You are on page 1of 8

Experimental Thermal and Fluid Science 47 (2013) 6067

Contents lists available at SciVerse ScienceDirect

Experimental Thermal and Fluid Science


journal homepage: www.elsevier.com/locate/etfs

Heat transfer and pressure drop during condensation of R152a in circular and square
microchannels
Na Liu a, Jun Ming Li a,, Jie Sun b, Hua Sheng Wang b
a
b

Key Laboratory for Thermal Science and Power Engineering of Ministry of Education, Department of Thermal Engineering, Tsinghua University, Beijing 100084, China
School of Engineering and Materials Science, Queen Mary University of London, Mile End Road, London E1 4NS, UK

a r t i c l e

i n f o

Article history:
Received 15 August 2012
Received in revised form 2 November 2012
Accepted 3 January 2013
Available online 16 January 2013
Keywords:
Condensation
Heat transfer
Microchannel
R152a
Pressure drop

a b s t r a c t
The paper reports experimental data for heat transfer and pressure drop during condensation of R152a in
circular and square microchannels with hydraulic diameters of 1.152 mm and 0.952 mm, respectively.
Saturation temperatures are 40 C and 50 C with mass uxes varying from 200 to 800 kg/m2 s and vapor
mass qualities from 0.1 to 0.9. Effects of mass ux, vapor mass quality and channel geometry on heat
transfer and pressure drop were investigated. The results show that heat-transfer coefcients and pressure drop both increase with increasing mass ux and vapor mass quality while decrease with increasing
saturation temperature. Channel geometry has much effect on heat transfer at low mass uxes while has
little effect on pressure drop. The present data were compared with earlier empirical correlations and a
theoretical solution. Heat transfer coefcients agree within experimental error with several correlations
and a theoretical solution for both circular and square microchannels. One pressure drop correlation
underestimates the data for the two microchannels while another pressure drop correlation overestimates the data for the square microchannel.
2013 Elsevier Inc. All rights reserved.

1. Introduction
Microchannels are increasingly used to improve heat-transfer
performance and to enable compact geometries in many applications. For condensation, owing to surface tension effects, methods
used to treat larger channels are not suitable when channel dimension is around 1 mm or less.
Recently, increasing attention on environmental issues has
brought substantial requirements and changes in refrigerants.
Though HFCs substitutes have no ozone depletion potential
(ODP), many of them have relatively high global warming potential
(GWP). Moreover, the European Unions F-gas regulations specify
beginning on January 1, 2011 new models and on January 1,
2017 new vehicles tted with air conditioning cannot be manufactured with uorinated greenhouse gases having GWP greater than
150 [1,2].
Su et al. [3] reviewed earlier experimental work during condensation in microchannels. In many of the earlier heat-transfer measurements, most uids are with similar properties, predominantly
R134a. Vapor-side heat-transfer coefcients have generally been
inferred from overall measurements by subtraction of thermal
resistance or using Wilson plot techniques [4]. Such data have high
uncertainty.
Corresponding author. Tel.: +86 10 62771001; fax: +86 10 62770209.
E-mail address: lijm@tsinghua.edu.cn (J.M. Li).
0894-1777/$ - see front matter 2013 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.expthermusci.2013.01.002

Recently, heat-transfer coefcients with various refrigerants


were obtained from direct measurement of wall temperatures during condensation in circular and non-circular microchannels. Matkovic et al. [58] measured heat-transfer coefcients and pressure
drop of R134a, R32, R1234yf and R245fa in a circular 0.96 mm
microchannel. Del Col et al. [9] measured heat-transfer coefcients
of R134a in a 1.18 mm side length square microchannel. Heattransfer coefcients of R134a were obtained by Derby et al. [10]
during condensation in 1 mm square, triangular and semicircular
multiple parallel minichannels. Mass ux and quality were determined to have signicant effects on the condensation process, even
at lower mass uxes, while saturation pressure, heat ux, and
channel shape had no signicant effects. Kim and Mudawar [11]
proposed a theoretical control-volume-based model based on the
assumptions of smooth interface between the annular liquid lm
and vapor core, and uniform liquid lm thickness around the channels circumference. The new model accurately captures the pressure drop and heat transfer coefcients in both magnitude and
trend. Kim et al. [12,13] experimentally investigated ow regimes,
pressure drop and heat transfer of FC-72 during condensation
along parallel, square micro-channels with a hydraulic diameter
of 1 mm. In their study, smooth-annular, wavy-annular, transition,
slug, and bubbly ow regimes were identied. A detailed pressure
model was presented and assessment of pressure drop correlations
was conducted. Comparing the data to predictions of previous
annular condensation heat transfer correlations showed correla-

61

N. Liu et al. / Experimental Thermal and Fluid Science 47 (2013) 6067

Nomenclature
a
a.m.
A
b
Cc
d
dh
G
f
h
hlv
I
k
l
m
N
Nu
U
p
Q
r.m.s.
Ra
Re
T

data referring to a or dp/dl, see Eqs. (12) and (13)


arithmetic mean error
cross sectional area of channel (m2)
side length of square channel (m)
coefcient of contraction, see Eq. (8)
diameter of circular channel (m)
hydraulic diameter (m)
mass ux (kg/m2 s)
friction factor
specic enthalpy (J/kg)
specic enthalpy of evaporation (J/kg)
electric current (A)
thermal conductivity (W/m K)
length (m)
mass ow rate (kg/s)
number of data points
Nusselt number, adh/kl
electric voltage (V)
pressure (Pa)
heat transfer rate (W)
root-mean-square error
arithmetical mean deviation of the assessed prole
(lm)
Reynolds number, Gdh/ll
temperature (K)

tions intended for macro-channels generally provided better predictions than correlations intended for mini/micro-channels. Besides, a new condensation heat transfer correlation was proposed
for annular condensation heat transfer in mini/micro-channels.
Agarwal et al. [14] used thermal amplication technique to measure heat-transfer coefcients of R134a in six non-circular horizontal microchannels.
Saturation and environmental properties of R22, R134a and
R152a are listed in Table 1. R152a and R134a have zero ozone
depletion potentials. Global warming potential of R152a is 120,
which is an order of magnitude lower than those of R22 and
R134a. It is noted that thermal conductivity, latent heat and surface tension are higher for R152a than for R22 and R134a. Wang
et al. [15] theoretically studied heat-transfer coefcients of R22,
R134a and R152a during condensation. According to their results,
heat-transfer coefcients of R152a are higher than those of R22
and R134a. Therefore, R152a is a potential substitute for R22. However, to the authors knowledge, there are no experimental studies
on heat transfer and pressure drop during condensation of R152a
in microchannels.
In the present paper, heat transfer and pressure drop are
investigated experimentally during condensation of R152a in a
circular and a square microchannels. Experiments are performed
at saturation temperatures of 40 C and 50 C with mass uxes
from 200 to 800 kg/m2 s and vapor mass qualities from 0.1 to
0.9. Effects of mass ux, vapor mass quality and channel geometry on heat transfer and pressure drop are presented in the
paper.

Greek symbols
heat-transfer coefcient (W/m2 K)
area ratio, Atest-section/Aheader
d
channel thickness (m)
l
viscosity (Pa s)
n
void fraction
q
density (kg/m3)
v
vapor mass quality

a
c

Subscripts
de
deceleration
exp
experimental
f
frictional
i
inside
in
inlet
l
liquid
o
outside
out
outlet
pre
predicted
r
refrigerant
s
saturation
v
vapor
w
wall

2. Apparatus and procedure


2.1. Test rig
Fig. 1a shows the schematic of the experimental rig designed for
heat transfer and pressure drop measurements. The rig consists of
one refrigerant loop and two cooling water loops. The sub-cooled
refrigerant from the reservoir is pumped to the Coriolis-effect mass
ow meter by the magnetic-driven gear pump after ltered. The
mass ux is controlled by a bypass loop. In the evaporator, the
refrigerant is heated to a desired vapor mass quality by adjusting
electric heating power. The two-phase refrigerant enters the test
section which is a counter-ow tube-in-tube heat exchanger. The
uid is partly condensed in the test section by the cooling water
outside. After the test section, the uid is sent to the post-condenser, where it is condensed and sub-cooled and nally ows
back to the reservoir. A 500 W electric cartridge heater is xed at
the bottom of the refrigerant reservoir to compensate for system
energy losses and make the system pressure stable. Two thermostatic water baths are used in the rig. One provides the cooling
water for the test section and the other for the post condenser.
The electric cartridge heater at the bottom of the reservoir was
used to change the system pressure to the test saturation pressure.
Meanwhile, the two thermostatic water baths were started to provide constant temperature water for the test section and the postcondenser. The test mass ux was initially set higher than the test
value because pressure drop in the test section increased with the
vaporization of the uid causing a decrease of the mass ux. Differ-

Table 1
Properties of R22, R134a and R152a.

R22
R134a
R152a

ts (C)

ps (MPa)

ql (kg/m3)

qv (kg/m3)

hlv (kJ/kg)

ll (lPa s)

kl (W/m K)

r (N/m)

ODP

GWP

50
50
50

1.943
1.318
1.177

1082
1102.3
830.8

86.0
66.3
37.1

154.2
151.8
245.4

123.0
141.8
122.2

0.0719
0.0704
0.0875

0.00474
0.00489
0.00648

0.05
0
0

1700
1300
120

62

N. Liu et al. / Experimental Thermal and Fluid Science 47 (2013) 6067

(a)

(b)
Fig. 1. Schematics of the apparatus (a) and the test section (b).

ent desired inlet vapor quality of the test section was obtained by
adjusting electric heating power in the evaporator. The saturation
state was maintained by the coincidence of measured saturation
temperatures with the saturation temperatures derived from the
saturation pressures. For different mass velocities and saturation
temperatures, the cooling water temperature at the inlet of the test
section was held constant while the cooling water ow rate was
adjusted to ensure water temperature difference in the test section
larger than 1 K. The data were collected by the data acquisition system when all the measured values were steady.
2.2. Test section
The test section is a counter ow tube-in-tube condenser. The
refrigerant condenses inside the test tube while the cooling water
ows in the annulus as shown in Fig. 1b. The test tubes are a stain-

(a) circular

less steel single circular microchannel (dh = 1.152 mm) and a stainless steel single square microchannel (dh = 0.952 mm). Cross
sectional dimensions were measured by a KYKY2800 series scanning electron microscope with the accuracy of 4.5 nm. The arithmetical mean deviations of the assessed prole Ra of the circular
and square inner surfaces are 2.0 lm and 3.2 lm. Fig. 2 shows
cross sectional views of the microchannels. The effective heat
transfer lengths of the circular and square microchannels are
0.336 m and 0.352 m.
The outer wall temperatures were measured with eight 76 lm
diameter type T thermocouples symmetrically attached on the
top and bottom outside of the microchannels. The four pairs of
thermocouples were uniformly soldered along the microchannel
as shown in Fig. 1b and the average data was used as the outer wall
temperature. The refrigerant and cooling water temperatures were
measured by inserting Pt100 temperature transducers into the
measured uids through tee ttings. All the thermocouples and
Pt100 transducers were calibrated using a 6020 Series high precision calibration bath before experiments. Mixers were set before
each uid temperature measurement point to ensure the uids
fully mixed. The pressures of the refrigerant were measured by
Trafag type 8251 pressure transducers with the accuracy of 0.3%
and the measuring range of 06 MPa. The refrigerant was drawn
out through tee ttings which connected to pressure transducers
with 3 mm inner diameter copper tubes. The pressure drop in
the test section was measured by an EJA110A differential pressure
transducer with the full scale accuracy of 0.08% and the measuring range of 0100 kPa. Mass uxes of the refrigerant and the cooling water were measured by two DMF-I Coriolis-effect mass ow
meters with the accuracy of 0.2% and the measuring range of
0.55 kg/h. All data were collected using an Agilent 34970A data
acquisition system with three 34901A cards in real time.
The saturation states at the inlet and outlet of the test section
are checked using the measured temperature and pressure. Deviations between the measured saturation temperatures and the saturation temperature derived from the saturation pressure are
within 0.12 C which can be attributed to the accuracy of the temperature and pressure transducers, the properties and purity supplied by R152a manufacturer. Therefore, the measured saturation
temperatures coincide well with the saturation temperatures derived from the saturation pressures. The deviation between measured saturation temperature and saturation temperature derived
from saturation pressure was beyond the Pt temperature sensor
accuracy, while the deviation between measured saturation pressure and saturation pressure derived from saturation temperature
was within the pressure transducer accuracy indicating that the
temperature sensor has relatively higher accuracy. Therefore, the
average saturation temperature between the inlet and outlet of
the test section was used to calculate heat transfer coefcients.

(b) square

Fig. 2. Cross sectional view of the microchannels.6.

63

N. Liu et al. / Experimental Thermal and Fluid Science 47 (2013) 6067

Aheader) and Cc is the coefcient of contraction, which is in turn a


function of the area ratio

2.3. Data reduction


The vapor mass quality of R152a at the inlet of the test section

Cc

vin is calculated from the energy balance in the evaporator

vin

UI  mr hl  hin
mr hfg

where U and I are the heating voltage and current, mr is the refrigerant mass ux, hl is specic enthalpy of the saturation liquid, hin is
specic enthalpy at the inlet of the evaporator and hlv is specic enthalpy of evaporation.
As single microchannels were used in the paper, heat transfer
rate was small when obtaining local heat transfer coefcients. In
order to ensure the cooling water temperature difference larger
than 1 K, vapor quality difference Dv in the test section was controlled to be 0.220.25 for lowest mass ux G = 200 kg/m2 s and
0.080.13 for highest mass ux G = 800 kg/m2 s during the experiments. Dv is calculated

Q
Dv water
mr hlv

2vin  Dv
2

The inner and outer wall temperature differences DTw for circular and square microchannels are expressed as

lndo =di
2pkl
d
DT w Q water
4blk

DT w Q water

4:a
4:b

where k is the thermal conductivity of the microchannel, l is the


effective heat transfer length, do and di are the outer and inner
diameters of the circular microchannel, b is the side length of the
square microchannel and d is the channel thickness.
The average heat-transfer coefcient a for the condensation of
R152a in the test section is determined by

Q water
a
AT s  T w  DT w

where Tw is the average outer wall temperature of the microchannel, Ts is the refrigerant saturation temperature and A is the cross
sectional area of the microchannel. The range of temperature difference between the saturation temperature Ts and the average outer
wall temperature Tw is 1.43.3 C.
The measured pressure drop Dpmeasured includes the frictional
loss Dpf, the expansion loss Dpexpansion, the contraction loss Dpcontraction and the deceleration loss Dpde caused by the vapor fraction
variation during condensation. Therefore, Dpmeasured is represented
as

Dpmeasured Dpf Dpexpansion Dpcontraction Dpde

Pressure drop caused by contraction Dpcontraction is estimated


using a homogeneous ow model recommended by Butterworth
and Hewitt [16]

Dpcontraction

G2

2q l

"

1
1
Cc

2

1c

#

1v



ql
1
qv

where G is the mass ux, ql and qv are the liquid and vapor densities, c is the area ratio of the test section and the header (Atest-section/

For the expansion loss, the following separated ow model recommended by Butterworth and Hewitt [13] is used

Dpexpansion

G2 c1  cws

ql

where ws, the separated ow multiplier, is a function of the two


phase densities and the vapor mass quality.
The pressure drop caused by uid deceleration Dpde is calculated using the model recommended by Carey [17]

Dpde

"
#
G2 v2 G2 1  v2

qv n
ql 1  n

vvout

"
#
G2 v2 G2 1  v2


qv n
ql 1  n

10
vvin

where the void fraction n is evaluated using Baroczy correlation [18]

"
n 1

where Qwater is the heat transfer rate of the cooling water in the test
section.
Therefore, the average vapor mass quality v in the test section is
calculated by

1
0:6391  c0:5 1


0:74  0:65  0:13 #1
1v
qv
ll

ql

11

lv

Therefore, the frictional pressure loss Dpf can be obtained by


subtracting the expansion loss Dpexpansion, the contraction loss
Dpcontraction and the deceleration loss Dpde from the measured
pressure drop Dpmeasured. Then the frictional pressure gradient is
dened as

dp Dpf

dl
l

12

According to the data reduction results, the fraction ranges of


the contraction loss Dpcontraction, the expansion loss Dpexpansion,
the frictional loss Dpf and the deceleration loss divided by measured total pressure loss Dpde are 7.315.4%, 105, 89.595.2%
and 1.75.9%. The expansion loss is small enough to neglect.
The physical properties of R152a were obtained from REFPROP
[19].
2.4. Uncertainty analysis
The uncertainty for the experimental results was determined by
a procedure proposed by Kline and McClintock [20]. Experimental
uncertainties of the measured parameters are listed in Table 2.
3. Results and discussion
3.1. Heat transfer and pressure drop of single-phase ow
Before any two phase measurements, single-phase heat transfer
and pressure drop experiments in the circular microchannel were
conducted to validate the experimental rig. The arithmetic mean
error, a.m., and root-mean-square error, r.m.s. are used to assess
Table 2
Experimental uncertainties for measured parameters.
Parameter

Uncertainty

Temperature (thermocouple)
Temperature (Pt100 transducer)
Refrigerant mass ow rate
Cooling water mass ow rate
Pressure
Pressure difference
Heat transfer coefcient

0.1 C
0.05 C
0.2% (5 kg/h)
0.2% (5 kg/h)
0.3% (6 MPa)
0.08% (100 kPa)
15.0% W/m2 K

Notes: The percentage uncertainties for refrigerant and cooling water ow rates,
pressure and pressure difference are relative to the end of scale values. The end of
scale values are reported in the table.

64

N. Liu et al. / Experimental Thermal and Fluid Science 47 (2013) 6067

the reliability of the test rig. The two parameters are dened as
follows:

80

1 X apre  aexp
a:m:
 100%
N
aexp

13

present

circular
R152a

70

Gnielinski (1976)

60

14

Nu

50

s

2
1 X apre  aexp
r:m:s:
 100%
N1
aexp

40
30
20
10
0

2000

4000

6000

8000

10000

12000

Re
Fig. 4. Comparison of measured Nu with predictions by Gnielinski correlation [19]
for single-phase ow.

20
o

18

ts = 40 C

16

t=

10 C

14

/ kW/ m2 K

where a refers to heat transfer coefcient a or pressure gradient dp/


dl and N is the number of data points.
Single-phase pressure drop experiments were conducted during
adiabatic ow of R152a with Reynolds number ranging from 800 to
1.01  104. Fig. 3 shows comparison between measured and predicted friction factors. The present data agree well with predictions
of Blasius equation [21] with the a.m. and r.m.s deviations within
1.7% and 8.8%, respectively.
Single-phase heat transfer experiments were performed with
Reynolds number ranging from 1.3  103 to 1.2  104 and temperature of liquid R152a from 27 C to 33 C. Fig. 4 shows comparison
between measured and predicted Nusselt numbers. The present
data agree well with predictions of Gnielinski correlation [22] with
the a.m. and r.m.s deviations within 4.1% and 12.1%, respectively.
Deviations of heat transfer rate between the refrigerant and
cooling water in the test section are within 5%. Single-phase
experiments show that the heat loss in the test section is rather
small and the experimental rig is appropriate for condensation
experiments.

12

circular square G / kg/(m2 s)


200
400
600
800

10
8
6

3.2. Heat transfer during condensation

3.2.1. Effects of mass ux and vapor mass quality


Fig. 5 shows the effects of mass ux and vapor mass quality on
heat transfer coefcients with the saturation temperature of 40 C
and the temperature difference between the refrigerant and the
cooling water of 10 C. During the experiments, mass uxes varied
from 200 to 800 kg/m2 s. Heat transfer coefcients increase with
increasing mass ux and vapor mass quality both in circular and
square microchannels, which agrees with the previous studies
[23]. Heat transfer coefcients become more sensitive to mass ux
at high vapor mass qualities implying the dominant effect of shear
stress. Error bars have been introduced for the heat transfer coefcients in Fig. 5.

2
0
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

1-
Fig. 5. Effects of mass ux, vapor mass quality and channel geometry on heattransfer coefcients.

20
o

circular
R152a

18
16

ts / C
40
50

G=400 kg/(m2 s)

/ kW/(m2 K)

14
0.10
present

circular
R152a

0.09

f =64/Re

0.08

f =0.3164/Re

0.25

0.07

10
8
6

0.06

0.05

0.04

0
0.0

0.03

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

1-

0.02

Fig. 6. Effect of saturation temperature on heat-transfer coefcients.

0.01
0.00

12

2000

4000

6000

8000

10000

12000

Re
Fig. 3. Comparison of measured and predicted friction factors.

3.2.2. Effect of saturation temperature


Experiments were conducted at two saturation temperatures of
40 C and 50 C in the circular microchannel. Fig. 6 shows the effect
of the saturation temperature on heat-transfer coefcients of

65

N. Liu et al. / Experimental Thermal and Fluid Science 47 (2013) 6067

3.2.4. Comparison with heat transfer correlations and a theoretical


solution
Four heat transfer correlations and a theoretical solution are
used to predict experimental data in the circular and square microchannels ([24,25,8,26,27]). Except for Wang and Rose [27] a theoretical solution, rest are based on experimental data of
refrigerants. The work of Wang and Rose [27] shows that there is
a region where heat transfer coefcient is governed only by surface
tension and the solution is derived based on the calculated results
in the region. Fig. 7 shows comparison between measured and predicted heat-transfer coefcients. Table 3 summarizes evaluation
results of heat transfer correlations and the theoretical solution
which were assessed in terms of the arithmetic mean error, a.m.,
and root-mean-square error, r.m.s.
The results for the circular microchannel agree within experimental error with Wang et al. [24] and Koyama et al. [25] correlations and the theoretical solution [27], while for the square
microchannel the results agree within experimental error with
Koyama et al. [25], Cavallini et al. [8] and Bandhauer et al. [26] correlations and the theoretical solution [27]. Fig. 7 shows that deviations between the data and predictions of Wang and Rose [27]
depend much on vapor mass quality.

150

(pre-exp)/exp100%

3.2.3. Effect of channel geometry


Fig. 5 also shows effects of channel geometry on heat-transfer
coefcients of R152a with various mass uxes. At G = 200 kg/m2 s
heat-transfer coefcients in the square microchannel are all higher
than those in the circular microchannel. On one hand, the heat
transfer enhancement is caused by effects of the corners in the
square microchannel. The condensate is pulled to the corners due
to the effect of surface tension, which reduces the average thermal
resistance across the cross section in the square microchannel. On
the other hand, the hydraulic diameter decrease of the square
microchannel (circular: dh = 1.152 mm, square: dh = 0.952 mm) increases heat-transfer coefcients according to the previous studies
of Zhang et al. [23]. At G = 400 kg/m2 s heat-transfer coefcients in
the square microchannel are still larger than those in the circular
microchannel though the enhancement is weaker compared with
that at G = 200 kg/m2 s. At G = 600 kg/m2 s heat-transfer coefcients of the two microchannels almost overlap each other. Shear
stress plays a much more important role than surface tension with
increase of mass ux. Effect of the square microchannel corners becomes weaker at higher mass uxes.

200

+30%

100
50
0
-50
-30%
-100
-150
-200
0.0

Wang et al. (2002)


Koyama et al. (2003)
Cavallini et al. (2011)
Bandhauer et al. (2006)
Wang and Rose (2011)

0.2

0.4

0.6

0.8

1.0

1-

(a) circular
200
150

(pre-exp)/exp100%

R152a. Heat-transfer coefcients decrease with increasing saturation temperature, which can be attributed to dependence of thermodynamic properties of the refrigerant on saturation
temperature. The saturation pressure and vapor density increase
with increasing saturation temperature, which increases the vapor
and liquid phase density ratio. The vapor phase velocity decreases
and the shear stress between the liquid and vapor phase decreases
with increasing saturation temperature leading to thicker condensate and lower heat-transfer coefcients.

100

+30%

50
0
-50
-100
-150
-200
0.0

-30%

Wang et al. (2002)


Koyama et al. (2003)
Cavallini et al. (2011)
Bandhauer et al. (2006)
Wang and Rose (2011)

0.2

0.4

0.6

0.8

1.0

1-

(b) square
Fig. 7. Comparison of measured and predicted heat-transfer coefcients.

Table 3
Performance of heat transfer correlations and a theoretical solution compared with
the present data.
Circular

Wang et al. [24] (2002) correlation


Koyama et al. [25] (2003) correlation
Cavallini et al. [8] (2011) correlation
Bandhauer et al. [26] (2006) correlation
Wang and Rose [27] (2011) theoretical
solution

Square

a.m.

r.m.s

a.m.

r.m.s

11.2
12.7
17.9
29.3
14.6

15.4
29.6
27.4
14.1
52.3

26.3
5.4
0.9
7.4
4.3

16.3
5.6
19.8
21.9
30.0

a.m.: Arithmetic mean error (%); r.m.s.: root-mean-square error (%).

3.3. Pressure drop during condensation


3.3.1. Effects of mass ux and vapor mass quality
Fig. 8 shows effects of mass ux and vapor mass quality on frictional pressure gradients both in circular and square microchannels. The frictional pressure gradients increase with mass ux
and vapor mass quality which is similar to heat-transfer coefcients shown in Fig. 5. Pressure gradient trends become smooth
at high vapor mass qualities. The mist ow regime with droplet entrained in the vapor core emerges at high vapor mass qualities
according to results of Garimella and Coleman [28] where pressure
gradients increase less with vapor mass quality.

3.3.2. Effect of saturation temperature


Fig. 9 shows effect of saturation temperature on frictional pressure gradients of R152a in the circular microchannel. The frictional
pressure gradient decreases with increasing saturation temperature similar to that for heat-transfer coefcients in Fig. 6. The saturation pressure increases with the saturation temperature leading
to higher reduced pressure and larger vapor and liquid phase density ratio. The decrease of the velocity difference between the two
phases decreases the frictional pressure gradient.

66

N. Liu et al. / Experimental Thermal and Fluid Science 47 (2013) 6067

300

200

[(dp/dl)pre-(dp/dl)exp]/(dp/dl)exp100%

dp/dl / kPa/m

200
400
600
800

ts =40 oC

250

240

G / kg/(m2 s)

circular square

R152a

150

100

50

0
0.0

0.2

0.4

0.6

0.8

180
120
60

+30%

0
-60

-30%
-120
-180
-240
0.0

1.0

Koyama et al. (2003)


Agarwal & Garimella (2009)
Cavallini et al. (2009)

0.2

0.4

1.0

circular

ts / C
2

G=400 kg/(m s)

240

[(dp/dl)pre-(dp/dl)exp]/(dp/dl)exp100%

120

40
50

80

dp/dl / kPa/m

0.8

(a) circular

Fig. 8. Effects of mass ux, vapor mass quality and channel geometry on frictional
pressure gradients.

100

0.6

1-

1-

60

40

180
120

+30%

60
0
-60

-30%

-120
-180
-240
0.0

20

Koyama et al. (2003)


Agarwal & Garimella (2009)
Cavallini et al. (2009)

0.2

0.4

0.8

0.6

1.0

1-
0
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

(b) square

1.0

1-

Fig. 10. Comparison of measured and predicted frictional pressure gradients.

Fig. 9. Effect of saturation temperature on frictional pressure gradients.

3.3.3. Effect of channel geometry


Effect of channel geometry on frictional pressure gradients is
shown in Fig. 8. Frictional pressure gradients of the square microchannel are slightly lower than those in the circular microchannel
at higher vapor mass qualities for the same operating conditions.
However, at lower vapor mass qualities, frictional pressure gradients are almost equivalent in the two microchannels. Therefore,
channel geometry has little effect on frictional pressure gradients.
3.3.4. Comparison with pressure drop correlations
Three pressure drop correlations [25,29,30] based on experimental data of R134a in microchannels were used to predict experimental data. As 2 Ra/Dh was larger than 0.0027, zero surface
roughness Ra was inserted in the Cavallini et al. [30] equations
when computing pressure gradients, while the surface roughness
was not needed in Koyama et al. [25] and Agarwal and Garimella
[29] equations. Comparison results are shown in Fig. 10. Table 4
summarizes the evaluation results of pressure drop correlations.
The performance of each correlation was assessed in terms of the
arithmetic mean error, a.m., and root-mean-square error, r.m.s.
Koyama et al. [25] underestimates the data for both microchannels
while Agarwal and Garimella [29] overestimate the data for the
square microchannel. Predictions of Cavallini et al. [30] show large

Table 4
Performance of pressure drop correlations compared with the present data.
Correlations

Koyama et al. [23] (2003) correlation


Agarwal and Garimella [29] (2009)
correlation
Cavallini et al. [6] (2009) correlation

Circular

Square

a.m.

r.m.s.

a.m.

r.m.s

78.1
4.2

6.9
20.5

64.5
34.2

14.0
20.1

25.5

43.9

11.7

41.0

a.m.: Arithmetic mean error (%); r.m.s.: root-mean-square error (%).

root-mean-square errors for data in both circular and square


microchannels.

4. Conclusions
Heat transfer and pressure drop during condensation of R152a
were investigated experimentally in a circular (d = 1.152 mm)
and a square (dh = 0.952 mm) microchannels. Experiments were
conducted with mass uxes from 200 to 800 kg/m2 s, saturation
temperatures of 40 C and 50 C and vapor mass qualities from
0.1 to 0.9. Experimental data were compared with empirical correlations and a theoretical solution. The following conclusions can be
drawn:

N. Liu et al. / Experimental Thermal and Fluid Science 47 (2013) 6067

1. Heat-transfer coefcients and pressure gradients during condensation increase with mass ux and vapor mass quality while
decrease with the saturation temperature both in circular and
square microchannels.
2. Heat-transfer coefcients of the square microchannel are higher
than those of the circular microchannel at G = 200 kg/m2 s and
400 kg/m2 s due to the effect of surface tension. The heat transfer enhancement decreases with mass ux as the shear stress
plays a much more important role at higher mass uxes. However, channel geometry has little effect on two phase pressure
gradients.
3. The heat transfer results for the circular microchannel agree
within experimental error with Wang et al. [24] and Koyama
et al. [25] correlations and the theoretical solution [27], while
for the square microchannel the results agree within experimental error with Koyama et al. [25], Cavallini et al. [8] and
Bandhauer et al. [26] correlations and the theoretical solution
[27]. For pressure drop results, Koyama et al. [25] underestimates the data for both microchannels while Agarwal and Garimella [29] overestimate the data for the square microchannel.

Acknowledgements
The work was nancially supported by the National Basic Research Program of China (973 Project, No. 2011CB706904),
Guangdong
Industry-Academia-Research
Project
(No.
2011A090200018) and the new energy vehicles industry Project
(2011) of Guangdong Special Funds for Strategic Emerging
Industries.
The work was also supported by EU research Grant FP7-2010IRSES-269205.
References
[1] Regulation (EG) No 842/2006 of the European Parliament and of the Council of
17 May 2006 on certain uorinated greenhouse gases, Ofcial Journal of the
European Union, 2006.
[2] Directive 2006/40/EC of The European Parliament and of the Council of 17 May
2006 relating to emissions from air-conditioning systems in motor vehicles
and amending Council Directive 70/156/EC, Ofcial Journal of the European
Union, 2006.
[3] Q. Su, G.X. Yu, H.S. Wang, J.W. Rose, Microchannel condensation: correlations
and theory, International of Refrigeration 32 (2009) 11491152.
[4] E.E. Wilson, A basis for rational design of heat transfer apparatus, ASME
Transactions 37 (1919) 4770.
[5] M. Matkovic, A. Cavallini, D. Del Col, L. Rossetto, Experimental study on
condensation heat transfer inside a single circular minichannel, International
Journal of Heat and Mass Transfer 52 (2009) 23112323.
[6] A. Cavallini, D. Del Col, M. Matkovic, L. Rossetto, Pressure drop during twophase ow of R134a and R32 in a single minichannel, Journal of Heat Transfer
131 (2009) 03310710331078.
[7] D. Del Col, D. Torresin, A. Cavallini, Heat transfer and pressure drop during
condensation of the low GWP refrigerant R1234yf, International Journal of
Refrigeration 33 (2010) 13071318.

67

[8] A. Cavallini, S. Bortolin, D. Del Col, M. Matkovic, L. Rossetto, Condensation heat


transfer and pressure losses of high and low pressure refrigerants owing in a
single circular minichannel, Heat Transfer Engineering 32 (2) (2011) 9098.
[9] D. Del Col, S. Bortolin, A. Cavallini, M. Matkovic, Effect of cross sectional shape
during condensation in a single square minichannel, International Journal of
Heat and Mass Transfer 54 (2011) 39093920.
[10] M. Derby, H.J. Lee, Y. Peles, M.K. Jensen, Condensation heat transfer in square,
triangular and semicircular mini-channels, International Journal of Heat and
Mass Transfer 55 (2012) 187197.
[11] S.M. Kim, I. Mudawar, Theoretical model for annular ow condensation in
rectangular microchannels, International Journal of Heat and Mass Transfer 55
(2012) 958970.
[12] S.M. Kim, J. Kim, I. Mudawar, Flow condensation in parallel micro-channels
Part 1: Experimental results and assessment of pressure drop correlations,
International Journal of Heat and Mass Transfer 55 (2012) 971983.
[13] S.M. Kim, I. Mudawar, Flow condensation in parallel micro-channels Part 2:
Heat transfer results and correlation technique, International Journal of Heat
and Mass Transfer 55 (2012) 984994.
[14] A. Agarwal, T.M. Bandhauer, S. Garimella, Measurement and modeling of
condensation heat transfer in non-circular microchannels, International
Journal of Refrigeration 33 (2010) 11691179.
[15] H.S. Wang, J. Ding, J.W. Rose, Heat transfer during annular lm condensation in
microchannels: calculations for R152a, R134a, R22, R410A, propane, ammonia
and carbon dioxide, Sixth International Conference on Enhanced, Compact and
Ultra-Compact Heat Exchangers: Science, Engineering and Technology 2
(2007) 1216.
[16] G.F. Hewitt, D. Butterworth, Two Phase Flow and Heat Transfer, Oxford Press,
1977.
[17] C.J. Carey, LiquidVapor Phase Change Phenomena, Taylor & Francis Series in
Mechanical Engineering, Hemisphere Publishing, 1992.
[18] C.J. Baroczy, Correlation of liquid fraction in two-phase ow with application
to liquid metals, Atomics International, 1963.
[19] NIST Thermodynamic and Transport Properties of Refrigerants and
Refrigerants Mixtures REFPROP, Version 7.0.
[20] S.J. Kline, F.A. McClintock, The description of uncertainties in single sample
experiments, Mechanical Engineering 75 (1953) 39.
[21] H. Blasius, Grenzschichten in ussigkeiten mit kleiner reibung, Zeit. Math.
Phys. 56 (1908) 137.
[22] V. Gnielinski, New equations for heat and mass transfer in turbulent pipe and
channel ow, International Chemical Engineering 16 (2) (1976) 359368.
[23] H.Y. Zhang, J.M. Li, N. Liu, B.X. Wang, Experimental investigation of
condensation heat transfer and pressure drop of R22, R410A and R407C in
mini-tubes, International Journal of Heat and Mass Transfer 55 (2012) 3522
3532.
[24] W.W. Wang, T.D. Radcliff, R.N. Christensen, A condensation heat transfer
correlation for millimeter-scale tubing with ow regime transition,
Experimental Thermal and Fluid Science 26 (2002) 473485.
[25] S. Koyama, K. Kuwahara, K. Nakashita, K. Yamamoto, An experimental study on
condensation of refrigerant R134a in a multi-port extruded tube, International
Journal of Refrigeration 24 (2003) 425432.
[26] T.M. Bandhauer, A. Agawal, S. Garimella, Measurement and modeling of
condensation heat transfer coefcients in circular microchannels, Journal of
Heat Transfer 128 (2006) 10501059.
[27] H.S. Wang, J.W. Rose, Theory of heat transfer during condensation in
microchannels, International Journal of Heat and Mass Transfer 54 (2011)
25252534.
[28] J.W. Coleman, S. Garimella, Characterization of two-phase ow patterns in
small diameter round and rectangular tubes, International Journal of Heat and
Mass Transfer 42 (1999) 28692881.
[29] A. Agarwal, S. Garimella, Modeling of pressure drop during condensation in
circular and noncircular microchannels, Journal of Fluids Engineering 131
(2009) 01130210113028.
[30] A. Cavallini, D. Del Col, M. Matkovic, L. Rossetto, Frictional pressure drop
during vapourliquid ow in minichannels: modeling and experimental
evaluation, International Journal of Heat and Fluid Flow 30 (2009) 131139.

You might also like