
used to capture the images. A timing board mounted on the PC
synchronized the CCD camera and the laser pulses.

2.3 Experimental Procedure. Following the fabrication process, the heaters electrical resistance was measured at different
temperatures in a controllable oven and a calibration curve was
generated. With the calibration curve, the heater was also used to
measure the average wall temperature during experiments.
After setting the water flow ratecontrolled by the tanks pressure and a valveand the air flow rate, the voltage and current
across the heater were recorded, and the heater power and average
temperaturebased on the calibration curvewere calculated.
To prevent gaseous cavitation in the microchannel, the distilled
water was initially degassed for several hoursusing a vacuum
pumpand then pressurized by helium.
To estimate heat loss through conduction, the device was vacuumed and the corresponding steady state heater temperatures
were recorded for a range of powers. Heat loss was calculated
based on the heater temperature and then subtracted from the total
power to obtain the effective heat dissipated to the flow.
To capture the sequence of bubble growth and detachment,
high speed camera recorded the images through a 5 magnification lens, while the test section was illuminated by the halogen
lamp. The recorded sequence of images was used to track the
motion of individual bubbles, and using the time interval between
the frames, parameters such as bubble velocity and frequency
were obtained. For frequency calculation, the time interval for 20
to 30 bubbles to pass a certain point was measured and average
frequency was calculated. For micro particle image velocimetry
(l-PIV) measurements, the main flow was seeded with 0.71 lm
fluorescent particles (peak emission wavelength of 612 nm) and
double pulses of Nd:Yag laser with wavelength of 532 nm illuminated the test section. Double frame images were recorded by the
CCD camera through a 10  magnification lens and a 570 nm
high pass filter to eliminate background light. Finally, double
frame images were processed in DavisV software through crosscorrelation algorithmsafter masking out the bubblesto extract
the velocity field. Initially, ensemble average of a set ofusually
500images (used as background image) was obtained and subtracted from individual images to reduce the background noise.
For two-phase images, the gas phase sections of the image were
masked out to provide accurate results in the surrounding area.
Finally, the images were cross-correlated in 64  64 and 32  32
R

Fig. 2 Micro devices package

022902-2 / Vol. 136, FEBRUARY 2014

Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/04/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 3 Schematics of experimental setup

interrogation windows with 50% overlap to extract the velocity


field.
2.4 Data Reduction. For single-phase experiments, Reynolds
number was defined as
Re

 h
qVD
l

(1)

where q (kgm3) is the density, l (kgm1s1) is the dynamic


viscosity, V (ms1) is the average velocity, and Dh (m) is the hydraulic diameter of the channel. Inlet temperature was used for
calculating the thermophysical properties of the liquid flow.
Superficial velocities were used to quantify the flow rates in the
two-phase regime according to
ji

Ci
Ac

(2)

where i denotes the fluid (i.e., water or air), Ci (m3s1) is the volumetric flow rate of fluid i, and Ac (m2) is the cross-sectional area
of the channel.
Based on the volumetric flow rate of the gas and the bubble frequency, the volume of an individual bubble was estimated according to
Cg
B
f

(3)

where B is the bubble volume and f is the bubble frequency. Since


often the bubble covered the entire height of the channel, planar
diameter of the bubbles was estimated as
r
4B
(4)
d
pH
where d is the planar diameter of the bubbles and H is the height
of the channel.
To measure the heat transfer coefficient, heat transfer rate and
the corresponding wall temperature were required. The net heat
transfer rate to the liquid was obtained by subtracting the heat loss
from the total applied power as
Journal of Heat Transfer

Q_ net Q_ tot  Q_ loss

(5)

To compensate for the heat conduction through the thin SiO2


layer, a one-dimensional conduction analysis was performed to
infer the surface temperature based on the heater temperature
according to
tSiO2
Tw Theater Q_ net
A  kSiO2

(6)

where tSiO2 (m) is the SiO2 layer thickness, A (m2) is the heater
area, and kSiO2 (Wm1K1) is the thermal conductivity of the
SiO2 layer. This correction for the wall temperature, however,
was very small (<1  C) in the present study. The average heat
transfer coefficient was then calculated based on the convection
heat transfer equation
 Tw  T0
Q_ net hA

(7)

and the corresponding average Nusselt number was


Nu

 h
hD
kl

(8)

where T0 is the inlet temperature and kl is the thermal conductivity


of the fluid. Because of the developing thermal boundary layer,
the inlet temperature was used in Eq. (7).
The effect of bubbles on the heat transfer coefficient was quantified by comparing the heat transfer coefficient at the presence of
the bubbles to the single-phase liquid flow at the same liquid flow
rate
E

hb  hsp
hsp

(9)

hb and hsp denote bubbly flow and single-phase heat transfer coefficients, respectively.
2.5 Heat Loss and Uncertainties. As discussed before, heat
losses were measured based on the heater temperature and subtracted from total power. The actual heat loss in the presence of
FEBRUARY 2014, Vol. 136 / 022902-3

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/04/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 4

Bubbles before detachment at different flow rates

Fig. 5 Sequence of high speed images from bubble growth and detachment at 6300 fps;
jl 5 0.95, jg 5 0.25 (m/s)

flow might be slightly different from the measured values as discussed in Ref. [14]. However, since the heat loss was between 5%
and 7% for the maximum and the minimum flow rates, this discrepancy was assumed to be negligible.
Standard uncertainty analysis methods were used to estimate
the uncertainties of the reduced data [15]. For flow ratesand
consequently superficial velocity and Reynolds numberuncertainties were associated with the accuracy of the rotameter measurement, which ranged between 63% and 66% for the twophase experiments. The uncertainty in temperature measurements
was less than 1  C. Based on Eq. (7), uncertainties in heat transfer
coefficient and Nusselt number were estimated to vary between
63% and 65% depending on DTs. These values were the total
estimation of the biased and unbiased uncertainties. The standard
deviation of the temperature fluctuations around the timeaveraged value is the unbiased part for the uncertainty, and therefore the true measure of the uncertainty in the enhancement factor,
E. This uncertainty was estimated to be 0.1  C, corresponding to
61% in the enhancement factor. Uncertainty in the velocity,
through the l-PIV measurements, was estimated to be less than
3%, and the uncertainty associated with the bubble frequency was
less than 2%.

gas velocities. A sequence of high speed images is shown in


Fig. 4 illustrating the shape, size, and frequency of injected bubbles over the heater area. As expected, the bubbles size increased
with increased gas flow rates and decreased with increased liquid
flow rates due to higher drag forces. As a result, the bubbles were
squeezed and stretched. The sequence of bubble growth and
detachment from the orifice (for one case) is presented in Fig. 5.
Frequencies of bubbles were measured for different liquid and
gas flow rates and the results are depicted in Figs. 6 and 7 for Orifices I and II, respectively. The frequency of the bubbles injected
from the smaller orifice (Orifice II) was higher compared to that
of larger orifice (Orifice I) by an average of 37% (for high gas
flow rates). This correlates well with the reciprocal ratio of the
orifice diameters (350 lm/250 lm 1.4). Considering the same
liquid and gas flow rates, smaller surface tension force (smaller
perimeter) withstood smaller drag force exerted by the liquid,

Results and Discussion

3.1 Bubble Dynamics. Four different liquid flow rates: 13.2,


18.8, 24.5, and 30.4 (ml/min), corresponding to Reynolds numbers
ranging from 310 to 720, were tested; the gas flow rates varied
from 0.5 to 7 (ml/min). These flow rates provided a range of bubble sizes with respect to heater dimension. As discussed in previous studies (e.g., Refs. [16, 17]), bubbles formed from an air
stream injected into a liquid flow were affected by the liquid and
022902-4 / Vol. 136, FEBRUARY 2014

Fig. 6 Bubble frequency (Orifice I)

Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/04/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

which led to smaller bubbles at higher frequencies. This force balance also explains the higher bubble frequencies at higher liquid
flow rates. At each liquid flow rate, the bubble frequency quickly
reached an asymptotic value with increased gas flow rates. For
very low gas flow rates, the bubble frequency and the bubble size
fluctuated, and therefore, some data points are not presented in the
figures. This effect can be attributed to slow pressure built-up in
the flow linedue to compressibility of the gasrequired to
overcome the initial stage of bubble formation. For heat transfer
measurements in these cases, the average values were used.

Fig. 7 Bubble frequency (Orifice II)

3.2 Heat Transfer and Flow Modification. The effect of gas


bubbles on the heat transfer coefficient was investigated for different gas and liquid flow rates. Heat transfer coefficients for twophase bubbly flow were compared to the single-phase results with
the same liquid flow rate. Since the convective heat transfer process is affected by the fluid flow, l-PIV measurements were performed to obtain the velocity field around the bubbles.

Fig. 8 Average Nusselt number for single-phase water flow

Fig. 9 Temperature change during the bubble injection (Orifice


I); jl 5 1.24, jg 5 0.30 (m/s)

Fig. 10 Effect of bubbles on the heat transfer (orifice I)

Journal of Heat Transfer

Fig. 11 Velocity field around the bubbles injected from


Orifice I

FEBRUARY 2014, Vol. 136 / 022902-5

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/04/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 12 Velocity field around developing bubbles injected from Orifice I; jl 5 0.95, jg 5 0.25 (m/s)

Subsequently, it was used to elucidate the mechanisms controlling


the heat transfer process.
Single-phase heat transfer coefficient measurements were initially obtained for Reynolds numbers ranging from 25 to 2000
(Fig. 8), corresponding to 0.05 < jl < 3.3 (m/s). To validate the
measurements, the results were compared to a simplified analytical solution for average Nusselt number for laminar flow between
parallel plates with a fully developed velocity boundary layer and
a developing thermal boundary layer and constant heat flux
boundary condition [18]. Since the channel had a high aspect ratio
and the heater was located relatively far from the side walls, this
flow condition did not deviate much from the analytical solutions
assumptions. The results compare well with the analytical
solution.
After conducting single-phase experiments, air bubbles were
introduced into the channel and heat transfer coefficients were
measured for different flow rates. Both liquid and gas streams
were maintained at the room temperature (23  C) to isolate the
effect of temperature difference of the secondary flow on the heat
transfer coefficient.
Figure 9 illustrates the average wall temperature changes during
the bubble injection period. After triggering the syringe pump, it
takes a short period for bubbles to reach a steady state condition at
which the stable pressure is built in the flow line. Smaller bubbles
were observed before the steady state condition, which indicates
lower gas flow rate condition. Similar behaviorbut more
gradualwas observed when the syringe pump was switched off,
namely, the bubbles gradually diminished until the single-phase
condition restored. It is worth nothing that Fig. 9 was plotted for a
specific flow combination at orifice I, however, because the injection mechanism was the same for all bubbles, a similar behavior
was observed for Orifice II.
The heat transfer coefficients of bubbly flow were compared to
single-phase heat transfer coefficients for the same liquid flow
rate, and the enhancement factor, E, was calculated. Figure 10
shows the results for bubbles introduced close to the heater from
Orifice I. At low liquid flow rates, the heat transfer coefficient initially increased and then began to decline. The downward trend
led to heat transfer coefficients even lower than those for singlephase liquid flow. At higher liquid flow rates, a similar trend was
observed, although the heat transfer coefficient was higher than
for liquid single-phase case for all the gas flow rates. A peak at
relatively low gas flow rate was observed indicating an optimum
gas flow rate for heat transfer enhancement. At gas flow rates
above the optimum condition, lower temperatures were observed
022902-6 / Vol. 136, FEBRUARY 2014

before and after the steady state condition indicating a higher heat
transfer coefficient at a lower gas flow rates.
l-PIV velocity measurements around the injected bubbles were
obtained to reveal the mechanism of the observed trends. The
results for six different cases taken at midplane of the channel are
presented in Fig. 11. Flow fields at different stages of the bubble
development were also captured and are presented in Fig. 12. (It
should be noted that in Fig. 12, because of the low frequency of
the CCD cameramaximum 15 Hzthe images were not taken
from the same bubble at different stages; instead images from various bubbles at different stages were obtained and the sequence
was reconstructed.) The results identified several mechanisms
controlling the enhanced heat transfer process during bubble formation: (a) acceleration of the carrier liquid flow corresponding to
reduced effective cross-sectional area because of the presence of
bubbles and (b) acceleration around the bubbles. On the other
hand, low velocity regions downstream of the bubbles hinder heat
transfer. The mixing in the shear layer between the high and low
velocity regions can enhance the heat transfer. Oscillations of the
bubbles displace the fluid around them and create high velocity
regions especially when the bubbles detach and the tail recoils
into the bubble (Fig. 12(b)). Another factor that affects the heat
transfer, especially at high gas flow rates, is related to the region
trapped beneath the bubbles. In these regions, a thin liquid film
forms between the wall and the bubble interface. The heat transfer

Fig. 13 effect of bubbles on heat transfer (Orifice II)

Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/04/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

It should be noted that the total effect of the bubbles on the heat
transfer was reflected through the area average heat transfer coefficient over the entire heated area.
At low liquid flow rates, the effect of bubbles on flow acceleration was comparably small and the bubbles were bigger. Consequently, a thin liquid layer over a large segment of the heater
formed and as the liquid layer stagnated, lower heat transfer coefficients were observed. At higher liquid flow rates, the flow acceleration became more appreciable and the bubbles were smaller.
As a result, a larger heater area was influenced by the favorable
effects of the bubbles on the heat transfer process, while a smaller
area was exposed to their adverse effects. In addition, higher bubble frequency helped replenish the liquid film near the heater
more frequently, thus, preventing it from heating up.
To examine the effect of bubbles introduced farther upstream
and presumably past the acceleration stagebubbles with the
same gas flow rates were injected from Orifice II, and their effect
on the heat transfer coefficient was investigated (Fig. 13). At low
liquid flow rates, after an initial increase, the heat transfer
enhancement decreased as the gas flow rate increased, but an
inflection point was observed. At high liquid flow rates, the
enhancement factors increased with increased gas flow rates. Observation of the velocity field (Fig. 14) revealed a low velocity
region in the center of the channelin the vicinity of the bubblesand a high velocity region on the sides. The average heat
transfer coefficient may increase or decrease depending on the
magnitude and the location of the high velocity region with
respect to the heater. Furthermore, as reported in several studies,
e.g., Refs. [19, 20], the liquid film thickness varies with capillary
number. Assuming constant properties, it can be inferred that the
film thickness increases with bubble velocity. This concurs with
the higher heat transfer enhancements at higher liquid flow rates.

Conclusion

Effect of air bubbles on heat transfer characteristics in micro


domains was studied experimentally. Heat transfer coefficient for
different water and air flow rates were measured and compared to
single-phase results. Enhancements up to 16% in the heat transfer
coefficient were observed. Moreover, bubble dynamics and velocity field around the bubbles were studied to elucidate the heat
transfer mechanisms. Results for the injected bubbles from two
different orifices upstream of the heater suggested two simultaneous and competing processes controlling the effect of bubbles on
the heat transfer. Depending on the dominance of each mechanism, the average heat transfer coefficient can be enhanced or hindered. The velocity field modification caused by the bubbles
played a key role in the heat transfer process. Although the impact
of adiabatically formed bubbles on the flow field and single-phase
heat transfer was discussed, and it helps elucidating one of the
many heat transfer mechanisms in the process, further investigation on subcooled boiling with similar approach can further
divulge the processes controlling heat transfer mechanisms. Mixing of the liquid flow caused by the bubbles was also believed to
be important in the heat transfer process, but the extent is dependent on the thickness of the thermal boundary layer.

Acknowledgment

Fig. 14 Velocity field around the bubbles injected from


Orifice II

process is mainly dependent on the thickness and velocity of the


liquid film. When the bubbles are attached to the wall, a stagnation region is formed in the film beneath the bubbles, and the liquid velocity is smaller compared to the single-phase case. This
thin liquid layer can heat up, and if not quickly removed, can elevate the surface temperature and impede the heat transfer process.
Journal of Heat Transfer

This work was supported by the Office of Naval Research (Program Manager Dr. Mark Spector). The authors would like to
acknowledge the staff of the Micro and Nano Fabrication Clean
Room (MNCR) at Rensselaer Polytechnic Institute for their assistance in fabrication of the micro devices.

Nomenclature
A
Ac
cp
Dh

heater area (m2)


cross-sectional area (m2)
specific heat (Jkg1K1)
hydraulic diameter (m)
FEBRUARY 2014, Vol. 136 / 022902-7

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/04/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

E
f
h
H
ji
kl
kSiO2
Nu
P
Q_ loss
Q_ net
Q_ tot
Re
t
T0
Theater
Tw
V
Vb

heat transfer enhancement


frequency (Hz)
average heat transfer coefficient (Wm1K1)
channel height (m)
superficial velocity of fluid i (ms1)
liquid conductivity (Wm1K1)
SiO2 conductivity (Wm1K1)
Nusselt number
pressure (kPa)
power loss (W)
net power (W)
total power (W)
Reynolds number
time (s)
inlet temperature ( C)
average heater temperature ( C)
average wall temperature ( C)
average velocity (ms1)
bubble velocity (ms1)

Greek Symbols
a
Ci
li
q
t
r

thermal diffusivity (m2s1)


volumetric flow rate of fluid i (m3s1)
viscosity of fluid i (kgm1s1)
density (kgm3)
kinematic viscosity (m2s1)
surface tension (Nm1)

Subscripts
b
g
l
sp

bubble
gas
liquid
single-phase

References
[1] Collier, J. G., and Thome, J. R., 1996, Convective Boiling and Condensation,
Oxford University Press, New York.

022902-8 / Vol. 136, FEBRUARY 2014

[2] Krishnamurthy, S., and Peles, Y., 2010, Flow Boiling Heat Transfer on Micro
Pin Fins Entrenched in a Microchannel, ASME J. Heat Transfer, 132(4),
p. 041007.
[3] Basu N., Warrier G. R., and Dhir V. K., 2005, Wall Heat Flux Partitioning
During Subcooled Flow Boiling: Part 1Model Development, ASME J. Heat
Transfer, 127(2), pp. 131140.
[4] Bertsch, S. S., Groll, E. A., and Garimella, S. V., 2009, Effects of Heat Flux,
Mass Flux, Vapor Quality, and Saturation Temperature on Flow Boiling Heat
Transfer in Microchannels, Int. J. Multiphase Flow, 35(2), pp. 142154.
[5] Kandlikar, S. G., 2004, Heat Transfer Mechanisms During Flow Boiling in
Microchannels, J. Heat Transfer, 126(1), pp. 816.
[6] Kim, J., 2009, Review of Nucleate Pool Boiling Bubble Heat Transfer Mechanisms, Int. J. Multiphase Flow, 35(12), pp. 10671076.
[7] Kuo, C.-J., and Peles, Y., 2009, Flow Boiling of Coolant (HFE-7000) Inside
Structured and Plain Wall Microchannels, ASME J. Heat Transfer, 131(12),
p. 121011.
[8] Kuo, C.-J., and Peles, Y., 2007, Local Measurement of Flow Boiling in Structured Surface Microchannels, Int. J. Heat Mass Transfer, 50, pp. 45134526.
[9] Krishnamurthy, S., and Peles, Y., 2010, Flow Boiling on Micro Pin Fins
Entrenched Inside a MicrochannelFlow Patterns and Bubble Departure Diameter and Bubble Frequency, ASME J. Heat Transfer, 132(4), p. 041002.
[10] Betz, A. R., and Attinger, D., 2010, Can Segmented Flow Enhance Heat
Transfer in Microchannel Heat Sinks?, Int. J. Heat Mass Transfer, 53(1920),
pp. 36833691.
[11] Asthana, A., Zinovik, I., Weinmueller, C., and Poulikakos, D., 2011,
Significant Nusselt Number Increase in Microchannels With a Segmented
Flow of Two Immiscible Liquids: An Experimental Study, Int. J. Heat Mass
Transfer, 54(78), pp. 14561464.
[12] Han, Y., and Shikazono, N., 2011, Stabilization of Flow Boiling in a Micro
Tube With Air Injection, Exp. Therm. Fluid Sci., 35(7), pp. 12551264.
[13] Xu, J., Liu, G., Zhang, W., Li, Q., and Wang, B., 2009, Seed Bubbles Stabilize
Flow and Heat Transfer in Parallel Microchannels, Int. J. Multiphase Flow,
35(8), pp. 773790.
[14] Browne, E. A., Michna, G. J., Jensen, M. K., and Peles, Y., 2010,
Experimental Investigation of Single-Phase Microjet Array Heat Transfer,
ASME J. Heat Transfer, 132, p. 041013.
[15] Kline, S. J., and McClintock, F. A., 1953, Describing Uncertainties in SingleSample Experiments, Mech. Eng. (Am. Soc. Mech. Eng.), 75(1), pp. 38.
[16] Elcock, D., Honkanen, M., Kuo, C., Amitay, M., and Peles, Y., 2011, Bubble
Dynamics and Interactions With a Pair of Micro Pillars in Tandem, Int. J. Multiphase Flow, 37(5), pp. 440452.
[17] Elcock, D., Jung, J., Kuo, C.-J., Amitay, M., and Peles Y., 2011, Interaction of
a Liquid Flow Around a Micropillar With a Gas Jet, Phys. Fluids, 23(12),
p. 122001.
[18] Shah, R. K., and London, A. L., 1978, Laminar Flow Forced Convection in
Ducts, Advances in Heat Transfer, (Supplement 1), Academic Press, New York.
[19] Han, Y., and Shikazono, N., 2009, Measurement of Liquid Film Thickness in
Micro Square Channel, Int. J. Multiphase Flow, 35(10), pp. 896903.
[20] Bretherton, F. P., 1961, The Motion of Long Bubbles in Tubes, J. Fluid
Mech., 10, pp. 166188.

Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/04/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Potrebbero piacerti anche