Sunteți pe pagina 1din 18

International Journal of Multiphase Flow 59 (2014) 84101

Contents lists available at ScienceDirect

International Journal of Multiphase Flow


j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / i j m u l fl o w

Review

Proposed models, ongoing experiments, and latest numerical


simulations of microchannel two-phase ow boiling
S. Szczukiewicz, M. Magnini, J.R. Thome
Laboratory of Heat and Mass Transfer (LTCM), Ecole Polytechnique Fdrale de Lausanne (EPFL), EPFL-STI-IGM-LTCM, Station 9, CH-1015 Lausanne, Switzerland

a r t i c l e

i n f o

Article history:
Received 16 July 2013
Received in revised form 28 October 2013
Accepted 29 October 2013
Available online 9 November 2013
Keywords:
Two-phase ow
Microchannels
Flow instability
Heat transfer
Evaporation
Numerical simulations

a b s t r a c t
A survey of the most recent work aimed at physically characterizing local heat transfer in ow boiling in
microchannels is presented. This includes recent experimental work, new ow boiling prediction methods, and numerical simulations of microchannel slug ows with evaporation. Some signicant developments in the measurement techniques provide simultaneous ow visualizations and measurements of
2D temperature elds of multi-microchannel evaporators. In particular, information on inlet micro-orices has been gained as well as better ways to reduce such heat transfer and pressure drop data for very
high resolution data (10,000 pixels at rate of 60 Hz). First of all, ow patterns are seen to have a significant inuence on the heat transfer trends in microchannels (just like in macrochannels), and thus need
to be accounted by visualization during experiments and during modeling. A clear distinction between
steady, unsteady, well- and maldistributed ows needs to be made to avoid any confusion when presenting and comparing the heat transfer coefcient trends. In reducing the raw data to local heat transfer
coefcients, the calculated values of several terms involved in the heat transfer coefcient determination
are inuenced by the data reduction procedure, especially the way to deduce the local saturation pressures/temperatures, and may lead to conicting trends and errors approaching 100% in local heat transfer
coefcients if done inappropriately. In addition to experiments, two-phase CFD simulations are emerging
as a tenable tool to investigate the local heat transfer mechanisms, especially those details not accessible
experimentally. In particular, a new prediction method based on numerical simulation results captures
the heat transfer in the recirculating liquid ow between elongated bubbles. Thus, it is shown here that
targeted computations can provide valuable insights on the local ow structures and heat transfer mechanisms, and thus be used to improve the mechanistic boiling heat transfer prediction methods.
2013 Elsevier Ltd. All rights reserved.

Contents
1.
2.

3.

4.

5.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
State-of-the-art of microscale two-phase flow boiling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.
Microchannel flow boiling heat transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.
Infra-red camera measurements applied to microchannels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Numerical simulations of two-phase flow boiling in microchannels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.
Numerical models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.
Literature review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Discussion on the most recent experimental and numerical results of heat transfer studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.
Contribution of numerical simulations to the heat transfer modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2.
Heat transfer coefficient data reduction in multi-microchannels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.3.
Stable and unstable two-phase heat transfer coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Corresponding author. Tel.: +41 021 6935981.


E-mail address: john.thome@ep.ch (J.R. Thome).
0301-9322/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijmultiphaseow.2013.10.014

85
85
85
88
89
89
90
91
91
93
95
99
99

S. Szczukiewicz et al. / International Journal of Multiphase Flow 59 (2014) 84101

1. Introduction
A noticeable global tendency towards miniaturization driven by
the micro-electronics industry is bringing ever greater attention to
multi-microchannel two-phase ow evaporation as the most
advantageous cooling process, utilizing the latent heat of evaporation to extract the heat in an energy efcient manner. As a result of
the enhanced thermal performance compared to other processes,
better axial temperature uniformity (Agostini et al., 2008b), reduced coolant ow rates, and thus smaller pumping powers (Agostini et al., 2007) are obtained. Therefore, two-phase ow cooling
provides an excellent opportunity to continue the progress relative
to Moores law (Moore, 1965) associated with a tremendous challenge of removing the continuously increasing heat uxes dissipated by modern CPUs. The large amount of experimental work,
theory and prediction methods have been reviewed in the past
few years by Thome (2004, 2006), Cheng et al. (2008), Thome
and Consolini (2010) and Baldassari and Marengo (2013). Consequently, the present review has a narrow scope to look at some
new emerging issues regarding experimentation and the targeted
use of numerical simulations to gain local, transient insight into
the two-phase evaporation process and improvement of its heat
transfer models.
Numerous micro-evaporators have been tested over the past
few years. Their reported heat transfer performances, quantied
in terms of local heat transfer coefcients, depend on the data
reduction methods and assumptions each study used. Several aspects, such as determination of the local uid saturation temperature, edge heat losses and heat spreading effects, and ow stability,
need to be more carefully taken into account when comparing and
modeling heat transfer coefcient results. Obviously, only the values calculated in the same manner, when merged together, will
bring adequate conclusions on microchannel cooling capabilities.
Moreover, the experimental techniques for measurements have
some technical limitations due to the small length and time scales
involved in ow boiling within microchannels. For instance, the
time response for thermocouples in point-wise temperature measurements is usually larger than the characteristic time of the
investigated phenomena, whilst experiments with Micro Particle
Image Velocimetry (MicroPIV) still remain a challenging task at
these high ow velocities.
On the other hand, the recent advances on multiphase Computational Fluid Dynamics (CFD) techniques, together with the
increasing processing power of computers, are making numerical
simulations an ever more powerful and reliable tool to provide
new and detailed insights into the local hydrodynamics and thermal features of ow boiling in microchannels. The accuracy of
the gasliquid interface tracking and modeling of interfacial effects
is of primary importance for microscale-aimed computational
methods, since the interface topology plays a fundamental role in
ows within microdevices. Volume Of Fluid (VOF) (Hirt and Nichols, 1981) and Level Set (LS) (Sussman et al., 1994) methods are
indeed the most widely used algorithms to model interfacial ows,
due to their accuracy, robustness and easiness of implementation.
In fact, the cited algorithms only add a color function equation
(to identify each phase) to the single-phase ow equation set,
which includes mass, momentum and energy equations, that are
then solved in a xed computational grid. However, it is important
to remark that while numerical simulations provide an advanced
tool to investigate two-phase ows which may also anticipate
experimental ndings, the development of such computational
methods requires detailed experimental measurements to validate
their new algorithms.
The present paper is organized as follows: rst the most recent
experimental ndings on microscale two-phase ows are reviewed

85

in Section 2, then Section 3 outlines the latest advances in multiphase numerical simulations in microchannels, next Section 4 discusses their mutual contribution and related issues of data
reduction, stable and unstable ow, and hydrodynamics to the heat
transfer coefcient trends, and nally Section 5 summarizes the
main conclusions of this work.

2. State-of-the-art of microscale two-phase ow boiling


In spite of the large number of papers published in the ow
boiling domain, many aspects still need to be better explained in
order to provide a fuller understanding of local two-phase ow
boiling characteristics. Such knowledge is essential to develop
more reliable prediction methods that can be used for designing
new high-performance microchannel heat spreaders for microelectronic and power electronic applications. This section presents
the most recent experimental results in microscale two-phase ow
research aiming to determine the contribution of geometrical
parameters and other two-phase ow aspects on the heat transfer
coefcient trends, which are then discussed in terms of two-phase
ow patterns and ow transitions.
2.1. Microchannel ow boiling heat transfer
Geometrical parameters, such as the hydraulic diameter and the
manifolds material and its shape, may signicantly inuence
microscale two-phase ow results (Hetsroni et al., 2005). For
example, several experimental studies reported signicant heat
transfer enhancement of ow boiling in small (Agostini and Bontemps, 2005; Karayiannis et al., 2010) and narrow channels (Su
et al., 2005) compared to conventional macrochannels. On the
other hand, the measurement reliability decreases with decreasing
tube diameter, as pointed out by Mishima and Hibiki (1996). Additionally, numerous differences between micro- and macrochannels
might be due to inaccurate dimensional measurements in the
microscale (Agostini et al., 2006), where the surface roughness effect on heat transfer at low to medium vapor qualities in the slug
ow regime is noticeable (Agostini et al., 2008d).
In particular, Agostini et al. (2003) showed that the ow boiling
heat transfer coefcient of R134a increased by a factor of 1.74
when decreasing the hydraulic diameter from 2.01 to 0.77 mm.
The increase of heat transfer coefcient at low values of vapor
quality with decreasing channel diameter is associated with the
decrease in the initial lm thickness between the elongated bubbles and the channel wall, as explained by Dupont and Thome
(2005) based on the three-zone model of Thome et al. (2004). For
example, Fig. 1 illustrates the local (width-averaged) heat transfer
coefcient trend versus local vapor quality from inlet to outlet for a
test section with 67 channels of 100  100 lm2 cross-section
(Szczukiewicz et al., 2012b, 2013b), which were measured with a
very ne resolution by means of a high-speed IR camera (for more
details, refer to the following section). In the isolated bubble (IB)
regime, in which bubbles might be smaller than the channel diameter or elongated, the heat transfer coefcient increases, and after
the local maximum, it starts to decrease in the coalescing bubble
(CB) regime. Then, when annular ow (AF) is formed, the heat
transfer coefcient climbs considerably again, dramatically illustrating the importance of ow patterns on the heat transfer process. In the IB regime, heat transfer increases without formation
of dry patches at the end of the elongated bubbles, while in the
CB regime, the heat transfer coefcient decreases due to the onset
of cyclical dryout as the vapor quality increases, which was observed visually by Borhani et al. (2010). The minimum coincides
with the churn ow regime (see the corresponding snapshot in

86

S. Szczukiewicz et al. / International Journal of Multiphase Flow 59 (2014) 84101

The three-zone model of Thome et al. (2004) and the Cioncolini


and Thome (2011) annular ow model for convective boiling were
joined together, applying a new heat ux-dependent transition
from the coalescing elongated bubble regime to the annular ow
regime:

 0:1 1:1
q
Bo
xCBAF 425 v
ql
Co0:5

r


where Bo q=Ghlv is the boiling number and Co r= g DqD2h

Fig. 1. Heat transfer coefcient trend for ow boiling of R236fa in a silicon multimicrochannel evaporator with orices at the inlet of each channel restricting the
ow by 50% (creating some ashed vapor to seed the evaporation process) at the
channel mass ux Gch 2299 kg m2 s1 and the base heat ux qb 48:6 W cm2
(Szczukiewicz et al., 2012b, 2013b). Flow visualization images were recorded by
Revellin (2005) in a sight glass at the exit of a single stainless steel tube of 0.79 mm
diameter.

Fig. 1). Afterwards, the heat transfer coefcient rises in the AF regime, when all the bubbles have coalesced and the liquid has
formed an annular ring with a continuous vapor core in the middle
of the channel. It is an effect of convective boiling across the thinning liquid lm. Therefore, the trend of the heat transfer coefcient
strongly depends on the ow pattern and the ow transitions between them. These transitions are preliminary controlled by the
rate of bubble coalescence and they are commonly dened by
the vapor quality x, e.g. Agostini et al. (2008a), Revellin et al.
(2008) and Ong and Thome (2011a).
It is worthwhile mentioning that the three-zone model of
Thome et al. (2004) predicts the heat transfer coefcient to decrease in the CB regime but to increase in the IB regime (two zones
without the third dryout intermittent zone). Recently, Costa-Patry
and Thome (2012, 2013) have presented a new ow pattern-based
prediction method for heat transfer coefcient in microchannels.

2200
2

17.4 W cm
2
20.8 W cm
2
25.1 W cm
28.5 W cm2
32.6 W cm2
2
36.3 W cm
40.2 W cm2
44.1 W cm2
2
48.1 W cm

1800
1600
1400

ch

[kg m2 s1]

2000

1200
1000
800
0

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.16

0.18

x []
Fig. 2. Vapor quality at the minimum heat transfer coefcient calculated based on
Eq. (1) for the two-phase ow of R1234ze(E) in the test section with the inlet
restrictions of expansion ratio ein;rest WW ch 2 and the base heat ux varying from
in;rest
17.4 to 48:1 W cm2 . The graph was prepared using the experimental data of
Szczukiewicz (2012) considering only stable ows, namely the single-phase ow
followed by two-phase ow without backow and the ashing two-phase ow
without backow operating regimes.

is the connement number, with hlv being the latent heat of vaporization, Dq the difference between the liquid and vapor densities
(respectively ql and qv ), r the surface tension coefcient, g the
gravity acceleration, and Dh the hydraulic diameter. According to
Eq. (1), which predicts the vapor quality at the minimum heat
transfer coefcient (see Fig. 2), the transition is a function of the liquid-to-vapor density ratio, as well as the mass ux G and the heat
ux q, giving results similar to those of Ong and Thome (2011a).
Some modications to the original heat transfer models of Thome
et al. (2004) for elongated bubble ow regime and the annular ow
model of Cioncolini and Thome (2011) were implemented to improve their performance in predicting heat transfer. For instance,
the three-zone model of Thome et al. (2004) was modied by setting the minimum lm thickness to the measured wall roughness
since the roughness breaks the liquid lm. This has been already
proposed in the previous studies of Agostini et al. (2008c) in a silicon test section, Ong and Thome (2011b) in three stainless steel
microtubes, and Vakili-Farahani et al. (2012) in an aluminum multiport tube, while the study of Costa-Patry et al. (2012a) included
both silicon and copper test sections. Table 1 gives more details
on the geometrical specications of the test sections and refrigerants they have investigated.
The above heat transfer prediction method of Costa-Patry and
Thome (2012, 2013) along with the Chen (1966)-like heat transfer
method of Bertsch et al. (2009) seem to be the most accurate ones
available today (Costa-Patry and Thome, 2012, 2013). Fig. 3(a)
demonstrates the heat transfer coefcient trends for both of these
methods in comparison to the experimental results of Szczukiewicz (2012). Firstly, it is seen that the Bertsch et al. (2009) model,
however, does not capture the increasing trend of heat transfer at
higher vapor qualities (corresponding to the AF regime). While, the
prediction method of Costa-Patry and Thome (2012, 2013) predicts
both the trend and the heat transfer coefcients well. This is especially true at low and high values of vapor quality and the location
of the local minimum of the heat transfer coefcient, given by Eq.
(1), representing the CB AF ow transition. The largest discrepancies between the predicted and the experimental values are noticeable at this transition (churn ow), which remains a region of
uncertainty in the multi-microchannel heat transfer studies, and
is very complex to mechanistically model. Together with the tests
listed in Table 1 and those in the recent work of Szczukiewicz et al.
(2012b, 2013b) shown in Fig. 1, their ow pattern based method so
far works for square and rectangular channels with aspect ratios
from about 1 to 10, for single circular channels, for multiport tubes,
and for numerous refrigerants.
Harirchian and Garimella (2012) also used the three-zone model of Thome et al. (2004) (with some modications) as the basis to
predict their experimental data (Harirchian and Garimella, 2008,
2009a,b, 2010) for 7 different microchannel heat sinks. The channel
locations where the ow transformed from bubbly to slug and consequently to annular ow were determined, and then the pressure
drop for each regime occurring along the channel was separately
calculated. They also proposed a ow regime-based method that
provided reliable results, but only for their one uid (FC-77) at
one saturation temperature, and hence only one set of physical
properties; thus its use with any other uid is an extrapolation.

87

S. Szczukiewicz et al. / International Journal of Multiphase Flow 59 (2014) 84101


Table 1
Surface roughness effect in experimental studies.
Reference

Test section

Channel and n geometry

Surface roughness

Test uid

Agostini et al.
(2008d)
Ong and Thome
(2011b)

Silicon micro-evaporator composed


of 67 channels
Single stainless steel microtube

Lch 13:2 mm; W ch 223 lm; Hch 680 lm; W f 178 lm

170 nm

D = 1.03 mm, 2.20 mm, and 3.04 mm

Vakili-Farahani
et al. (2012,
2013)

Extruded aluminum multiport tube


composed of 7 channels

Dh 1:4 mm

595.85 nm, 826.99 nm,


and 796.81 nm
(respectively)
810 nm

R245fa and
R236fa
R245fa,
R236fa, and
R134a
R245fa,
R1234ze(E),
and R134a

Costa-Patry and
Thome (2012,
2013)

Copper micro-evaporator composed


of 52 channels (Costa-Patry et al.,
2012b)
Silicon micro-evaporator composed
of 135 channels (Costa-Patry et al.,
2011a)
Single stainless steel microtube (Ong
and Thome, 2011b)

Lch 20 mm; W ch 163 lm; Hch 1560 lm; W f 80 lm

450 nm

Lch 12:7 mm; W ch 85 lm; Hch 560 lm; W f 47 lm

67 to 90 nm

See above

See above

R245fa,
R1234ze(E),
and R134a
R245fa and
R236fa
See above

Lch : channel length Hch : channel height W ch : channel width W f : n width.

(a)

x 10

x 10
15

2.5
Tran et al. (1996)

w [W m2 K ]

w [W m

K ]

Experimental
2

1.5

Experimental
CostaPatry and Thome (2012)
Bertsch et al. (2009)

0.5

Yu et al. (2002)

10

0.05

0.1

0.15

0.2

0.25

0.3

(b)

0
0

x []

0.05

0.1

0.15

0.2

0.25

0.3

x []

x 10

Fig. 4. Comparison of the experimental results of Szczukiewicz (2012) with the


models of Tran et al. (1996) and Yu et al. (2002).

w [W m2 K1]

2.5

1.5

1
Experimental
New buffer for CostaPatry and Thome (2012)
Bertsch et al. (2009)

0.5

0.05

0.1

0.15

0.2

0.25

0.3

x []
Fig. 3. Experimental two-phase ow heat transfer coefcients of R236fa in the
100  100 lm2 multi-microchannel with ein;rest 4 for Gch 525 kg m2 s1 and
qw 155:3 kW m2 compared to the prediction method of Bertsch et al. (2009), (a)
original model of Costa-Patry and Thome (2012, 2013), and (b) the latter with the
new vapor quality buffer xbuffer 2xCBAF . Figure extracted from Szczukiewicz
(2012).

They did not unfortunately present any graph with trend lines of
the heat transfer coefcient data versus local vapor quality, nor
the trends of their predictions. Hence, the method of Harirchian
and Garimella (2012) was not compared with the experimental

results in Figs. 3 and 4 because an initial guess of the liquid lm


thickness is required to start their heat transfer calculation (and
they note that the stability of the calculation can depend on the
choice).
To better handle the churn ow regime separating the CB and
AF regimes and reect the experimental U-shaped heat transfer
coefcient trend, as presented in Fig. 3(b), Szczukiewicz (2012)
proposed a new vapor quality buffer for the width of the CB-AF
transition as an update to the heat transfer prediction method of
Costa-Patry and Thome (2012, 2013). Still, the rise in the AF heat
transfer coefcients is more rapid than suggested by their model
and this aspect needs to be further investigated. Also in Fig. 4,
the U shape of the heat transfer coefcient trend can be observed,
however, in case of the method of Tran et al. (1996), it is hardly
noticeable due to the large scale of the y axis. In this gure, the
models of Tran et al. (1996) and its modied version presented
by Yu et al. (2002) are compared with the experimental data of
Szczukiewicz (2012) and the equations to determine the heat
transfer coefcient for both of them are respectively given below:

a 840; 000 Bo2 Wel

0:3

ql
qv

0:27

a 6; 400; 000 Bo2 Wel

0:4
2

ql
qv

0:2
3

88

S. Szczukiewicz et al. / International Journal of Multiphase Flow 59 (2014) 84101

where Bo is the boiling number dened above and


Wel G2 Dh =ql r is the Weber number. In Eq. (2), the lead constant
was changed accordingly to Thome (2010), who noticed a probable
typographical error in the original publication of Tran et al. (1996).
Secondly, these constants for Eqs. (2) and (3) are different, which
explains the difference in their extrapolated simulations presented
in Fig. 4. One order of magnitude of difference in this constant is signicant especially taking into account the fact that the tested geometry and the range of experimental conditions were similar,
although it is worthwhile to note that the refrigerant-based correlation was adapted to t to the experimental results obtained for
water. Finally, both of them state that heat transfer coefcient is
not a function of G and x, which is in contrast with common trends
of many studies as pointed out in Thome (2010).

and the transient boundaries between them were given. Analogous


observations were made in the intercrossed array of triangular
microchannels and transverse trapezoidal microchannels (Xu
et al., 2006).
An experimental investigation of ow boiling of 2-propanol and
water in 50  50 lm2 Cyclo Olen Polymer COP parallel channels
was performed by Hardt et al. (2007). In their experiments, the
temperature measurements were done by means of an IR camera,
which was calibrated using a single thermocouple that had an
accuracy of 1.5 K. Although, the authors stated that their camera
itself measured the temperature with a thermal resolution of

(a)

2.2. Infra-red camera measurements applied to microchannels


As a replacement of point-wise thermocouple and diode temperature measurements, IR thermography has started to be more
extensively explored in the past few years for microchannel heat
transfer experiments, providing a very ne spatial resolution and
instantaneous measurement of heat transfer coefcients. Some
applications of IR cameras to convective heat transfer experiments
are highlighted below.
Hapke et al. (2002) used IR thermography to determine local
heat transfer coefcients of water and n-heptane evaporating in
rectangular microchannels with sizes ranging from 300 to
700 lm applying a classical one-dimensional (1D) heat conduction
approach. Diaz et al. (2005) and Diaz and Schmidt (2007b,a) extended the data base of Hapke et al. (2002) to include additional
geometries, channel sizes, and uids. In order to improve the accuracy of their IR temperature measurements, the test sections were
painted black with an emissivity of 0.95. Due to technical difculties, the change in uid saturation temperature T sat along the channel was neglected and the uid temperature was estimated as an
average of the inlet and outlet temperatures (that is a xed value
of T sat for calculating all local heat transfer coefcients), which severely inuences the values and trends when the fall in T sat is signicant with respect to the local wall superheat.
Hetsroni et al. (2001, 2006) focused on investigating triangular
parallel multi-microchannels down to hundreds of microns in size
under uniform and non-uniform heat ux conditions. An IR camera
with an accuracy of 1 C was employed to monitor the temperature variations across the uniformly heated test section wall. These
were found to be associated with hydraulic instabilities within the
test section and they were signicantly enhanced for non-uniform
heating. IR camera temperature measurements were also carried
out by Hetsroni et al. (2001, 2003) to study the explosive vaporization of water in microchannels with periodic wetting and dryout
behavior. Furthermore, Hetsroni et al. (2002) showed that temporal temperature and pressure uctuations of the uid Vertrel XF
(DuPontTM ) evaporating in their test section corresponded to each
other and they caused a reduction in heat transfer, i.e. unstable
ow penalized heat transfer.
Xu et al. (2005, 2006) measured (by means of an IR camera)
thermal oscillations for a uniformly heated surface of their silicon
test sections, although the recording rate of their IR image system
was not sufcient to observe the thermal ow patterns in greater
detail. In particular, Xu et al. (2005) examined ow boiling of acetone in silicon parallel triangular microchannels, each having a
hydraulic diameter of 155.4 lm. Similar to previous studies, the
back of the test section was covered by a thin layer of black lacquer
with an emissivity of 0.94 that improved the accuracy of the IR
temperatures to within 0.4 C. Three zones of a full boiling cycle
were described: (i) liquid relling stage, (ii) bubble nucleation,
growth and coalescence stage, and (iii) transient annular ow stage

Created based
on the flow
observation
from the top

(b)

Obtained by means of the


IR camera placed at the bottom

(c)

(d)

Flow direction
Fig. 5. (a) Two-phase ow operational map for R236fa in the micro-evaporator
with the 50 lm-wide, 100 lm-deep, and 100 lm-long inlet restrictions ein;rest 2,
where:
single-phase ow,
single-phase ow followed by two-phase ow
with backow,
(b) single-phase ow followed by two-phase without backow
(desirable operating regime),
two-phase ow with backow triggered by
bubbles formed in the ow loop before the test section, and
(c) ashing twophase ow without backow (the most desirable operating regime). (b) Photograph
of the experimental ow boiling test facility with the optical system. This
operational map and the photo of the test rig were extracted from Szczukiewicz
et al. (2013a). The thermal maps (c) and (d) were recorded for the two-phase ow of
R245fa in the test section of ein;rest 2 (Szczukiewicz, 2012).

S. Szczukiewicz et al. / International Journal of Multiphase Flow 59 (2014) 84101

0.08 K at 30 C. Patil and Narayanan (2005, 2006) utilized IR thermography to obtain spatially-resolved temperature measurements
in a single silicon, uniformly heated microchannel of the same
width as Hardt et al. (2007), but having a higher aspect ratio. Both
the wall and near-wall uid temperatures (here water) were measured, depending on the opacity of the channel wall. The IR camera
calibration was done implementing intensity maps during water
ow at the temperature of 23.5 C, although this does not explain
measurements at other temperatures. An IR camera was also used
(together with a high-speed ow visualization camera) by Barber
et al. (2009, 2011) to study ow boiling instabilities of n-pentane
in a single rectangular microchannel having a hydraulic diameter
of 727 lm under uniform heating. The accuracy of their IR temperatures was reported to be 1 C.
Recently, Szczukiewicz et al. (2012a,b, 2013b) introduced a new
in situ pixel by pixel technique to calibrate the raw infra-red image
signals with an accuracy of 0.2 C, and thus converting them into
accurate two-dimensional temperature elds of 10,000 pixels over
the heated surface of the silicon micro-evaporator. The test facility
and the test sections were designed such that simultaneous twophase ow patterns (through a transparent Pyrex cover plate) and
temperature visualizations in 67 microchannels of 100  100 lm2
cross-sectional areas were possible. To this aim, they used a highspeed video camera placed above the test section and a high-speed
infra-red (IR) camera below, as shown in Fig. 5(b). Their results for
R245fa, R236fa, and R1234ze(E) at a variety of the channel mass
ux Gch and the base heat ux qb were categorized into 8 different
two-phase ow operating regimes, among which 2 examples are
illustrated in Fig. 5, where T IR is a temperature measured at the base
of the test section, y indicates an axis perpendicular to the ow
direction, while z denotes the distance form the channel entrance.
This description applies also to the other thermal maps in this paper. The ashing two-phase ow without backow operating regime, shown in Fig. 5(d), was identied as the most desirable one,
since it provided the best spatio-temporal temperature and ow
uniformities. Moreover, several new two-phase ow operational
maps were developed for the two-phase ow of refrigerants
R245fa, R236fa, and R1234ze(E) owing in the test sections with inlet restrictions (used for ow stabilization, as suggested by Kosar
et al. (2006); Agostini et al. (2008c); Park et al. (2009)) and without
any inlet restrictions. An example of such an operational map for
R236fa owing in the test section with the inlet restrictions of the
expansion ratio of ein;rest 2 is presented in Fig. 5(a). From an engineering stand-point of view, these operational maps are very
important for specifying the most advantageous operating conditions, at which the ow is always going to be stable for eventual actual electronics cooling applications. More details on the two-phase
ow operational study of refrigerants in multi-microchannel
evaporators for future high-performance 3D-ICs can be found in
Szczukiewicz (2012) and Szczukiewicz et al. (2013a).

3. Numerical simulations of two-phase ow boiling in


microchannels
In spite of the high resolution of IR cameras and high-speed video cameras, many dynamic and localized aspects of these twophase ows and their heat transfer mechanisms remain elusive,
or based on conjecture, rather than based on proven principles.
This is specically where numerical simulation of these ows can
provide strategic insight into this complex process and provide
the basis for proposing models that incorporate these newly identied aspects. This section introduces this recent work with rst an
overview of numerical two-phase models and the following section gives some insight into their use to improve a mechanistic
model.

89

3.1. Numerical models


The computation of evaporating ows in microchannels requires the conventional multiphase algorithms for CFD to be coupled with specic physical and numerical models to accurately
capture the interfacial effects which become dominant in the
microscale. These models are briey summarized below along with
a unied formulation of the ow equations for the Volume Of Fluid
and Level Set methods. Then, a review of the pertinent numerical
ndings so far is presented.
The VOF and LS methods adopt a single-uid mathematical representation of the two-phase ow, where the gas and liquid phases
are treated as a single uid with variable properties across the
interface. A color function c is used to identify each phase on the
computational domain and a Heaviside step function I is built
according to the values of c to compute the uid properties throughout the computational domain. For instance, for each computational
cell of the ow domain the density is calculated as follows:

q ql qv  ql I

where 0 6 I 6 1 and qv ; ql are the vapor and liquid phase specic


densities. The liquidvapor interface is then identied as a transition region for the uid properties and it has a nite thickness of
23 computational cells. When the ow problem is treated as
incompressible, which is true provided that the variation of the vapor density due to the pressure drop that occurs along the microchannel is negligible, the mass conservation equation is expressed
as:



_
rum

qv

ql

dS

where u denotes the uid velocity. The term at the r.h.s. represents
_ is the interphase
the mass source due to the evaporation, where m
mass transfer and dS is a delta function which is non-zero only at
the interface and its expression depends on the specic multiphase
algorithm adopted.
The momentum equation for Newtonian uids in laminar ow,
appropriate for microchannels, takes the following form:

@qu
r  qu  u rp r  lru ruT  qg
@t
rjndS

with t being the time, p the pressure, q and l the single-uid density and viscosity to be computed as it is shown in Eq. (4), and g the
gravity vector. The last term on the r.h.s. represents the surface tension force for a constant surface tension coefcient, where j and n
identify the local interface curvature and unit normal vector. The
interface topology is not available explicitly in VOF and LS methods,
but it can be derived by means of the color function eld as
n rc=jrcj and j r  n as originally proposed by Brackbill
et al. (1992). However, the computation of the interface curvature
by means of derivatives of the color function is known to be of poor
accuracy for VOF methods as pointed out by Cummins et al. (2005),
thus generating errors in the estimation of the surface tension force.
Since the surface tension is a dominant force in the microscale, the
accuracy of its calculation is fundamental to obtain reliable numerical results. Hence, it is preferable that VOF schemes include specic
algorithms for the reconstruction of the interface topology such as
the parabolic tting of Renardy and Renardy (2002), the Height
Function method (Cummins et al., 2005), or coupled LS and VOF
schemes (CLSVOF) (Sussman and Puckett, 2000).
The energy equation to be solved is given by:

@qcp T
r  qcp uT r  krT s
@t
_ lv  cp;v  cp;l TdS
: ru  mh

90

S. Szczukiewicz et al. / International Journal of Multiphase Flow 59 (2014) 84101

where T indicates the temperature, cp and k are respectively the single-uid specic heat at constant pressure and thermal conductivity, while s represents the shear stress tensor. The second term on
the r.h.s. represents the viscous heating which may become important for very small channel sizes (negligible for the sizes considered
here), while the third term implements the enthalpy sink due to
evaporation, that of the vapor created and of the liquid that disappears. Usually, the temperature variations are sufciently small
such that the physical properties of the phases are considered constant. The set of the ow equations is then completed by the transport equation for the color function:

_
m
@c
u  rc dS
@t
q

which is used to update the position of the interface and then the
mixture uid properties as time elapses.
To complete the formulation, an evaporation model to express
_ as a function of the local temperathe interphase mass transfer m
ture and pressure is necessary. If microscale effects on mass transfer are neglected, the interface is assumed to be at the saturation
temperature, such that the temperature eld is continuous across
the interface, and the mass transfer can be computed as
_ krT  n=hlv (Mukherjee and Kandlikar, 2005; Mukherjee,
m
2009; Mukherjee et al., 2011; Lee et al., 2012; Suh et al., 2008).
However, as the scale of the problem is reduced, interphase resistance and disjoining and capillary pressures tend to create a discontinuity in the temperature and pressure elds across the
interface, thus generating an interfacial resistance to mass transfer
which decreases the evaporation rate. In this direction, Schrage
(1953) derived a relationship to express the interphase mass transfer as a function of the local temperatures and pressures of the liquid and gas phases at the interface, while Wang et al. (2007)
showed that the following linearized expression is a good
approximation:

_ C T T i  T v C p pl  pv
m

as long as the temperature difference between the liquidvapor


interface and the equilibrium saturation temperature is below 5 K.
C T and C p are constants of the model which only involve the uid
properties, see Wang et al. (2007). T i is the liquidvapor interface
temperature, T v is the vapor temperature at the interface, and
pl  pv is the liquidvapor pressure jump. Eq. (9) has been implemented by various authors to set-up a CFD solver for boiling and
condensing ows in microchannels such as Magnini et al. (2013a),
Kunkelmann and Stephan (2009) and Nebuloni and Thome (2010).
In ow boiling in microchannels, liquid dryout at the channel
wall may occur as a consequence of the evaporation of the liquid
lm which surrounds an elongated vapor bubble (Thome et al.,
2004). In CFD computations, the modeling of wall adhesion involves the implementation of a (static or dynamic) contact angle
model. The contact angle is a condition on the direction of the
interface normal vector at the solid-liquidvapor three-phase contact line, and hence on the local color function eld for LS and VOF
methods. In the former, the contact angle is introduced as a boundary condition for the color function eld at the wall (Mukherjee
and Kandlikar, 2005; Suh et al., 2008; Lee et al., 2012), while in
the latter it is usually enforced by adjusting the components of
the unit normal vector involved in the surface tension force term
in Eq. (6) (Brackbill et al., 1992; Renardy et al., 2001; Afkhami
and Bussmann, 2008). Another issue connected to wall adhesion
is the microlayer evaporation process, which is known to increase
dramatically the heat transfer in the contact line region in nucleate
boiling, and hence may be a signicant heat transfer mechanism at
the perimeter of intermittent dry patches in microchannels. Usually, the contact line evaporation is modeled by solving a fourth

order differential equation for the lm thickness evolution in the


so-called micro-region (Stephan and Busse, 1992), which is discretized by a separated computational grid. This solution is then included in the ow problem for the macro-region by means of
specic source terms (Kunkelmann and Stephan, 2009; Li and Dhir,
2007). However, its short life in the cyclic heat transfer process and
small footprint in elongated bubble ow tends to reduce it to only
a small inuence.
3.2. Literature review
The numerical investigations of ow boiling in microchannels
have focused initially on the fundamental aspects of the ow, such
as bubble dynamics and the ow eld induced by the evaporation
process. In this light, Mukherjee and Kandlikar (2005) simulated
the growth of a spherical bubble during ow boiling in a square
microchannel and observed that the presence of the channel walls
tended to elongate the bubble, which grew with an exponential
time-law, while vapor patches appeared at the centerlines of the
channel walls due to the dryout of the trapped liquid lm. Li and
Dhir (2007) analyzed the bubble growth and detachment from a
at wall in ow boiling conditions and observed that the departure
diameter of the bubble decreased as the bulk ow velocity increased and it increased with the wall tilt angle. Augmentation of
the ow velocity suppressed the gravity effect on the bubble
dynamics.
Subsequently, as the experimental ndings highlighted specic
issues which required more detailed knowledge of the local phenomena to clarify, e.g. ow instabilities, dominant heat transfer
mechanisms, heat transfer enhancement, more targeted numerical
studies began to be conducted. The stability of the ow during
evaporation in parallel microchannels was investigated by Suh
et al. (2008), who showed that backow occurred when the formation of vapor bubbles was not simultaneous in adjacent microchannels, and thus led to a distribution of the heat ux which was not
uniform on the surface of the heater. Dong et al. (2012) studied the
effect of single and multiple growing vapor bubbles on the uid
ow behavior and heat transfer, and reported that the bubble formation process induced a ow resistance which increased with the
growth of the bubble, disappeared with the bubble departure, and
was strongly affected by the presence of multiple bubbles. Mukherjee and Kandlikar (2009) analyzed the effect of inlet constrictions to prevent the backow growth of vapor bubbles in a
square microchannel and showed that, despite the positive effect
in stabilizing the ow, they generated a high pressure drop and reduced the efciency of the thin-lm evaporation mechanism.
Diverging microchannels in the direction of the desired ow were
proposed as a better solution. However, the analysis of pressure
drop needs to be applied to the entire two-phase cooling loop to
ascertain the true impact of such orices.
The contribution of the evaporation of the thin liquid lm surrounding an elongated bubble during ow boiling was the subject
of the work of Mukherjee (2009). He simulated the growth of a vapor bubble in contact with the heated surface of a microchannel
and observed that the formation of a thin layer of liquid between
the bubble and the channel wall, which was promoted by smaller
contact angles, increased the wall heat transfer and thus reinforcing the interpretation that the heat transfer in microchannels is
thin-lm evaporation dominated. Magnini et al. (2013a) studied
the hydrodynamics and heat transfer given by the ow boiling of
single elongated bubbles in a circular microchannel and observed
a strong increase of the heat transfer coefcient in the vapor bubble region due to the evaporation of the thin liquid lm, including a
thermally developing length effect. They obtained good predictions
of the heat transfer coefcient by means of a model based on transient heat conduction across the lm, which was an extension of

91

S. Szczukiewicz et al. / International Journal of Multiphase Flow 59 (2014) 84101

4. Discussion on the most recent experimental and numerical


results of heat transfer studies
Some critical discussion on experimental and numerical work is
presented below, whose objective is to highlight some of the
achievements but also pitfalls of the current state-of-the-art and
also to illustrate how experimental, theoretical modeling and
numerical simulations can work hand-in-hand to resolve microchannel evaporation research issues.
4.1. Contribution of numerical simulations to the heat transfer
modeling
After almost ten years since its publication, the three-zone
model of Thome et al. (2004) with its related updates (Agostini
et al., 2008d; Harirchian and Garimella, 2012; Costa-Patry and
Thome, 2012, 2013; Costa-Patry et al., 2012b) is still the best performing boiling heat transfer prediction method for slug ow boiling in microchannels in circular and non-circular channels,
covering numerous uids and channel sizes down to 85 micron
widths. Even so, numerous simplications are assumed in the
two-phase ow structure in developing this mechanistic model
and numerical two-phase simulations constitute a unique tool to
investigate the local ow phenomena inuencing the micro-heat
transfer processes involved, and thus can contribute to the
improvement of the sub-models. For example by means of computations, Magnini et al. (2013a) already proved that, for a liquid lm
thickness of 1/20 of the channel diameter, the thermal inertia

effect is not negligible when modeling the heat transfer in the liquid lm region by assuming one-dimensional heat conduction.
By adding a transient term to the original three-zone models formulation (which was developed for much thinner liquid lms, on
the order of 1/100 of the channel diameter), the time-law of the
heat transfer coefcient given by CFD simulations was predicted
satisfactorily. This inuence is more signicant for shorter bubbles
since the thermal boundary layer development length then plays a
larger role; hence this effect is more important in the IB regime
where bubbles are still relatively short L < 2D but less so in the
CB regime where coalescence of bubbles results in mostly long
bubbles L > 5D.
With the aim of improving the modeling of the heat transfer in
the liquid slug zone of the three-zone model, consider two bubbles
owing and evaporating in sequence in a circular microchannel.
They were simulated using the numerical framework already discussed above by Magnini et al. (2013a), taking advantage of a 2D
axisymmetrical formulation to limit the computational time of
the simulations. For horizontally oriented channels, this is a valid
assumption provided that Co > 1, as observed experimentally by
Ong and Thome (2011a), and hence only working conditions
matching such criterion were chosen. The reliability of the solver
in modeling the ow of axisymmetrical elongated bubbles was assessed by a positive comparison of the liquid lm thickness
trapped between the bubbles and the wall in adiabatic conditions
with the Han and Shikazono (2009) correlation, the latter based on
their measurements with a highly accurate oscillating microscope
technique. The validation of the numerical results in ow boiling
conditions were also proven by comparing the position of the nose
of the growing bubble against time with the theoretical model derived by Consolini and Thome (2010) for the ow of coalescing
bubbles in microchannels. Fig. 6 shows the computational and theoretical results for three different operating uids, namely R113,
R245fa and R134a, under similar operating conditions. The comparison is quite positive, where the increasing underestimations
of the model at the highest time steps are due to the assumption
that the bubble grows only by absorbing the wall heat ux, while
in the numerical simulations the bubble also grows because it receives the sensible heat of the superheated liquid by evaporation
across the nose of the bubble.

12

Position of the bubble nose [mm]

Thome et al. (2004) steady-state three-zone model, and conrmed


the thin-lm evaporation dominance on heat transfer within
microchannels. Mukherjee et al. (2011) also attempted to explain
some of the experimentally observed trends for the boiling heat
transfer coefcient, see Agostini and Thome (2005) for an earlier
review. The results of their computational study suggested that
the heat transfer coefcient increased with the heat ux as it augmented the bubble growth rate and then the velocity of the liquid,
which was pushed against the channel wall by the growing bubble.
The authors observed that the liquid mass ux had only a little effect on the heat transfer and this was justied by the high velocities generated by the bubble growth process, which suppressed the
effect of the liquid inlet velocity. This can also be explained by
assuming that the heat conduction is the dominant heat transfer
mechanism across the liquid lm. Their conclusions also match
those that can be deduced from analysis of the three-zone model,
where bubble frequency and velocity play important roles while
mass velocity does not inuence heat transfer (completely counter-intuitive to single-phase ows). Magnini et al. (2013b) performed simulations of the ow boiling of multiple bubbles
within a microchannel and showed that the interaction among
sequential bubbles generated bubbles of different lengths, velocities and thickness of the liquid lm trapped between the liquidvapor interface and the channel wall. This led to different heat
transfer performances, in particular the time-averaged heat transfer coefcient for the trailing bubble cycle was much higher than
that of the leading one (about 60% higher).
The enhancement of the heat transfer performance of a microchannel of square cross-section was the objective of the work of
Lee et al. (2012), who performed numerical simulations to optimize the design of transverse ns on the channel. It was found that
the heat transfer was signicantly improved by those solutions
which promoted thinner liquid lms trapped between the bubble
and the channel walls and larger liquidvaporsolid interface contact regions. It is not known however if this will promote premature dry patch formation and CHF.

10

4
R113
R245fa
2
R134a

10

12

14

16

Time [ms]
Fig. 6. Positions of the bubble nose against time in ow boiling conditions, given by
simulations (solid lines) and Consolini and Thome (2010) model (dashed lines).
Simulation conditions: D = 0.5 mm, q 20 kW m2 ; G 600 kg m2 s1 (R113,
R245fa) and 500 kg m2 s1 (R134a), T sat 50  C (R113, R245fa) and 31 C (R134a).

92

S. Szczukiewicz et al. / International Journal of Multiphase Flow 59 (2014) 84101

In the present simulations, R245fa is employed as the working


uid. The microchannel has a circular cross-section with a diameter of D = 0.5 mm and length of 72 diameters, split into an initial
adiabatic region of 16 diameters, then a heated section of 22 diameters and nally a terminal adiabatic zone. Two elongated vapor
bubbles at the saturation temperature T sat 31  C are initialized
at the upstream of the initial adiabatic region of the channel and
the bubbles are 6 diameters apart. The channel is fed with a mass
ux of G 550 kg m2 s1 of saturated liquid, which pushes the
bubbles downstream into the heated region. A constant and uniform heat ux of q 5 kW m2 is applied. The initial conditions
for the temperature and velocity eld of the liquid are obtained
by means of a preliminary liquid-only steady-state simulation. As
the simulation of the two-phase ow starts, the bubbles quickly
achieve a steady ow in the adiabatic region of the channel and
the ow is characterized by a time-averaged cross-sectional void
fraction of 0.28 and a vapor quality of 0.02. As the bubbles enter
into the diabatic section of the channel, they begin to grow and
to accelerate downstream due to the evaporation of the superheated liquid present in the thermal boundary layer at the wall
and at the nose. The dynamics of the bubbles during the evaporation process and the induced wall heat transfer is different between the leading and the following bubble, because the second
bubble comes across a region which had already been cooled down
by the transit of the leading one. Hence, the trailing bubble shows a
lower evaporation rate and velocity, which results in a 13% thinner
liquid lm than the bubble ahead. As an illustration, Fig. 7 plots the
heat transfer coefcient versus time after 21 heated diameters for
the simulation and a model which is going to be discussed below.
The heat transfer coefcient is computed as:

at

q
T w t  T sat

10

where T w is the wall temperature. The numerical results show that


the heat transfer increases during the transit of the bubbles, as it is
expected due to the evaporation of the thin liquid lm trapped between the bubble and the channel wall, then it reaches a maximum
and decreases smoothly as the liquid plugs following each bubble
pass by. The second bubble and its liquid slug have signicantly
higher heat transfer than the rst bubble, not only relevant for
the heat transfer model but also illustrating the limitation of
single-bubble simulations. The local heat transfer behavior in the

liquid slugs cannot be captured by single-phase methods for heat


transfer, such as the Shah and London (1978) correlation implemented in the three-zone model, as it is generated by a specic circulation ow induced by the two-phase ow at the tails of the
bubbles. The ow pattern of the liquid within the slug trapped between the bubbles is obtained by computing the streamlines of the
velocity eld relative to the velocity of the nose of the trailing bubble, which are displayed in Fig. 8 for the time instant t 22:4 ms at
which both the bubbles are growing due to the evaporation.
It is observed that, in agreement with the Thulasidas et al.
(1997) experimental measurements, the ow within the liquid
slug can be split into a wall-adherent liquid layer which bypasses
the bubbles and a recirculating ow which occurs in the core region of the channel.
By assuming that the heat is transferred by one-dimensional
heat conduction from the channel wall to the recirculating ow region, the transient-heat-conduction-based boiling heat transfer
model for the liquid lm region of Magnini et al. (2013a) can be extended to model the heat transfer in the liquid slug region as well,
to thus obtain an improved two-zone (liquid dryout is not modeled
here) boiling heat transfer model for slug ow. A schematical representation of the decomposition of the ow domain adopted by
such two-zone model is depicted in Fig. 8(b). In the liquid slug region, the heat transfer coefcient is modeled by solving the Fourier
equation:

@2T
@T
qcp
@y2
@t

within the wall-adherent lm region bounded by the channel wall


at y 0 and the recirculating ow region at y ds . An analytical
expression for the thickness of the wall-adherent liquid layer ds
was provided by Thulasidas et al. (1997). At the boundary, a constant heat ux condition is applied at y 0, while a convection condition is employed to model the heat transfer at the ctitious
boundary between the wall-adherent and recirculating regions:

k

@T
as T  T sat
@y

k
D

Heat transfer coefficient [W/m2K]

liquid
slug

leading
bubble

trailing
bubble

liquid

3000

2500

2000

1500

Simulation
Model

1000
15

20

25

30

35

40

45

Time [ms]
Fig. 7. Heat transfer coefcient after 21 heated diameters given by the simulation
of two bubbles owing in the microchannel and by the proposed analytical model
for the heat transfer. The black vertical lines identify the transit of the bubbles nose
and rear.

12

where as is the heat transfer coefcient between these regions,


which is presently evaluated by means of the following correlation
originally proposed by He et al. (2010):

as 24:7 0:54Pe0:45 Ls =D1:34


liquid

11

13

where Pe is the Peclet number of the liquid within the slug and Ls is
the length of the liquid slug. Note that, in Eq. (12), the recirculating
ow region is assumed to be at the saturation temperature. The initial temperature prole for the liquid slug region is obtained by the
model itself, as the temperature prole at the end of the previous
liquid lm region. The so-dened ow problem allows an analytical
expression for the heat transfer coefcient in the liquid slug region
to be obtained, which is then coupled with the solution presented
by Magnini et al. (2013a) for the liquid lm zone. The prediction given by this updated model is plotted in red in Fig. 7 and it estimates
the heat transfer trends and magnitude in satisfactory agreement
with the results of the computations. This model can be developed
further to decrease the overestimation observed in Fig. 7 as shown
in Magnini et al. (2013b), and to be made fully stand-alone by
means of correlations available in the literature to estimate those
parameters (e.g. thin liquid lm thickness, wall-adherent liquid
layer thickness, etc.) provided here by the numerical simulation results. Hence, this case study provides a good example of how
numerical two-phase simulations can be used to identify new aspects of the heat transfer process in microchannel slug ow and
provide the input to the development of new theoretical models.

93

(a)

Nondimensional radial
position

S. Szczukiewicz et al. / International Journal of Multiphase Flow 59 (2014) 84101


0.5
0.4
0.3
0.2
0.1
0
24

25

26

27

28

29

30

31

32

33

34

Nondimensional axial position

(b)

liquid slug
region
l T =h s (TTs)
y

liquid film
region

recirculating zone

vapor bubble
T=Tsat

y
s

adherent film

liquid film
z

Fig. 8. (a) Streamlines of the velocity eld relative to the velocity of the nose of the trailing bubble, at t = 22.4 ms. The black lines identify the bubbles proles which are
superimposed to the streamlines plot. (b) Scheme of the decomposition of the ow eld within the microchannel adopted by the proposed two-zone boiling heat transfer
model for slug ow. Since d; ds
R, the radial coordinate is here replaced by the vertical coordinate y.

4.2. Heat transfer coefcient data reduction in multi-microchannels


In order to characterize the heat transfer performance of a multi-microchannel evaporator, the local heat transfer coefcient
needs to be determined. Equation (14) expresses the denition of
the local wall heat transfer coefcient:

aw z

qw z
T w z  T fl z

14

where z denotes the distance from the channel entrance. To begin


with, the local saturation temperature of the refrigerant T fl is
determined by the local saturation pressure, which might signicantly decrease along a microchannel. For instance, Szczukiewicz
(2012) showed that the saturation temperature starting at
37.5 C at the inlet of the channel may fall at the end of the channel anywhere from 28 C predicted by the homogeneous pressure
drop model up to 35 C obtained by Lee and Garimella (2008) prediction method, leading to a difference of 50% in heat transfer
coefcient at a local wall temperature of T w 42  C when applying
Eq. (14) to reduce the data. Thus, the channel pressure drop, Dpch ,
needs to be precisely quantied in order to accurately reduce raw
data to heat transfer coefcients. Nonetheless, due to numerous
technical difculties in multi-microchannel experiments, the pressure drop is commonly measured between the inlet and the outlet
manifolds plenums (referring to the total pressure drop). While
some methods of direct local pressure measurements within microchannels are available in the literature, such as the one of Kohl et al.
(2005), applied in a single microchannel down to 25 lm, in the case
of multi-microchannels the applicability of such a single-channel
method is quite difcult.
When direct experimental measurements are not plausible, the
next best way is to either experimentally measure or compute the
inlet pressure at the beginning of the channels, and then the local
saturation pressure (and temperature) prole is obtainable by
applying a well-established two-phase ow pressure drop prediction method for the channels. For instance, the prediction method
of Cioncolini et al. (2009) was found by Costa-Patry (2011) to be
the best method in his tests when comparing with their experimentally measured channel pressure drops. Such models with

predictive error bands of 2030%, however, may still have a crucial


effect on the determination of the local heat transfer coefcients
(Szczukiewicz, 2012) when applying Eq. (14).
Alternatively, the value of Dpch can be calculated by subtracting
the inlet Dpin;rest and the outlet restriction pressure losses
Dpout;rest from the total experimental pressure drop measured between the two plenums. The value of Dpin;rest (usually single-phase
ow) can be obtained by tting subcooled liquid experimental results to the well known formula of Idelcik (1999) or computed
according to Lee and Garimella (2008), as was done by Szczukiewicz (2012). The two-phase ow outlet restriction pressure losses,
Dpout;rest , can be obtained by employing the method of Costa-Patry
et al. (2011b), knowing the absolute pressure measured in the outlet manifolds plenum and the saturation pressure at the outlet of
the channel. The latter can be obtained by imposing an adiabatic
zone at the outlet (when using local heaters) to get the exit wall
temperature, which matches the local saturation temperature,
and thus indirectly yields the local saturation pressure of the
microchannels in a partially heated test section. These temperatures might be measured by, for instance, thermal diode sensors
(Costa-Patry et al., 2011b) or self-calibrated IR camera (Szczukiewicz, 2012).
To get the local saturation pressures along the microchannels,
one can assume a linear pressure drop over the length of the channel using the values of Dpch and then determine T fl based on the vapor pressure curve. However, this is only appropriate when the
pressure drops are small. Instead of assuming such a linear temperature gradient along the channel, Szczukiewicz (2012) reduced
her heat transfer data using the annular pressure drop prediction
method of Cioncolini et al. (2009) combined with the model of
Lockhart and Martinelli (1949) for vapor qualities below the isolated bubble to coalescing bubble (IB-CB) transition of Ong
(2010) and then applying a ratio of the experimental and the predicted pressure drop values to make it match the experimental
pressure at the exit in each case. This emulated the expected
non-linear variation in the pressure gradient and saturation temperature, accounting also for the accelerational pressure gradient
(all of which would emulate what has to be done in the actual simulation/design of a micro-evaporator). Fig. 9(a) demonstrates an
example of the prorated uid temperatures compared to the linear

94

S. Szczukiewicz et al. / International Journal of Multiphase Flow 59 (2014) 84101

(a)

39

(a)

x 10

3D

linear
prorated

Tfl [ oC]

q ft [W m2]

37

35

1D

33
2

10

z [mm]

(b)

x 10

(b)

3.5
3

10

x 10
3.5

2.5

83%

2
1.5
1

[W m2 K1]
w

[W m2 K1]

z [mm]

2.5
2
1.5
1

linear

1D

homogeneous pch

0.5

0.5

prorated

3D

10

z [mm]
Fig. 9. (a) Prorated uid temperature (Szczukiewicz, 2012), and (b) local heat
transfer coefcient trends assuming linear, homogeneous, and prorated pressure
(and consequently uid saturation temperature) proles along the channel for
R236fa owing in the test section with ein;rest 4, Gch 1692 kg m2 s1 ,
qb 47:8 W cm2 for the experimental channel pressure drop of 46.3 kPa. Note
that in Fig. 9(b), as explained later in the text, the heat transfer coefcients affected
by edge effects are excluded.

temperature prole for determining the local heat transfer coefcients (Szczukiewicz, 2012). Fig. 9(b) highlights the important differences in the local heat transfer coefcients using three different
approaches: (i) assuming linear, (ii) homogeneous, and (iii) prorated pressure (and consequently saturation temperature) proles.
The homogeneous pressure drop model (taken here as an example)
overpredicts the values of the channel pressure drop and thus lowers the value of the local saturation temperature of the refrigerant
at the exit and the value of the local heat transfer coefcient, leading to a difference of 83% between the approaches (ii) and (iii) at
z = 9.5 mm for this test case. Furthermore, the linear pressure drop
assumption articially brings the local heat transfer coefcients to
lower values (except the inlet and the outlet temperatures which
are experimentally measured), which might severely affect the
heat transfer coefcients along the length by up to 10%, when comparing to the approach (iii), i.e. at the CB-AF ow transition. Therefore, the proration method for simulating the uid temperature is
recommended as the most appropriate one in order to provide the
most accurate estimation of the local heat transfer coefcient (future experimental studies should take note of this).
Turning now to another common data reduction procedure, the
conventional 1D heat conduction approach does not take into account the heat spreading towards the colder surrounding regions
that can be observed due to the strong variation in the local heat
transfer coefcient with vapor quality along the channel and at

0
0

0.05

0.1

0.15

0.2

x []
Fig. 10. Two-phase ow of R1234ze(E) in the test section with ein;rest 2 for
Gch 1705 kg m2 s1 and qb 32 W cm2 : (a) actual heat ux at the root of the
ns along the channel length qft , and (b) local wall heat transfer coefcients, aw ,
obtained using the 1D and 3D conduction schemes in function of x. Figure extracted
from Szczukiewicz (2012).

the boundaries. Costa-Patry (2011) noted that the lateral non-uniform heat ux distribution changes the local pressure drop and
evaporation rates. Consequently, the calculated values of the local
wall temperatures and heat uxes and local vapor qualities are
inuenced by the data reduction procedure. The heat spreading effects can be accounted for by using the pragmatic 3D thermal conduction scheme of Costa-Patry (2011), where the temperature and
heat ux values at the test sections base are found by spatially discretizing the 3D domain and then solving an energy balance for
each control volume (CV). Afterwards, assuming the external walls
of the silicon test section to be adiabatic, the various nodes are
linked with each other, such that: Q L n Q R n  1,
Q F n Q B n  1, and Q D n Q U n  1, where Q is the heat
ow rate and n is a natural number indicating the layers number.
The notations above are as follows: D for down (base for the rst
layer of CVs), U up, L left, R right, F front, and B back. This
procedure is quite fast and yields comparable results to a full 3D
heat conduction simulation.
Fig. 10 presents the comparison between the 1D and 3D heat
conduction schemes (Szczukiewicz, 2012), where the silicon wafer
was discretized in 100  100 (set to the pixels of the IR temperature measurements) 140 control volumes (and taking into account the thermal conductivity change with respect to
temperature). The biggest discrepancy is noticeable at the corners,
where the edge effects are most signicant and they were better
captured by the 3D heat conduction model. As can be seen, the rst

S. Szczukiewicz et al. / International Journal of Multiphase Flow 59 (2014) 84101

and the last 5 pixels of the IR array include edge effects, and thus
they are discarded. Moreover, the heat spreads towards the colder
surrounding regions (regions of higher heat transfer coefcients
and slope in T fl ), here at the inlet and the outlet, and the actual heat
ux in those zones is rst higher and then lower than that assumed
in 1D calculations. Therefore, the heat transfer coefcients determined assuming 1D conduction from the base of the silicon micro-evaporator to the root of the ns are underpredicted at the
inlet and the outlet of the channel, as demonstrated in Fig. 10(b),
where the local vapor quality is calculated with corresponding local heat ux prole of the 1D and 3D calculations. If the heat
spreading is not considered, the heat transfer coefcient is overpredicted at the local minimum, which corresponds to a ow transition. The present silicon test section has only a 0.28 mm base
thickness, but still exhibits some heat spreading, which will be
even more signicant for thicker micro-evaporators. Fig. 11 shows
an example of the local wall heat transfer coefcients of R236fa obtained through 1D and 3D heat conduction schemes at a wall heat
ux of qw 71 kW m2 and three different mass uxes, where for
instance for Gch 2099 kg m2 s1 at z = 9.5 mm the heat transfer
coefcient accounting for 3D heat spreading is 14% higher compared to 1D heat conduction scheme. Thus, 3D heat spreading
should be included in all future data reduction procedures in
experimental studies to obtain the most accurate results for the
values of aw and x, and thus the data trends. Since a real microevaporator cold plate is subject to heat spreading effects, it is
important that test data are reduced and then modeled in the same
manner as the actual application.
Fig. 12(a) illustrates the wall heat transfer coefcient, aw , obtained using 3D heat spreading in the silicon micro-evaporator of
Szczukiewicz (2012), plotted versus the longitudinal channel location, z, and compared to the corresponding ow pattern map transition (vertical blue line). The descending trend of aw at the
beginning of the channel (low vapor quality range) corresponds
to the coalescing bubble (CB) region, where the elongated bubbles
coalesce and the local intermittent dry-out patches are formed
(Thome et al., 2004). The local minimum of the heat transfer coefcient is well located by the coalescing bubble annular ow (CB
AF) transition of Costa-Patry and Thome (2012, 2013), computed
according to Eq. (1). Fig. 12(b) shows a video image of the ow
aligned with the graph at the top. The color changes along the
channel from almost black at the entrance (subcooled ow/bubbly
ow) to nearly white (transition region) and becomes gray at the

3.5

x 10

[W m
w

(a)

(b)

Fig. 12. The ashing two-phase ow without backow operating regime of R236fa
in the micro-evaporator with ein;rest 4 for Gch 2096 kg m2 s1 and
qb 47 W cm2 : (a) wall heat transfer coefcient, aw , and (b) video image of ow.
Figure extracted from Szczukiewicz (2012).

exit. The blue line indicating the CB AF transition of Costa-Patry


and Thome (2012, 2013) falls between the second and the third
zone. This means that more than half of the channel is in the annular ow (AF) regime. Due to the convective evaporation across the
annular lm, this provides an increasing heat transfer rate with
increasing vapor quality. Thus, experimental databases should be
coupled with ow visualizations in order to capture the ow pattern effects on the major trends in the data, or at least evaluated
and characterized using the best ow pattern map for the application. In essence, one could consider the experiment to be incomplete if this is not done. . .imagine presenting single-phase ow
data without indicating the transition point from laminar to turbulent ow or the location of fully developed ow.
4.3. Stable and unstable two-phase heat transfer coefcients

2.5

K ]

95

2
1.5

G = 911 kg m2 s1, 1D
ch
2 1
Gch = 911 kg m s , 3D
G = 1514 kg m2 s1, 1D
ch
2 1
Gch = 1514 kg m s , 3D
G = 2099 kg m2 s1, 1D
ch
2 1
Gch = 2099 kg m s , 3D

1
0.5
0
0

0.1

0.2

0.3

0.4

0.5

x []
Fig. 11. Two-phase wall heat transfer coefcient as a function of local vapor quality
for R236fa owing in the test section with the inlet restrictions of ein;rest 4 for
qw 171 kW m2 . The calculations of the local vapor quality here included the 3D
heat spreading effect.

Two-phase ow stability is the next important issue to be considered (but often ignored in the microchannel ow boiling literature), which needs to be addressed in any credible heat transfer
study, since its effect might signicantly change the heat transfer
coefcient trends and values with respect to the vapor quality, as
illustrated by a specic experimental comparison performed by
Consolini and Thome (2009). Figure 13(a) and (b) shows the twophase ow boiling of R245fa and the time-averaged temperature
distribution in the micro-evaporator without any inlet restrictions
(micro-orices). As expected, signicant ow instabilities, vapor
backow, and ow maldistribution occur, which are eventually
governed by the pressure drop in each individual channel, which
lead to high-amplitude and high-frequency temperature and pressure oscillations (see for instance Wu and Cheng (2003)). It is
worthwhile to mention that in such a situation, the assumption
of uniform mass ux in all the channels is invalid and it is not

96

S. Szczukiewicz et al. / International Journal of Multiphase Flow 59 (2014) 84101

(a)

(b) 10

IR

[ C]
54
52

y [mm]

50
6

48
46

44
2
42
40
2

10

z [mm]

(c)

TIR [oC]

(d)

54

10

y [mm]

50

OUTLET

Flow direction

Flow uniformity improved

No back flow

INLET

52
8

48
46

44
2

42
40
2

10

z [mm]

Fig. 13. Snapshots of the high-speed ow visualization and the time-averaged IR temperature maps of the test sections base provided by the two-phase ow boiling of
R245fa for Gch 2035 kg m2 s1 and qb 36:5 W cm2 : (a), (b) without any inlet restrictions, and (c), (d) with the 50 lm-wide, 100 lm-deep, and 100 lm-long inlet microorices (Szczukiewicz et al., 2012b; Szczukiewicz et al., 2013b). The ow is from left to right in all the presented images for both ow and temperature.

advised to be used to reduce the data for local heat transfer coefcients. This illustrates the necessity of developing new measurement techniques for determining the mass ux in each individual
channel within multi-microchannel test sections.
As demonstrated in Fig. 13(c), such undesired phenomena
might be prevented by using an inlet slit to create restrictions at
the entrance of each channel, which tend to stabilize the twophase ow (Agostini et al., 2008c). For the same reason, Park
et al. (2009) used inlet restrictions for each channel in their copper
test section. These restrictions and the channels in their case were
fabricated separately and then aligned within the multiple-microchannel element tested. Nonetheless, such a solution is not applicable for silicon chips, where the restrictions are generally
fabricated along with the channels, applying the same manufacturing process. Recently, Szczukiewicz et al. (2012b, 2013b), used
rectangular orices at the inlet of each channel (they were manufactured in one etching process together with the microchannels)
with expansion ratios varying from ein;rest 1:33 to 4. However,
they saw that the micro-orices of ein;rest 1:33 did not stabilize
the two-phase ow of R236fa within the range of the tested experimental conditions. Such geometries were also previously studied
by Peles and co-workers, i.e. Kosar et al. (2006), Schneider et al.
(2006, 2007), Kosar and Peles (2007).
Fig. 14 illustrates the two-phase ow and temperature patterns
of R236fa for three different micro-orices sizes tested by Szczukiewicz (2012) at the same experimental conditions. The channels
were 100  100 lm2 of cross-sectional area, while the expansion
ratio of the inlet restrictions was varied from 1.33 to 4. As can be
seen, the overall two-phase ow stability improves with increasing
the expansion ratio of the inlet restrictions. In Fig. 14(b), the test
sections base temperature appears to have medium values,

although the two-phase ow is always unstable within the tested


range of the experimental conditions. Whereas, as noticeable in
Fig. 14(d) and (f), the temperature increases with increasing the
expansion ratio from 2 to 4 (experimentally, the outlet temperature was controlled to a xed value). As pointed out by Mukherjee
and Kandlikar (2009), the inlet restriction increases the velocity of
the liquid and consequently the liquid ow rate which by-passes
the bubble through the liquid lm. This generates a hydrodynamic
effect which squeezes the bubble, thus thickening the liquid lm
and reducing the thin-lm evaporation heat transfer rate. In order
to explore the effect of the size of the inlet restriction on the thermal behavior of the micro-evaporator under stable conditions, Table 2 lists the spatio-temporal average temperatures of the test
sections base for ein;rest 2 and 4, for a base heat ux ranging from
25 to 48 W cm2 . It is seen that the difference of the average base
temperature T b;av e between these cases decreases with increasing
heat ux. This may be interpreted as a consequence of the higher
evaporation rate which tends to thin the liquid lm, thus opposing
and then suppressing the hydrodynamic effect of the inlet
restriction.
In order to determine the effect of the expansion ratio on the local heat transfer coefcient, a comparison was performed for
R236fa owing in the micro-evaporators with different inlet
restrictions. Fig. 15 reports that the average level of the heat transfer coefcient decreases when increasing the expansion ratio of the
inlet restriction. The red vertical line corresponds to the predicted
CB AF ow transition of Costa-Patry and Thome (2012, 2013). The
represented points are from the ashing two-phase ow without
backow operating regime (Szczukiewicz et al., 2012b, 2013b).
This regime was found to provide the best ow and temperature
stability. Based on the high-speed ow visualization videos, it

97

S. Szczukiewicz et al. / International Journal of Multiphase Flow 59 (2014) 84101

(a)

TIR [oC]

(b) 10

46
8

y [mm]

44
6
42
4

40

38
36
2

10

z [mm]

(c)

(d)

TIR [ C]
10
46
8

y [mm]

44
6
42
4

40

38
36
2

10

z [mm]

(e)

(f)

TIR [oC]
10
46
8

y [mm]

44
6
42
4

40

38
36
2

10

z [mm]
Fig. 14. Snapshots of the high-speed ow visualization and the time-averaged IR temperature maps of the test sections base provided by the two-phase ow boiling of
R236fa for Gch 1100 kg m2 s1 and qb 36 W cm2 : (a), (b) ein;rest 1:33, (c), (d) ein;rest 2, and (e), (f) ein;rest 4. The ow is from left to right in all the presented images for
both ow and temperature.

Table 2
Spatio-temporal average temperatures of the test sections base for
Gch 2094 kg m2 s1 of channel mass ux, the base heat ux, qb , ranging from 25
to 48 W cm2 and the expansion ratios of the inlet restrictions of ein;rest 2 and 4,
where DT T b;av e ein;rest 4  T b;av e ein;rest 2.
qb
W cm2

T b;av e ein;rest 2
C

T b;av e ein;rest 4
C

DT
K

25.0
29.0
32.6
36.6
40.0
44.4
48.0

42.6
43.5
44.3
45.3
46.1
47.0
47.6

45.2
46.0
46.8
47.3
48.1
48.9
49.5

2.6
2.5
2.5
2.0
2.0
1.9
1.9

was determined that bubbly ow was initiated at the beginning of


each channel from the ashing and almost immediately developed
into slug and consequently annular ow.

In Fig. 15 at low vapor qualities, the descending trend of aw corresponds to the coalescing elongated bubble region, as predicted
by the three-zone model of Thome et al. (2004). In this zone, the
higher expansion ratio leads to a drop of the heat transfer coefcient, as explained previously. After reaching the local minimum,
which corresponds relatively well to the CB AF ow transition
of Costa-Patry and Thome (2012, 2013), the heat transfer increases
with increasing vapor quality. In the annular ow regime, the heat
transfer coefcient for the higher expansion ratio grows more signicantly. This can be ascribed to the instability of the vaporliquid interface triggered by the higher velocities associated with
the smaller restriction that gives a more irregular bubble interface,
thus enhancing the time-averaged heat transfer coefcient (Tibirica et al., 2012). In addition the perturbed interface promotes more
liquid to be entrained in the vapor core, which was shown by
Cioncolini and Thome (2011) to have a positive effect on the heat
transfer performance (of course as long as dryout is avoided).

98

S. Szczukiewicz et al. / International Journal of Multiphase Flow 59 (2014) 84101


4

x 10

[W m2 K1]

2.5
2
1.5
e

in,rest

in,rest

0.5

2 1

Gch = 2099 kg m

s , qb = 16.9 W cm

Gch = 2096 kg m2 s1, qb = 48.0 W cm2

2.5
2
1.5

K ]

Gch = 906 kg m2 s1, qb = 16.8 W cm2

CBAF transition of CostaPatry


and Thome (2012)

w [W m

x 10

3.5

3.5

=2

=4

1
0.5
0

0
0

0.05

0.1

0.15

0.2

0.25

0.05

0.1

0.15

The further analysis includes the impact of two-phase ow


instability on the value and trend of the heat transfer coefcient.
Fig. 16 shows a comparison of three sets of data: two at similar
heat ux (green and red lines) and two at similar mass ux (red
and blue lines). For the lowest heat ux and mass ux (green color), signicant ow instability, vapor backow, and ow maldistribution were present. In the second case (red color), the ow
instabilities were prevented and the vapor backow was suppressed. However, the ow experienced a severe ow maldistribution, which increased and then decreased the heat transfer
coefcients by about 30%, showing the maldistribution is mostly
detrimental to heat transfer. As the mass ux and the heat ux increased (blue color), the stable two-phase ow with a ashed vapor injected into the channel was achieved. The solid lines
correspond to experimental results, whereas the dashed ones represent the simulated (steady) heat transfer coefcients for comparison. The experimental trends of the heat transfer coefcient are Ushaped (over the applicable ranges of vapor qualities) in all three
cases. The stable two-phase ow with ashing (blue line) lead to
the highest rise of the heat transfer coefcient as the vapor quality
increases in the annular ow regime. Furthermore, since in the
simulations the local heat transfer coefcient in the annular ow
regime does not depend on the heat ux, the dashed red and blue
lines overlap. This tendency is usually observed experimentally,
but it is not present here due to the maldistributed ow. Therefore,
a clear distinction between steady, unsteady, well- and maldistributed ows needs to be made when presenting the heat transfer
coefcient trends.
For unstable two-phase ow (green line in Fig. 16), the experimental heat transfer coefcient trend is nearly at over a wide
range of vapor quality (0.02 < x < 0.17 in the present conditions)
and the unstable values are about the same as those predicted
for stable ow in this case. This may confuse the situation when
a limited number of temperature sensors are adopted and the
experimental trends are extrapolated. For instance, Consolini and
Thome (2009) showed that the local heat transfer coefcients under unsteady conditions were independent of local vapor quality.
Nonetheless, they used only 5 sensors along the channel length,
which is much less than the present case, and thus the increase
of the heat transfer coefcient toward the inlet and outlet could
not be captured in their tests.
The results presented above are time-averaged; however, a
temporal analysis is needed to provide a full overview on the multi-microchannel heat transfer performance and reliability during

0.25

0.3

Fig. 16. Two-phase wall heat transfer coefcient of R236fa in the micro-evaporator
with the inlet restrictions of the expansion ratio of ein;rest 4 as a function of local
vapor quality, where solid lines correspond to experimental results, while the
dashed ones represent the simulated (steady) values obtained using the ow
pattern-based heat transfer model of Costa-Patry and Thome (2012, 2013).

its whole lifetime in practical applications. To this aim, the temperature uctuations for a selected pixel of the IR cameras sensor array are assessed in Fig. 17. Note that the highest standard deviation
of the temperature uctuation, 0.82 K, is detected for the unstable
two-phase ow with backow developing into jet ow operating
regime (c), but as shown in Szczukiewicz (2012) it can be much
higher. On the other hand, the ashing two-phase ow without
backow, dened as the most optimum operating regime, provides
very stable ow with a temperature standard deviation of 0.04 K.
Additionally, the current analysis reveals that the frequency of the
temperature oscillations is about 2 to 6 Hz for unstable two-phase
ow (c). The frequency increases with mass and heat uxes (g, h),
simultaneously with decreasing the amplitude of the oscillations,
such that above a certain threshold for heat and mass ux the
oscillations disappear and the ow can be regarded as stable. The

60

55
(h)

(g)

50

Tb [oC]

Fig. 15. Two-phase wall heat transfer coefcient as a function of local vapor quality
for R236fa in the test section with the inlet restrictions of ein;rest 2 and 4 for
Gch 2081 kg m2 s1 and qw 173 kW m2 (stable two-phase ows).

0.2

x []

x []

(b)

45

(a)

(c)

(e)
(d)

(f)

40

35

30
0

10

20

30

40

50

t [s]
Fig. 17. Temporal temperature uctuations at the base of the silicon microevaporator of ein;rest 2: (a) single-phase ow in the entire test section with the
vapor bubbles appearing only at the manifolds outlet plenum, (b) single-phase ow
followed by two-phase ow with backow into the inlet header, (c) unstable twophase ow with backow developing into jet ow, (d) jet ow, (e) single-phase ow
followed by two-phase ow without backow, (f) two-phase ow with backow
triggered by bubbles formed in the ow loop before the test section, (g) ashing
two-phase ow (at exit of inlet micro-orices) with backow, and (h) ashing twophase ow without backow. Figure extracted from the Ph.D. thesis of Szczukiewicz
(2012).

S. Szczukiewicz et al. / International Journal of Multiphase Flow 59 (2014) 84101


4

2.8

x 10

z = 2 mm, ein,rest = 2
z = 2 mm, e
=4
in,rest
z = 5 mm, e
=2
in,rest
z = 5 mm, ein,rest = 4
z = 8 mm, e
=2
in,rest
z = 8 mm, ein,rest = 4

2.4

w [W m2 K ]

2.6

2.2
2
1.8
1.6
1.4
0

10

y [mm]
Fig. 18. Width-wise heat transfer coefcient proles for two-phase ows of R236fa
in ein;rest 2 (unsteady case) and ein;rest 4 (steady case) at the longitudinal channel
location of z 2, 5, and 8 mm for Gch 1292 kg m2 s1 and qb 43:2 W cm2 ,
considering an array of 90  90 IR temperatures.

frequency of the present temperature oscillations for unstable


ows agrees quite well with the ones given by Consolini and
Thome (2009).
As a last point, the width-wise heat transfer coefcient proles
for steady (in ein;rest 4) and unsteady (in ein;rest 2) two-phase
ows of R236fa at the longitudinal channel location of z = 2, 5,
and 8 mm for Gch 1292 kg m2 s1 and qb 43:2 W cm2 are presented in Fig. 18. The rst two locations correspond to the churn
ow, while the last one represents the annular ow. It is evident
that average level of the heat transfer coefcient decreases with
increasing the expansion ratio of the inlet restriction, which is
associated with higher ow velocities in ein;rest 4. Those ones extend the region of churn ow. Whereas, in annular ow for
z = 8 mm, the difference in heat transfer coefcient between
ein;rest 2 and 4 becomes less signicant. Secondly, the heat transfer coefcient for the unsteady case is characterized by signicant
uctuations in the lateral direction. Whereas, in ein;rest 4, where
the ow is very stable in time (see the ashing two-phase ow
without backow operating regime in Fig. 17), the heat transfer
coefcient prole is more uniform between y = 2 mm and 8 mm,
with y being a coordinate perpendicular to the ow direction. In
this test section, due to the heat spreading in the lateral direction,
the local heat transfer coefcient at the boarders rises proportionally to the dissipated heat ux, since the heat tends to conduct towards the cooler regions (for more detail, refer to Section 4.2). This
highlights again that 3D heat conduction analysis needs to be taken into account to provide the most accurate heat transfer results
and data trends.
5. Conclusions
This paper illustrates the most recent experimental and numerical outcomes on two-phase ow boiling in microchannels. The
numerical simulations of boiling ows performed by the authors
show that the proper modeling of the thermal inertia of the liquid
lm trapped between an elongated bubble and the channel wall,
and of the ow recirculation in the liquid slug between two bubbles, provides very valuable local information on the heat transfer
coefcient. This may aid to improve physics-based boiling heat
transfer prediction methods, for instance the broadly-used threezone model of Thome et al. (2004). A better understanding of the
linkage between ow patterns and heat transfer trends in microchannels is also obtained by coupling the very ne spatial

99

temperature measurements allowed by IR thermography with ow


visualization techniques. Such analysis conrmed that ow pattern
transitions need to be further investigated to clarify the observed
trends in the heat transfer coefcient with respect to the local vapor quality. High emphasis is here given to the heat transfer coefcient data reduction process which, in order to conduct a fair
comparison among independent experimental studies, has to account for peculiar microscale effects associated with the miniaturization of the tested evaporators. In particular, when the pressure
drop is signicant, the saturation temperature involved in the heat
transfer coefcient estimation cannot be assumed to be constant or
linear along the channel length; it was illustrated how linear, prediction method-based, and prorated reconstructions of the pressure prole along the microchannel may lead to remarkable
differences in heat transfer trends and magnitudes. Moreover, the
measured wall temperature and the actual local heat ux seen
by the two-phase ow are affected by the heat spreading across
the micro-evaporator due to the slopes in the heat transfer coefcients themselves and by the edge effects in proximity of its manifold. Therefore, an accurate 3D modeling of the heat conduction
from the base of the heat sink to the channels walls is recommended, in order to obtain a heat transfer coefcient which properly reects the performance of the microchannel two-phase ow.
In addition, the spatio-temporal local heat transfer coefcient analysis revealed that the two-phase ow instability and ow maldistribution among the channels may have predominant inuence on
the experimental results and data trends, and thus add to confuse
of the situation. Therefore, a clear distinction between mal- and
well-distributed ows, unsteady, and steady ows needs to be
made in order to properly reduce and compare the results from
independent studies. Finally, it was shown that the inlet micro-orices are an effective solution to stabilize the two-phase ow in
microchannels, and thus greatly extend the reliable range of operating conditions of the micro-evaporator.

References
Afkhami, S., Bussmann, M., 2008. Height functions for applying contact angles to 2D
VOF simulations. Int. J. Numerical Methods Fluids 57, 453472.
Agostini, B., Bontemps, A., 2005. Vertical ow boiling of refrigerant R134a in small
channels. Int. J. Heat Fluid Flow 26, 296306.
Agostini, B., Bontemps, A., Thonon, B., 2006. Effects of geometrical and
thermophysical parameters on heat transfer measurements in small-diameter
channels. Heat Transfer Eng. 27, 1424.
Agostini, B., Bontemps, A., Watel, B., Thonon, B., 2003. Boiling heat transfer in
minichannels: inuence of the hydraulic diameter. In: Int. Congress of
Refrigeration.
Agostini, B., Fabbri, M., Park, J.E., Wojtan, L., Thome, J.R., Michel, B., 2007. State of the
art of high heat ux cooling technologies. Heat Transfer Eng. 28, 258281.
Agostini, B., Revellin, R., Thome, J.R., 2008a. Elongated bubbles in microchannels.
Part I: Experimental study and modeling of elongated bubble velocity. Int. J.
Multiphase Flow 34, 590601.
Agostini, B., Thome, J.R. 2005. Comparison of an extended database for boiling heat
transfer coefcients in multi-microchannels elements with the three-zone
model. In: ECI Heat Transfer and Fluid Flow in Microscale. Castelvecchio Pascoli,
Italy.
Agostini, B., Thome, J.R., Fabbri, M., Michel, B., 2008b. High heat ux two-phase
cooling in silicon multimicrochannels. IEEE Trans. Compon. Packag. Technol. 31,
691701.
Agostini, B., Thome, J.R., Fabbri, M., Michel, B., Calmi, D., Kloter, U., 2008c. High heat
ux ow boiling in silicon multi-microchannels Part I: heat transfer
characteristics of refrigerant R236fa. Int. J. Heat Mass Transfer 51, 54005414.
Agostini, B., Thome, J.R., Fabbri, M., Michel, B., Calmi, D., Kloter, U., 2008d. High heat
ux ow boiling in silicon multi-microchannels Part II: heat transfer
characteristics of refrigerant R245fa. Int. J. Heat Mass Transfer 51, 54155425.
Baldassari, C., Marengo, M., 2013. Flow boiling in microchannels and microgravity.
Prog. Energy Combustion Sci. 39, 136.
Barber, J., Brutin, D., Seane, K., Gardarein, J.L., Tadrist, L., 2011. Unsteady-state
uctuation analysis during bubble growth in a rectangular microchannel. Int. J.
Heat Mass Transfer 54, 47844795.
Barber, J., Seane, K., Brutin, D., Tadrist, L., 2009. Hydrodynamics and heat transfer
during ow boiling instabilities in a single microchannel. Appl. Thermal Eng. 29,
12991308.

100

S. Szczukiewicz et al. / International Journal of Multiphase Flow 59 (2014) 84101

Bertsch, S.S., Groll, E.A., Garimella, S.V., 2009. A composite heat transfer correlation
for saturated ow boiling in small channels. Int. J. Heat Mass Transfer 52, 2110
2118.
Borhani, N., Agostini, B., Thome, J.R., 2010. A novel time strip ow
visualisation technique for investigation of intermittent dewetting and
dryout in elongated bubble ow in a microchannel evaporator. Int. J.
Heat Mass Transfer 53, 48094818.
Brackbill, J.U., Kothe, D.B., Zemach, C., 1992. A continuum method for modeling
surface tension. J. Comput. Phys. 100, 335354.
Chen, J.C., 1966. Correlation for boiling heat transfer to saturated uids in
convective ow. Ind. Eng. Chem. Process Des. Dev. 5, 322329.
Cheng, L., Ribatski, G., Thome, J.R., 2008. Two-phase ow patterns and ow-pattern
maps: fundamentals and applications. Appl. Mech. Rev. 61, 128.
Cioncolini, A., Thome, J.R., 2011. Algebraic turbulence modeling in adiabatic and
evaporating annular two-phase ow. Int. J. Heat Fluid Flow 32, 805817.
Cioncolini, A., Thome, J.R., Lombardi, C., 2009. Unied macro-to-microscale
method to predict two-phase frictional pressure drops of annular ows.
Int. J. Multiphase Flow 35, 11381148.
Consolini, L., Thome, J.R., 2009. Micro-channel ow boiling heat transfer of R-134a,
R-236fa, and R-245fa. Microuids Nanouids 6, 731746.
Consolini, L., Thome, J.R., 2010. A heat transfer model for evaporation of coalescing
bubbles in microchannel ow. Int. J. Heat Fluid Flow 31, 115125.
Costa-Patry, E. 2011. Cooling High Heat Flux Micro-Electronic Systems Using
Refrigerants in High Aspect Ratio Multi-Microchannel Evaporators, Ph.D. thesis.
Ecole Polytechnique Federale de Lausanne.
Costa-Patry, E., Nebuloni, S., Olivier, J., Thome, J.R., 2012a. On-chip two-phase
cooling with refrigerant 85 lm-wide multi-microchannel evaporator under
hot-spot conditions. IEEE Trans. Compon. Packag. Manuf. Technol. 2, 311320.
Costa-Patry, E., Olivier, J., Michel, B., Thome, J.R., 2011a. Two-phase ow of
refrigerant in 85 lm-wide multi-microchannels: Part II heat transfer with 35
local heaters. Int. J. Heat Fluid Flow 32, 464476.
Costa-Patry, E., Olivier, J., Nichita, B.A., Michel, B., Thome, J.R., 2011b. Two-phase
ow of refrigerant in 85 lm-wide multi-microchannels: Part I pressure drop.
Int. J. Heat Fluid Flow 32, 451463.
Costa-Patry, E., Olivier, J., Thome, J.R., 2012b. Heat transfer characteristics in
a copper micro-evaporator and ow pattern-based prediction method for
ow boiling in microchannels. Frontiers Heat Mass Transfer 3, 114.
Costa-Patry, E., Thome, J.R. 2012. Flow pattern based ow boiling heat transfer
model for microchannels. In: ECI 8th International Conference on Boiling and
Condensation Heat Transfer.
Costa-Patry, E., Thome, J.R., 2013. Flow pattern based ow boiling heat transfer
model for microchannels. Int. J. Refrigeration 36, 414420.
Cummins, S.J., Francois, M.M., Kothe, D.B., 2005. Estimating curvature from volume
fractions. Comput. Struct. 83, 425434.
Diaz, M.C., Boye, H., Hapke, I., Schmidt, J., Staate, Y., Zhekov, Z., 2005.
Investigation of ow boiling in narrow channels by thermographic
measurements of local wall temperatures. Microuidics Nanouidics 1,
111.
Diaz, M.C., Schmidt, J., 2007a. Experimental investigation of transient boiling heat
transfer in microchannels. Int. J. Heat Fluid Flow 28, 95102.
Diaz, M.C., Schmidt, J., 2007b. Flow boiling of n-hexane in small channels heat
transfer measurements and ow pattern observations. Chem. Eng. Technol. 3,
389394.
Dong, Z., Xu, J., Jiang, F., Liu, P., 2012. Numerical study of vapor bubble effect on ow
and heat transfer in microchannel. Int. J. Thermal Sci. 54, 2232.
Dupont, V., Thome, J.R., 2005. Evaporation in microchannels: inuence of the
channel diameter on heat transfer. Microuidics Nanouidics 1, 119127.
Han, Y., Shikazono, N., 2009. Measurement of the liquid lm thickness in microtube
slug ow. Int. J. Heat Mass Transfer 30, 842853.
Hapke, I., Boye, H., Schmidt, J., 2002. Flow boiling of water and n-heptane in micro
channels. Microscale Thermophys. Eng. 6, 99115.
Hardt, S., Schilder, B., Tiemann, D., Kolb, G., Hessel, V., Stephan, P., 2007. Analysis of
ow patterns emerging during evaporation in parallel microchannels. Int. J.
Heat Mass Transfer 50, 226239.
Harirchian, T., Garimella, S.V., 2008. Microchannel size effects on local ow boiling
heat transfer to a dielectric uid. Int. J. Heat Mass Transfer 51, 37243735.
Harirchian, T., Garimella, S.V., 2009a. Effects of channel dimension, heat ux, and
mass ux on ow boiling regimes in microchannels. Int. J. Multiphase Flow 35,
349362.
Harirchian, T., Garimella, S.V., 2009b. The critical role of channel cross-sectional
area in microchannel. Int. J. Multiphase Flow 35, 904913.
Harirchian, T., Garimella, S.V., 2010. A comprehensive ow regime map for
microchannel ow boiling with quantitative transition criteria. Int. J. Heat
Mass Transfer 53, 26942702.
Harirchian, T., Garimella, S.V., 2012. Flow regime-based modeling of heat transfer
and pressure drop in microchannel ow boiling. Int. J. Heat Mass Transfer 55,
12461260.
He, Q., Hasegawa, Y., Kasagi, N., 2010. Heat transfer modelling of gas-liquid slug
ow without phase change in a micro tube. Int. J. Heat Fluid Flow 31, 126136.
Hetsroni, G., Mosyak, A., Pogrebnyak, E., Segal, Z., 2005. Explosive boiling of water in
parallel micro-channels. Int. J. Multiphase Flow 31, 371392.
Hetsroni, G., Mosyak, A., Pogrebnyak, E., Segal, Z., 2006. Periodic boiling in parallel
micro-channels at low vapor quality. Int. J. Multiphase Flow 32, 11411159.
Hetsroni, G., Mosyak, A., Segal, Z., 2001. Nonuniform temperature distribution in
electronic devices cooled by ow in parallel microchannels. IEEE Trans.
Compon. Packag. Technol. 24, 1623.

Hetsroni, G., Mosyak, A., Segal, Z., Pogrebnyak, E., 2003. Two-phase ow patterns in
parallel micro-channels. Int. J. Multiphase Flow 29, 341360.
Hetsroni, G., Mosyak, A., Segal, Z., Ziskind, G., 2002. A uniform temperature heat sink
for cooling of electronic devices. Int. J. Heat Mass Transfer 45, 32753286.
Hirt, C.W., Nichols, B.D., 1981. Volume of uid (VOF) method for the dynamics of
free boundaries. J. Comput. Phys. 39, 201225.
Idelcik, I.E. 1999. Memento des pertes de charge, Eyrolles, Paris.
Karayiannis, T.G., Shiferaw, D., Kenning, D.B.R., Wadekar, V.V., 2010. Flow patterns
and heat transfer for ow boiling in small to micro diameter tubes. Heat
Transfer Eng. 31, 257275.
Kohl, M.J., Abdel-Khalik, S.I., Jeter, S.M., Sadowski, D.L., 2005. An experimental
investigation of microchannel ow with internal pressure measurements. Int. J.
Heat Mass Transfer 48, 15181533.
Kosar, A., Kuo, C.J., Peles, Y., 2006. Suppression of boiling ow oscillations in parallel
microchannels by inlet restrictors. J. Heat Transfer 128, 251260.
Kosar, A., Peles, Y., 2007. Critical heat ux of R-123 in silicon-based microchannels.
Trans. ASME 129, 844851.
Kunkelmann, C., Stephan, P., 2009. CFD simulation of boiling ows using the
Volume-of-Fluid method within Open FOAM. Numer. Heat Transfer, Part A 56,
631646.
Lee, P.S., Garimella, S.V., 2008. Saturated ow boiling heat transfer and pressure
drop in silicon microchannel array. Int. J. Heat Mass Transfer 51, 789806.
Lee, W., Son, G., Yoon, H.Y., 2012. Direct numerical simulation of ow boiling in a
nned microchannel. Int. Commun. Heat Mass Transfer.
Li, D., Dhir, V.K., 2007. Numerical study of single bubble dynamics during ow
boiling. J. Heat Transfer 129, 864876.
Lockhart, R.W., Martinelli, R.C., 1949. Proposed correlation of data for isothermal
two-phase, two-component ow in pipes. Chem. Eng. Prog. 45, 3948.
Magnini, M., Pulvirenti, B., Thome, J.R., 2013a. Numerical investigation of
hydrodynamics and heat transfer of elongated bubbles during ow boiling in
a microchannel. Int. J. Heat Mass Transfer 59, 451471.
Magnini, M., Pulvirenti, B., Thome, J.R., 2013b. Numerical investigation of the
inuence of leading and sequential bubbles on slug ow boiling within a
microchannel. Int. J. Thermal Sci. 71, 3652.
Mishima, K., Hibiki, T., 1996. Some characteristics of air-water two-phase ow in
small diameter vertical tubes. Int. J. Multiphase Flow 22.
Moore, G.E., 1965. Cramming more components onto integrated circuits. Electronics
38, 114117.
Mukherjee, A., 2009. Contribution of thin-lm evaporation during ow boiling
inside microchannels. Int. J. Thermal Sci. 48, 20252035.
Mukherjee, A., Kandlikar, S.G., 2005. Numerical simulation of growth of a vapor
bubble during ow boiling of water in microchannel. Microuids Nanouids 1,
137145.
Mukherjee, A., Kandlikar, S.G., 2009. The effect of inlet constriction on bubble
growth during ow boiling in microchannels. Int. J. Heat Mass Transfer 53,
52045212.
Mukherjee, A., Kandlikar, S.G., Edel, Z.J., 2011. Numerical study of bubble growth
and wall heat transfer during ow boiling in a microchannel. Int. J. Heat Mass
Transfer 54, 37023718.
Nebuloni, S., Thome, J.R., 2010. Numerical modeling of laminar annular lm
condensation for different channel shapes. Int. J. Heat Mass Transfer 53, 2615
2627.
C.L. Ong, Macro-to-Microchannel Transition in Two-Phase Flow and Evaporation,
Ph.D. thesis, Ecole Polytechnique Fedeerale de Lausanne, Switzerland, 2010.
Ong, C.L., Thome, J.R., 2011a. Macro-to-microchannel transition in two-phase ow:
Part 1 two-phase ow patterns and lm thickness measurements. Exp.
Thermal Fluid Sci. 35, 3747.
Ong, C.L., Thome, J.R., 2011b. Macro-to-microchannel transition in two-phase ow:
Part 2 Flow boiling heat transfer and critical heat ux. Exp. Thermal Fluid Sci.
35, 873886.
Park, J.E., Thome, J.R., Michel, B. 2009. Effect of inlet orices on saturated CHF and
ow visualisation in multi-microchannel heat sinks. In: 25th IEEE
Semiconductor Thermal Measurement and Management (SEMI-THERM)
Symposium.
Patil, V.A., Narayanan, V. 2005. Measurement of near wall liquid temperatures in
single-phase ow through silicon microchannels. In: Proceedings of the 3rd
International Conference on Microchannels and Minichannels.
Patil, V.A., Narayanan, V., 2006. Spatially resolved temperature measurement in
microchannels. Microuidics Nanouidics 2, 291300.
Renardy, M., Renardy, Y., Li, J., 2001. Numerical simulation of moving contact line
problems using a volume-of-uid method. J. Comput. Phys. 171, 243263.
Renardy, Y., Renardy, M., 2002. PROST: a Parabolic Reconstruction of Surface
Tension for the volume-of-uid method. J. Comput. Phys. 183, 400421.
Revellin, R. 2005. Experimental Two-Phase Fluid Flow in Microchannels, Ph.D.
thesis. Ecole Polytechnique Federale de Lausanne.
Revellin, R., Agostini, B., Thome, J.R., 2008. Elongated bubbles in microchannels. Part
II: experimental study and modeling of bubble collisions. Int. J. Multiphase Flow
34, 602613.
Schneider, B., Kosar, A., Kuo, C.J., Mishra, C., Cole, G.S., Scaringe, R.P., Peles, Y., 2006.
Cavitation enhanced heat transfer in microchannels. J. Heat Transfer 128, 1293
1301.
Schneider, B., Kosar, A., Peles, Y., 2007. Hydrodynamic cavitation and boiling in
refrigerant (R-123) ow inside microchannels. Int. J. Heat Mass Transfer 50,
28382854.
Schrage, R.W., 1953. A Theoretical Study of Interphase Mass Transfer. Columbia
University Press, New York.

S. Szczukiewicz et al. / International Journal of Multiphase Flow 59 (2014) 84101


Shah, R.K., London, A.L., 1978. Laminar Flow Forced Convection in Ducts. Academic,
New York.
Stephan, P.C., Busse, C.A., 1992. Analysis of the heat transfer coefcient of grooved
heat pipe evaporator walls. Int. J. Heat Mass Transfer 35, 383391.
Su, S., Huang, S., Wang, X., 2005. Study of boiling incipience and heat transfer
enhancement in forced ow through narrow channels. Int. J. Multiphase Flow
31, 253260.
Suh, Y., Lee, W., Son, G., 2008. Bubble dynamics, ow, and heat transfer during ow
boiling in parallel microchannels. Numer. Heat Transfer, Part A 54, 390405.
Sussman, M., Puckett, E.G., 2000. A coupled level set and volume-of-uid method
for computing 3D and axisymmetric incompressible two-phase ows. J.
Comput. Phys. 162, 301337.
Sussman, M., Smereka, P., Osher, S., 1994. A level set approach for computing
solutions to incompressible two-phase ow. J. Comput. Phys. 114, 146159.
Szczukiewicz, S. 2012. Thermal and Visual Operational Characteristics of MultiMicrochannel Evaporators Using Refrigerants, Ph.D. thesis. Ecole Polytechnique
Federale de Lausanne, Switzerland.
Szczukiewicz, S., Borhani, N., Thome, J.R. 2012a. Two-phase ow boiling in a single
layer of future high-performance 3D stacked computer chips. In: 13th IEEE
Intersociety Conference on Thermal and Thermomechanical Phenomena in
Electronic Systems. San Diego, California, USA.
Szczukiewicz, S., Borhani, N., Thome, J.R. 2012b. Two-phase heat transfer and highspeed visualization of refrigerant ows in 100  100 lm2 silicon multimicrochannels. In: ECI 8th International Conference on Boiling and
Condensation Heat Transfer, Lausanne, Switzerland.
Szczukiewicz, S., Borhani, N., Thome, J.R., 2013a. Two-phase ow operational maps
for multi-microchannel evaporators. Int. J. Heat Fluid Flow 42, 176189.
Szczukiewicz, S., Borhani, N., Thome, J.R., 2013b. Two-phase heat transfer and highspeed visualization of refrigerant ows in 100  100 lm2 silicon multimicrochannels. Int. J. Refrigeration 36, 402413.
Thome, J.R., 2004. Boiling in microchannels: a review of experiment and theory. Int.
J. Heat Fluid Flow 25, 128139.
Thome, J.R., 2006. State-of-the-art overview of boiling and two-phase ows in
microchannels. Heat Transfer Eng. 27, 419.

101

Thome, J.R. 2010. Wolverine Engineering Data Book III. <http://www.wlv.com>.


Thome, J.R., Consolini, L., 2010. Mechanisms of boiling in micro-channels: critical
assessement. Heat Transfer Eng. 31, 288297.
Thome, J.R., Dupont, V., Jabobi, A.M., 2004. Heat transfer model for evaporation in
microchannels. Part I: presentation of the model. Int. J. Heat Mass Transfer 47,
33753385.
Thulasidas, T.C., Abraham, M.A., Cerro, R.L., 1997. Flow patterns in liquid slugs
during bubble-train ow inside capillaries. Chem. Eng. Sci. 52, 29472962.
Tibirica, C.B., Ribatski, G., Thome, J.R., 2012. Flow boiling characteristics for
R1234ze(E) in 1.0 and 2.2 mm circular channels. J. Heat Transfer 134, 18.
Tran, T.N., Wambsganss, M.W., France, D.M., 1996. Small circular- and rectangularchannel boiling with two refrigerant. Int. J. Multiphase Flow 22, 485498.
Vakili-Farahani, F., Agostini, B., Thome, J.R. 2012. Experimental study on ow
boiling heat transfer of multiport tubes with R245fa and R1234ze(E). In: ECI 8th
International Conference on Boiling and Condensation Heat Transfer.
Vakili-Farahani, F., Agostini, B., Thome, J.R., 2013. Experimental study on ow
boiling heat transfer of multiport tubes with R245fa and R1234ze(E). Int. J.
Refrigeration 36, 335352.
Wang, H., Garimella, S.V., Murthy, J.Y., 2007. Characteristics of evaporating thin lm
in a microchannel. Int. J. Heat Mass Transfer 50, 39333942.
Wu, H.Y., Cheng, P., 2003. Visualization and measurement of periodic boiling in
silicon microchannels. Int. J. Heat Mass Transfer 46, 26032614.
Xu, J., Shen, S., Gan, Y., Li, Y., Zhang, W., Su, Q., 2005. Transient ow pattern based
microscale boiling heat transfer mechanisms. J. Micromech. Microeng. 15,
13441361.
Xu, J.L., Zhang, W., Wang, Q.W., Su, Q., 2006. Flow instability and transient ow
patterns inside intercrossed silicon microchannel array in a micro-timescale.
Int. J. Multiphase Flow 32, 568592.
Yu, W., Wambsganss, D.M., Hull, J.R., 2002. Two-phase pressure drop, boiling heat
transfer, and critical heat ux to water in a small-diameter horizontal tube. Int.
J. Multiphase Flow 28, 927941.

S-ar putea să vă placă și