Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

International Journal of Heat and Mass Transfer 53 (2010) 4141–4151

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


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

Characterization of heat transfer modes of tube side convective condensation


under the influence of an applied DC voltage
H. Sadek, C.Y. Ching *, J. Cotton
Dept. of Mechanical Engineering, McMaster University, Hamilton, Canada L8S 4L7

a r t i c l e i n f o a b s t r a c t

Article history: An experimental study was performed to determine the modes of heat transfer in horizontal tube side
Received 12 March 2010 convective condensation under an applied DC high voltage. Experiments were performed with R134a
Received in revised form 13 May 2010 in a horizontal, single-pass, counter-current heat exchanger. A 8 kV DC voltage was applied through a
Available online 11 June 2010
central rod electrode with the tube wall grounded for mass flux in the range 45–156 kg/m2 s. The average
quality across the test section was 45%, which corresponds to an initially stratified flow with the liquid
Keywords: level below the central electrode. The high-voltage electric field results in an interfacial electric body
Two-phase flow
force which extracts the liquid from the bottom stratum towards the vapour core. This results in a phase
Electrohydrodynamic
Heat transfer
redistribution with a resultant increase in heat transfer. The increase in heat transfer can be attributed to
Condensation a change of the heat transfer mode in the lower section of the tube from convective condensation to film
condensation.
Ó 2010 Elsevier Ltd. All rights reserved.

   
1. Introduction f 000 ¼ q E  1 E2 re þ 1 r qE2 @ e : ð1Þ
e e
2 2 @q T
There has been significant interest in condensation heat transfer
The three terms on the right-hand side of (1) represent the electro-
enhancement techniques due to the advantages it offers in terms of
phoretic, dielectrophoretic, and electrostrictive components of the
increased performance and the possibility of more compact heat ex-
electric body force respectively. The electrophoretic component
changer designs. Among the various enhancement techniques, the
represents the force acting on the free charges in the presence of
use of electrohydrodynamics (EHD) is appealing because it can be
an electric field (also known as the Coulomb force). This force is
configured relatively non-intrusively using simple electrode de-
usually dominant for single-phase flow and can contribute to heat
signs. It is a relatively robust technique since it is non-mechanical,
transfer in two phase flows by increasing convection in the individ-
and allows a high degree of local control. Electrohydrodynamics
ual phases. The dielectrophoretic component represents the force
refer to the coupling between the flow field and a high-voltage elec-
due to the spatial change of the permittivity of the dielectric med-
tric field. This coupling induces electric body forces within the fluid
ium as a result of temperature gradients and/or differences in the
which can affect the two-phase flow and therefore heat transfer and
phases. The electrostrictive force is caused by both the inhomoge-
pressure drop. These effects can be divided into interfacial effects
neity in the electric field strength and the variation in the dielectric
and bulk effects in the separate phases [1–4]. The effect of EHD at
constant of the medium with temperature and density. In two-
the interface includes inducing perturbations and waviness into
phase flows such as in convective boiling and condensation, the
the interface [5–8], causing liquid extraction which causes thinning
dielectrophoretic and electrostrictive forces (polarization forces)
of the liquid phase thickness [9,10], inducing droplet entrainment
are dominant and significant at the vapour–liquid interface due to
[8], and promoting pseudo-dropwise condensation [9,11]. For the
the large difference in permittivity between the two phases
bulk of the phases, the effect of EHD includes enhancement of con-
[14,15]. This force can cause interfacial instabilities and force the li-
vection in both phases [1], generation of turbulence [2], and
quid with higher permittivity to move to regions of higher electric
enhancement of vapour bubble/liquid droplet motions [12].
field. This phenomenon is usually referred to as liquid extraction
The total EHD force induced by the electric fields on the fluid is
[16] and can result in phase redistribution in two-phase flows.
expressed as [1,13]
The relative importance of the different components of the EHD
body force can be determined by evaluating the key dimensionless
numbers in the governing differential momentum equation. The
additional electric body force in the momentum equation can be
* Corresponding author. Tel.: +1 905 525 9140x24998. represented by the EHD number, Ehd, and the Masuda number,
E-mail address: chingcy@mcmaster.ca (C.Y. Ching). Md [1,17];

0017-9310/$ - see front matter Ó 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijheatmasstransfer.2010.05.037
4142 H. Sadek et al. / International Journal of Heat and Mass Transfer 53 (2010) 4141–4151

Nomenclature

A area (m2) Tsbottom average bottom surface temperature (K)


AEHD area occupied by the liquid in contact with the heat Tstop average top surface temperature (K)
transfer surface with EHD (m2) Tsat saturation temperature (K)
cppw specific heat of the water through the preheater (J/kg K) Tw average water temperature in the test section water
cpref specific heat of the refrigerant (J/kg K) jacket (K)
cpw specific heat of the water through the test section (J/ TwPRE,i water temperature at preheater inlet (K)
kg K) TwPRE,o water temperature at preheater outlet (K)
D diameter (m) V voltage (V)
E electric field strength (V/m) x mass vapour quality
Ehd EHD number DTRTD RTD temperature difference across the test section (K)
Ex extraction ratio
fi interfacial surface roughness Greek
fe000 electric body force per unit volume (N/m3) a void fraction
g acceleration due to gravity (m/s2) ah homogeneous void fraction
G mass flux (kg/m2s) ara Rouhani and Axelsson void fraction
h heat transfer coefficient (W/m2 K) d liquid film thickness (m)
hc convective condensation heat transfer coefficient (W/ dEHD liquid film thickness with EHD (m)
m2 K) e dielectric permittivity (N/V2)
hf Nusselt film condensation heat transfer coefficient (W/ eo dielectric permittivity of free space (8.854  1012) (N/
m2 K) V2)
hL stratum liquid level height (m) es dielectric constant
hlv latent heat of vaporization (kJ/kg) h angle (rad)
ho heat transfer coefficient without no applied voltage (W/ hstrat stratified angle around upper perimeter of the tube
m2 K) (rad)
I current (A) l dynamic viscosity (kg/ms)
k thermal conductivity (W/m K) le ion mobility (m2/V s)
L length scale (m) m kinematic viscosity (m2/s)
Md Masuda number q density (kg/m3)
m_w water mass flow rate in the test section (kg/s) qe charge density (C/m3)
m_ ref refrigerant mass flow rate (kg/s) r surface tension (N/m)
m_ wPRE water mass flow rate in the water preheater (kg/s)
Pr Prandtl number Subscripts
q heat transfer rate (W) b, bottom bottom
qPRE heat transfer rate to the refrigerant in the electric and in inlet
water preheaters (W) l, L liquid
qw test section heat transfer rate (W) o reference
R thermal resistance (W/K) out outlet
Re Reynolds number based on the hydraulic diameter ref refrigerant
T temperature (K) t, top top
TRsub refrigerant inlet temperature (subcooled) (K) v, V vapour
TRCONDo refrigerant temperature at condenser outlet (K) w water
T Sav g average surface temperature (K)

I o L3 E2 eo L2 T o ð@ es =@TÞ Seyed-Yagoobi and Bryan [19], Laohalertdecha et al. [4]. The effect
Ehd ¼ Md ¼ : ð2Þ of DC electric fields on in-tube convective condensation heat trans-
q o m 2 le A 2qo m2
fer using a concentric wire electrode configuration was studies by
The dimensional analysis indicates that the electrohydrodynamic Singh et al. [20], Gidwani et al. [21], Feng and Seyed-Yagoobi [22],
forces will have a significant effect on the fluid flow if Md/Re2  1 Sadek et al. [23], Laohalertdecha et al. [24]. In general, EHD enhances
and/or Ehd/Re2  1. The evaluation of Md and Ehd numbers is heat transfer, and the augmentation level decreases with an increase
dependant on the nature of the flow (single phase or two phase), of the mass flux and the vapour quality. The heat transfer enhance-
electrode geometry, fluid properties and the applied voltage. In ment in two-phase flow is mainly attributed to the liquid extraction
the current investigation, the vapour EHD number and the vapour from the proximity of the heat transfer surface to the bulk of the flow
interfacial Masuda number (where the electric field strength is eval- caused by the induced interfacial electric forces. The amount of li-
uated for the vapour phase at the liquid–vapour interface) were quid extracted to the core, however has not been quantified. While
used to determine the effect of the different components of the there is agreement in the literature that EHD presents a promising
EHD force on the forced convective flow. The comparison between active technique for heat transfer enhancement, there is a lack of
the vapour EHD number and the interfacial Masuda number shows understanding of how the high voltage electric fields affect the heat
that the interfacial Masuda number is four order of magnitude high- transfer modes. In order to apply EHD as a heat transfer enhance-
er than the vapour EHD number which implies that the dominant ment/control technique in practical heat exchangers, a better phys-
force acting on the liquid–vapour interface are the dielectrophoretic ical understanding of the heat transfer mechanisms under the effect
and electrostrictive forces. of applied high voltage is required.
There have been several studies on EHD heat transfer enhance- Conventional condensation heat transfer modes (without EHD)
ment and detailed reviews are given by Allen and Karayiannis [18], are classified as: (i) gravity dominated where the gravity forces
H. Sadek et al. / International Journal of Heat and Mass Transfer 53 (2010) 4141–4151 4143

dominate over the vapour shear forces and (ii) shear dominated of the condenser. The refrigerant inlet quality to the test section
where the vapour shear forces overcome the gravity forces. In can be controlled by two separate heating sections; direct electric
the gravity dominated flow regime (stratified, wavy, and slug heating and water heating in a plate-type heat exchanger. The
flow), the heat transfer modes are typically laminar film condensa- refrigerant exiting the heat exchanger is directed into the horizon-
tion in the top section of the tube and forced-convective condensa- tal test section via a straight length of 0.55 m (50 diameters) tubing
tion in the bottom section of the tube. This regime is characterized to achieve fully developed conditions. Leaving the test section, the
by heat transfer coefficients that depend on the wall-to-refrigerant refrigerant enters a 30 kW coaxial single-pass condenser to return
temperature difference but nearly independent of mass flux. In the the refrigerant two-phase mixture to its original liquid state before
shear-dominated flow regime (annular flow), the dominant heat entering the pump. The test section consists of a horizontal single
transfer mechanism is forced-convective condensation. This re- tube counter-current heat exchanger as depicted in Fig. 2. The out-
gime is characterized by heat transfer coefficients that are inde- er water jacket is constructed from a 30 cm length of clear PVC
pendent of temperature difference but very dependent on mass with inner diameter of 19.1 mm and wall thickness of 3.8 mm.
flux and quality. The objective of the current investigation is to The inner tube containing the two phase refrigerant is a stainless
characterize the heat transfer modes during convective condensa- steel tube with an outer diameter of 12.7 mm and an inner diam-
tion with and without EHD. The dominant heat transfer modes un- eter of 10.2 mm. A 3.2 mm diameter stainless steel rod electrode
der the effect of DC applied voltage are identified by comparing the was used to apply the electric field across the annular gap formed
experimental heat transfer values to the ones predicted by the con- by the electrode and the surface of the inner stainless steel tube.
ventual heat transfer models. The electrode is concentric with the inner tube forming an annulus
as shown in Fig. 2. The width of the annular gap is maintained
2. Experimental facility along the entire length of the test section by three non-conducting
spacers at intervals of 75 mm. The spacers were machined out of
The present investigation utilizes an existing two-phase flow non-conducting material Delrin (dielectric constant of approxi-
test facility most recently used by Sadek et al. [23]. The test section mately 2.5) and occupy approximately 20% of the cross-sectional
used by Sadek et al. [23] was shortened from 150 cm to 30 cm, flow area. The spacer design is shown in Fig. 2. The outside surface
resulting in smaller changes of vapour quality across the test sec- temperature of the inner test-section tube is measured with
tion. Thus, the advantage of this test section is that it allows a 0.5 mm O.D. T-Type thermocouples embedded in the tube at three
study of the effect of EHD on condensation with a nearly constant axial locations, 75 mm apart. Each location contains four thermo-
average quality across the test section. For example, the difference couples, at the bottom, top and the two sides. The inlet and outlet
between the inlet and outlet qualities was less than 7% for a heat temperatures of the refrigerant and the water are measured by
flux of 6 kW/m2. The test facility is a closed loop charged with 1.6 mm O.D. T-Type thermocouple probes for the refrigerant side
refrigerant R-134a as shown schematically in Fig. 1. Refrigerant and 6 mm O.D. resistance temperatures detectors (RTD) for the
is circulated through the loop by a gear pump located at the exit water side, as shown schematically in Fig. 2. The water flow rate

T
Relief Valve
~ Rotameter
T Condenser
Water Water Out
I
In
Chiller
Pump Tank
T
Sight
T Tube
T
P
Water Out Water In

Condenser
Turbine PID
Water Outlet
Flow Controller
Meters

T
~
Test Section T P Relief Valve

T
T
High Voltage Pressure Transducer
Water
Amplifier Preheater
T

Relief Valve P T
~ Ceramic Isolation Fittings
Constant
Preheated
P T Sections
Heat Flux ~
Evaporators P T Relief Valve

Filter
Turbine Flow Meters

Gear
Pump

Fig. 1. Schematic diagram of test facility.


4144 H. Sadek et al. / International Journal of Heat and Mass Transfer 53 (2010) 4141–4151

Dimensions in mm
200 250 300
Water
outlet
R134a
inlet

T1s T2s T3s


Viewing
Refrigerant Water
window for 75 75 75 75
outlet inlet
highspeed
camera
Pressure
Transducer

Ø 3.18 Wire Electrode


Ts-t

Ts-s2 Ts-s1 10.2 12.7 19.05 26.67

Ts-b

Spacer Design

Fig. 2. Schematic diagram of the test section.

to the test section is measured using a turbine flow meters. The vective thermal resistance, the cross sectional thermal resistance
pressure drop across the test section is measured using a differen- circuit can be represented by Fig. 3 where bottom refers to the
tial pressure transducer, with an accuracy of 0.25% of the full scale stratified region. The thermal resistances can be expressed as
connected to the test section via 6.35 mm O.D. tubing. The tubing
1 1
is passed through a water jacket to condense any vapour in the Rwtop ¼ Rwbottom ¼ ;
Atop hwtop Abottom hwbottom
tubes in order to equalize the hydrostatic head on each side of
the transducer. 1 1
Rref top ¼ Rref bottom ¼ ; ð3Þ
A viewing section is located at the outlet of the test section for Atop href top Abottom href bottom
visual observations of the flow. A high speed camera with a maxi- where the Abottom and Atop are the inner tube surface area in contact
mum recording rate of 2000 fps is used for visualizing and record- with the liquid stratum and the falling film respectively. The water
ing the flow pattern. The viewing section was made of transparent side heat transfer coefficient is assumed uniform, and therefore the
quartz tube coated with an electrically transparent conductive film ratio between the overall heat transfer from the top to the bottom
of tin oxide. The grounded transparent tube ensures the continua- sections can be calculated as
tion of the electric field within the viewing section. The electric
field was established by connecting the voltage output from a high qt Atop hwtop ðT stop  T w Þ Atop ðT stop  T w Þ
voltage amplifier to the center electrode. ¼ ¼ ; ð4Þ
qb Abottom hwbottom ðT sbottom  T w Þ Abottom ðT sbottom  T w Þ
The chilled water loop, needed to supply the cooling water to
the test section, consists of a chiller, turbine flow meters and resis- where Tsbottom and Tstop are obtained by averaging the surface
tance temperatures detectors (RTD). The cooling capacity of the temperatures over the bottom and top sections respectively.
chiller is from 3.2 kW to 7.3 kW for a temperature range from The total overall heat removed by the cooling water from the
5 °C to 35 °C. The inlet and outlet water temperatures to the test test section is determined by
section are measured using 6 mm O.D RTDs with uncertainty
_ w cpw ðDT RTD Þ
qw ¼ qt þ qb ¼ AhðT sat  T Sav g Þ ¼ m ð5Þ
of ±0.03 °C. The heat flux extracted from the test section can be
controlled by controlling either or both the water flow rate and/ where T Sav g is the circumferentially and axially averaged surface
or the outlet temperature from the chiller. The water flow rate is temperature over the test section. Rearranging Eqs. (4) and (5) re-
controlled by diverting a portion of the flow exiting the chiller back sults in
to the chiller reservoir via a bypass as illustrated in Fig. 1. The
 
water outlet temperature from the chiller can be controlled in Atop ðT stop  T w Þ
the range 5–35 °C with resolution of 0.5 °C. qb 1 þ ¼ AhðT sat  T Sav g Þ ð6Þ
Abottom ðT sbottom  T w Þ

or
3. Data reduction
AhðT sat  T Sav g Þ
Excluding the thermal resistance of the stainless steel tube, qb ¼ Atop ðT stop T w Þ
: ð7Þ
1þA
which is one order of magnitude lower than that for the inner con- bottom ðT sbottom T w Þ
H. Sadek et al. / International Journal of Heat and Mass Transfer 53 (2010) 4141–4151 4145

Water
Tw jacket
Ts-t Rref-top Ts-top Rw-top

Tsat

Tsat
Tw Ts-s1 Ts-s2 Tw Tw

-
R ref-bottom Ts-bottom Rw-bottom
Tw

Fig. 3. Cross sectional thermal circuit for the test section.

Therefore the overall top, bottom and total heat transfer coefficients To confirm the accuracy of the experimental measurements, the
are calculated as power extracted from the refrigerant by the condenser and the test
qt qb section condenser was compared with the power input by the pre-
ht ¼  ; hb ¼ ; heaters and the surroundings. The energy balance was found to be
Atop T sat  T stop Abottom ðT sat  T sbottom Þ
within ±15% for all test conditions, providing confidence in the effi-
qw
h¼  ; ð8Þ cacy of the measurement and analysis techniques. An uncertainty
A T sat  T Sav g
analysis of the experimental results was performed as outlined by
where Kline and McClintock [26], and are summarized in Table 1.

Atop h
¼ ð9Þ 4. Results and discussion
Abottom 2p  h
and h (Fig. 3) is estimated using the correlation from [25] as 4.1. Two-phase flow patterns
h ¼ hstrat
8 h i9 Identifying the flow pattern is necessary to predict heat transfer
>
< pð1  aÞ þ ð1:5pÞ1=3 1  2ð1  aÞ þ ð1  aÞ1=3  a1=3 >
= in two phase flows as the heat transfer characteristics depend on
¼ 2p  2 h i : the distribution of the two phases. The flow patterns deduced from
>
:  200
1
ð1  aÞa½1  2ð1  aÞ 1 þ 4ðð1  aÞ2  a2 Þ : >;
flow visualization and the surface temperature measurements in
ð10Þ this study were in good agreement with the flow pattern map of
El-Hajal et al. [25] as modified for an annular geometry (Fig. 4).
The void fraction, a, is calculated using the logarithmic mean void The mass flux for the current experiments is in the range 45 kg/
fraction (LMe) proposed by El-Hajal et al. [25], m2 s to 156 kg/m2 s, with inlet and outlet vapour qualities of 48%
ah  a ra and 42%, respectively. The relatively small change of vapour qual-
a¼ ah
; ð11Þ ity across the test section has no significant effect on the two-
ln ara phase flow patterns and therefore no transitions are expected
along the test section. The two-phase flow pattern for these flow
where ah and ara can be expressed as [25]
conditions is stratified wavy for Gh147 kg/m2 s and annular flow
  !1
x x 1x 1:18ð1  xÞ½g rðql  qv Þ0:25 for G > 147 kg/m2 s as shown in Fig. 4. At the lower mass flux,
ara ¼ ½1 þ 0:12ð1  xÞ þ þ the flow pattern is close to the stratified transition line and moves
qv qv ql Gq0:5
l
toward the transition line between stratified wavy and annular
ð12Þ flow as the mass flux is increased. The flow patterns are corrobo-
The refrigerant inlet quality is obtained from the energy balance rated from the tube side circumferential surface temperature dis-
across the electric and water preheaters such that tribution. The time averaged top, side and bottom surface
subcooled temperatures (i.e. difference between the saturation
qPRE
m_ ref  cpref ðT sat  T Rsub Þ temperature and the surface temperature) averaged over the three
xin ¼ ; ð13Þ axial locations for the different mass fluxes are shown in Fig. 5.
hlv ;in
Representative flow images and schematic diagrams of the flow
where qPRE is the net heat input by the electric and water preheaters patterns are also shown as. In general, a higher surface subcooled
as temperature indicates a higher thermal resistance at the heat

transfer surface. The bottom surface subcooled temperatures are
qPRE ¼ ½IðV 1 þ V 2 Þ þ m_ wPRE cppw ðT wPRE;i  T wPRE;o Þ : ð14Þ
higher than those for the top and side temperatures which indi-
The maximum pressure drop across the test section is less than
0.4% of the absolute pressure, therefore the effect of the change in Table 1
the pressure between the inlet and the exit of the test section on Uncertainties in calculated parameters.
the two-phase refrigerant properties can be neglected. The refrig-
Parameter Maximum percentage error
erant outlet quality is calculated by applying an energy balance
Heat flux (W/m2) ±8.32%
across the test section such that
Heat transfer coefficient (W/m2 K) ±14%
qw Reynolds number ±3%
xout ¼ xin  : ð15Þ Average quality ±10.38%
_ ref hlv
m
4146 H. Sadek et al. / International Journal of Heat and Mass Transfer 53 (2010) 4141–4151

400

350

300

Mass flux (kg/m2s)


Intermittent
250

200

150 Annular

100 Stratified-
Stratified
Wavy
50
Stratified
0
0.0 0.2 0.4 0.6 0.8 1.0
Vapour quality
Fig. 4. Flow pattern map for R-134a adopted for annular geometries. Do = 10.92 mm, Di = 3.18 mm, Tsat = 25 °C. —— transition lines from El-Hajal et al. [25], M stratified wavy
experimental data.  electrode boundaries.

Fig. 5. Time-averaged subcooled surface temperatures without EHD forces for a constant heat flux of q00 = 5.7 kW/m2 and average quality of xavg = 45%. h top, } bottom, M
side.

cates that the flow is either stratified or stratified wavy with a extraction from the bottom stratum and from the side film conden-
resultant higher as shown in the flow images. The side surface sub- sate which diminish the thermal resistance and decreases the sub-
cooled temperatures are higher than the top surface subcooled cooled temperatures. For mass flux less than 120 kg/m2 s, the top
temperatures due to the thicker liquid condensate film at the sides surface subcooled temperatures at 8 kV are less than those at
compared to the top. Increasing the mass flux increases the inertial 0 kV which can be attributed to liquid extraction from the top film
forces which increases the interfacial instabilities and causes liquid condensate as well as the bottom stratum and the side liquid film
droplet entrainment, and therefore the bottom liquid stratum get to the core of the two-phase flow. For high mass flux (more than
thinner and the top, bottom and side subcooled temperatures get 120 kg/m2 s), the top surface subcooled temperatures at 8 kV are
closer to each other. greater than those at 0 kV. This is because the vapour inertial forces
The average top, side and bottom surface subcooled tempera- drive the extracted liquid towards the heat transfer surface area
tures for different mass fluxes and fixed average quality of 45% and therefore the top liquid condensate film gets thicker with a
with an 8 kV applied voltage are shown in Fig. 6. All three temper- resultant increase in the subcooled surface temperatures.
atures are similar for a given mass flux for the entire range of mass
flux, suggesting the presence of a relatively uniform annular liquid 4.2. Heat transfer
film along the circumference, as corroborated by the flow images
(Fig. 6). The bottom and side surface subcooled temperatures at Two principal heat transfer mechanisms can be identified for
8 kV are less than those without EHD force (Fig. 5) for the entire stratified and stratified wavy flow when there is no applied voltage
range of mass flux. This implies that the EHD forces cause liquid [27]: (i) convective condensation associated with the axial flow of
H. Sadek et al. / International Journal of Heat and Mass Transfer 53 (2010) 4141–4151 4147

4.5

ed surface temperature (oC)


4

3.5

2.5

1.5
Subccoled

0.5

0
0 20 40 60 80 100 120 140 160 180

Mass flux (kg/m2s)


Fig. 6. Time-averaged subcooled surface temperatures with 8 kV applied voltage, for a constant heat flux of q00 = 5.7 kW/m2 and average quality of xavg = 45%. h top, } bottom,
M side.

the bottom liquid stratum and (ii) film condensation associated heat transfer coefficients from the model predications being 6.5%
with the thin liquid condensate falling around the periphery of and 14.7%, respectively. The model under predicts the top heat
the tube. The experimental results for the top and bottom heat transfer, which can be attributed to neglecting the effect of the axial
transfer coefficients for average quality of 45% are compared to vapour shear on the film condensation heat transfer.
the model predictions of Thome et al. [27] in Fig. 7. The two-phase The top heat transfer coefficient is higher than that at the bot-
flow pattern along the test section is stratified wavy and therefore tom over the entire mass flux range. This indicates that the heat
the overall top heat transfer coefficient is calculated assuming film transfer of the falling film condensate is greater than the convec-
condensation as [27] tive condensation of the liquid stratum for this range of mass flux.
" #0:25 Increasing the mass flux increases the top heat transfer coefficient
q ðq  qv Þghlv k3l till a mass flux of 100 kg/m2 s and then it remains almost constant
hf ¼ 0:728 l l ; ð16Þ
Dll ðT sat  T stop Þ for higher mass fluxes. Film condensation heat transfer is mainly a
function of the temperature difference between the wall tempera-
while the overall bottom heat transfer coefficient is calculated ture and the saturation temperature which is dependant on the li-
assuming convective condensation as [27] quid film thickness. Increasing the mass flux has two opposing
effects on the top heat transfer coefficient. Firstly, it increases the
kl
hc ¼ 0:003Re0:74
l Pr 0:5
l fi : ð17Þ axial vapour shear perpendicular to the direction of the falling con-
d densate. This leads to a thinner film thickness and therefore higher
In general, there is good agreement between the experimental and heat transfer. On the other hand, increasing the mass flux sweeps
model results, with the maximum deviation of the bottom and top part of the liquid from the bottom stratum towards the top and

Fig. 7. Heat transfer coefficient without EHD forces for j experimental-bottom, N experimental-top, h model-bottom (Eq. (17)), M model-top (Eq. (16)).
4148 H. Sadek et al. / International Journal of Heat and Mass Transfer 53 (2010) 4141–4151

Fig. 8. The effect of mass flux on heat transfer, h 0 kV, M EHD-8 kV, and  ratio. Heat flux of q00 = 5.7 kW/m2 and average quality of xavg = 45%.

sides of the tube. This causes the liquid film at the top to thicken mass flux of 156 kg/m2 s. These results are consistent with the
and therefore lowers the top heat transfer. For mass flux less than dimensionless analysis which shows that the effect of EHD will
100 kg/m2 s, it is likely that increasing the mass flux, increases the be significant if the interfacial Masuda number (Md) is comparable
vapour shear, leading to thinner film thickness and therefore high- to the square of liquid Reynolds number ðRe2l Þ. The effect of mass
er heat transfer coefficients. For mass flux greater than 100 kg/ flux on ðMd=Re2l Þ and the heat transfer ratio is illustrated in
m2 s, the thickening of the film thickness due to the liquid being Fig. 9. The ratio ðMd=Re2l Þ is calculated for the liquid stratum based
swept from the bottom stratum towards the top and sides of the on the maximum interfacial electric field for the corresponding li-
tube offsets the thinning due to the vapour shear and therefore quid height. At mass flux = 57 kg/m2 s, the dimensionless ratio
the heat transfer remains nearly constant. For the heat transfer ðMd=Re2l Þ is approximately 0.7 which corresponds to a heat transfer
at the bottom, increasing the mass flux increases the heat transfer ratio of 2.8. Increasing the mass flux decreases the ratio ðMd=Re2l Þ
for the entire range tested. The convective heat transfer mecha- till it reaches a value of 0.04 with a corresponding heat transfer ra-
nism is a function of Reynolds number and the liquid film thick- tio of 1.2 at a mass flux of 156 kg/m2 s.
ness. Increasing the mass flux increases Reynolds number of the The experimental EHD top and bottom heat transfer coefficients
bottom stratum and decreases the height of the liquid stratum are compared to those predicted using models for dropwise con-
which increases heat transfer in this case. densation [28], film condensation and convective condensation
The heat transfer coefficient without EHD, with 8 kV DC applied [27] in Fig. 10. The experimental results are in best agreement with
voltage and the heat transfer ratio for different mass fluxes and the film condensation model results which suggests that the mode
fixed average quality of 45% are shown in Fig. 8. The heat transfer of heat transfer in this case is film condensation. There is some dis-
ratio, which is defined as the ratio of the heat transfer coefficient crepancy between the film condensation model predictions and
with applied voltage to that without applied voltage, is maximum the experimental values can be attributed to the additional EHD ef-
(h/ho = 2.8) at the lowest mass flux (G = 55 kg/m2 s) and decreases fects such as induced secondary motions within the liquid film,
with an increase of the mass flux till it reaches a value of 1.2 at a droplet interactions with the liquid film, and mixing/exchange of

Fig. 9. The effect of mass flux on } Md=Re2l , and M heat transfer ratio (8 kV). Heat flux of q00 = 5.7 kW/m2 and average quality of xavg = 45%.
H. Sadek et al. / International Journal of Heat and Mass Transfer 53 (2010) 4141–4151 4149

Fig. 10. Heat transfer coefficient for M 8 kV DC-top, h 8 kV DC-bottom, N–N film condensation (Eq. (16)), j – j convective condensation (Eq. (17)), –dropwise
condensation.

liquid between the vapour core and the liquid film. These addi- be estimated assuming film condensation, using the corresponding
tional EHD effects can increase heat transfer and is not accounted experimental heat transfer coefficients. The liquid extraction ratio
for in the model. Ex, defined as the percentage ratio of the area of the liquid in con-
Identifying the mode of condensation under the applied DC tact with the heat transfer surface without EHD force to that with
voltage allows for the estimation of the annular film thickness EHD force as
and therefore the amount of liquid extracted from the vicinity of
the heat transfer surface to the bulk of the flow due to the induced Al  AEHD
Ex ¼  100; ð20Þ
EHD forces. Without EHD, the bottom liquid stratum height can be Al
calculated from the geometric expression [27] Al ¼ Að1  aÞ; ð21Þ
  
2p  hstart ph 2 i
hL ¼ 0:5D 1  cos : ð18Þ AEHD ¼ D  ðD  2dEHD Þ2 ; ð22Þ
2 4

The top liquid film thickness can be evaluated using the film con- where Al, AEHD and dEHD are the area occupied with the liquid in con-
densation model which assumes that the heat transfer across the li- tact with the heat transfer surface area with and without EHD and
quid film is due to conduction only, therefore the liquid film thickness with EHD respectively.
The top and bottom heat transfer coefficients without EHD
kl
d¼ ; ð19Þ (0 kV) and with 8 kV DC applied voltage and the corresponding
h extraction ratio for different mass fluxes and fixed average quality
where h, in this case, is the experimental top heat transfer coeffi- of 45% are presented in Figs. 11 and 12, respectively. With EHD, the
cient. With EHD, both top and bottom liquid film thicknesses can top and the bottom heat transfer coefficients are almost equal for

Fig. 11. The effect of mass flux on the overall heat transfer coefficient. h 0 kV-top, j 0 kV-bottom, M 8 kV-top, N 8 kV-bottom. Heat flux of q00 = 5.7 kW/m2 and xavg = 45%.
4150 H. Sadek et al. / International Journal of Heat and Mass Transfer 53 (2010) 4141–4151

Fig. 12. The effect of mass flux on extraction ratio due to 8 kV applied voltage. Measurements taken for a constant heat flux of q00 = 9.6 kW/m2 and xavg = 45%.

the entire range of mass flux. This indicates that the applied volt- the liquid–vapour interface. The EHD forces increase the interfacial
age minimizes the effect of the gravitational forces and causes instabilities and cause liquid to be extracted away from the heat
the heat transfer to be more uniform circumferentially. At low transfer surfaces towards the vapour core. Liquid extraction causes
mass fluxes, both the top and the bottom heat transfer coefficients an increase in heat transfer due to the change in the mode of heat
at 8 kV are higher than those at 0 kV. Before the application of the transfer at the bottom of the tube from convective condensation
high voltage, the film thickness is 1.55 mm and 0.03 mm for the into film condensation. At mass flux of 55 kg/m2 s, the induced
bottom stratum and the top and side liquid condensate respec- electric forces caused liquid extraction of approximately 94% and
tively. After applying the high voltage, the film thickness reduces which corresponds to heat transfer enfacement of factor 2.8.
to 0.01 mm along the entire circumference. Applying 8 kV induces Increasing the mass flux, decreases the ratio Md=Re2l and therefore
EHD interfacial forces which cause liquid extraction of almost 94% decreases the extraction ratio to 87% and heat transfer enhance-
from both the bottom liquid stratum and the top and side film con- ment ratio at mass flux 156 kg/m2 s and 1.2, respectively.
densate, Fig. 12. This diminishes the thermal resistances for both
the bottom stratum and the film condensate at the top and sides
of the tube and therefore increases heat transfer. Increasing the References
mass flux at 8 kV DC applied voltage decreases the ratio
[1] J.S. Chang, A. Watson, Electromagnetic hydrodynamics, IEEE Trans. Dielect.
ðMd=Re2l Þ which decreases the effect of EHD on liquid extraction Electr. Insul. 1 (5) (1994) 871–895.
and therefore decreases the top and bottom heat transfer coeffi- [2] A. Yabe, Y. Mori, K. Hijikata, Active heat transfer enhancement by utilizing
cients. The liquid extraction decreases from 94% at the mass flux electric fields, Ann. Rev. Heat Transfer 7 (1995) 193–244.
[3] I.W. Eames, H.M. Sabir, Potential benefits of electrohydrodynamic
55 kg/m2 s to 87% at mass flux 156 kg/m2 s. For mass flux greater enhancement of two-phase heat transfer in the design of refrigeration
than approximately 130 kg/m2 s, the effect of EHD on the top heat systems, Appl. Therm. Eng. 17 (1) (1997) 79–92.
transfer is negligible and cause enhancement only for the bottom [4] S. Laohalertdecha, P. Naphon, S. Wongwises, A review of electrohydrodynamic
enhancement of heat transfer, Renew. Sustain. Energ. Rev. 11 (5) (2007) 858–
heat transfer. This can be attributed to the higher ðMd=Re2l Þ interfa- 876.
cial EHD forces for the bottom liquid stratum compared to the top [5] J.R. Melcher, Continuum Electromechanics, MIT Press, 1981.
and sides liquid film. The closer the interface to the high voltage [6] V.M. Budov, V.A. Kir’yanov, I.A. Shemagin, Heat transfer in the laminar-wave
section of condensation of a stationary vapor, J. Eng. Phys. 52 (6) (1987) 647–
electrode, the higher the values of the interfacial electric field
649.
intensity and therefore the induced interfacial EHD forces and [7] V. Penev, S.B. Krylov, C.H. Boyadjiev, V.P. Vorotilin, Wavy flow of thin liquid
interfacial Md number. film, Int. J. Heat Mass Transfer 15 (1968) 1389–1406.
[8] J.S. Chang, Stratified gas–liquid two-phase electrohydrodynamics in horizontal
pipe flow, IEEE Trans. Indust. Appl. (1989) 241–247.
5. Conclusions [9] A. Yabe, Active heat transfer enhancement by applying electric fields, Proc.
ASME/JSME Therm. Eng. Joint Conf. 3 (1991) 15–23.
[10] K. Yamashita, M. Kumagai, S. Sekita, A. Yabe, T. Taketani, K. Kikuchi, Heat
Experiments were performed in a 30 cm long horizontal, single- transfer characteristics on an EHD condenser, in: 3rd ASME/JSME Joint
pass, counter-current heat exchanger to investigate the variation in Thermal Engineering Conference, vol. 3, 1991, pp. 61–67.
[11] K. Sunada, A. Yabe, T. Taketani, Y. Yoshizawa, Experimental study of EHD
the heat transfer mode during tube-side convective condensation pseudo-dropwise condensation, ASME–JSME Therm. Eng. Proc. 3 (1991) 47–
due to the application of a high DC voltage. The experimental re- 53.
sults were compared to the predictions of the conventional heat [12] J. Ogata, A. Yabe, Augmentation of boiling heat transfer by utilizing the EHD
effect-EHD behaviour of boiling bubbles and heat transfer characteristics, Int. J.
transfer models for mass flux in the range of 45–156 kg/m2s and
Heat Mass Transfer 36 (3) (1993) 783–791.
average quality of 45%. With no applied voltage, the heat transfer [13] W.K.H. Panofsky, M. Phillips, Classical Electricity and Magnetism, second ed.,
mechanisms for stratified/stratified wavy flow are film condensa- Addison-Wesley Pub. Co., 1962.
tion for the falling film at the top section of the tube and convective [14] J. Cotton, A.J. Robinson, M. Shoukri, J.S. Chang, A two-phase flow pattern map
for annular channels under a dc applied voltage and the application to
condensation for the bottom liquid stratum. Applying a DC voltage electrohydrodynamic convective boiling analysis, Int. J. Heat Mass Transfer 48
induces electrohydrodynamic (EHD) forces which primarily act at (25–26) (2005) 5536–5579.
H. Sadek et al. / International Journal of Heat and Mass Transfer 53 (2010) 4141–4151 4151

[15] J.S. Cotton, D. Brocilo, J.S. Chang, M. Shoukri, T.S. Pollard, Numerical simulation [22] Y. Feng, J. Seyed-Yagoobi, Mechanism of annular two-phase flow heat transfer
of electric field distributions in electrohydrodynamic two-phase flow regimes, enhancement and pressure drop penalty in the presence of a radial electric
IEEE Trans. Dielect. Electr. Insul. 10 (1) (2003) 37–51. field – turbulence analysis, J. Heat Transfer 125 (3) (2003) 478–486.
[16] A. Yabe, K. Kikuchi, T. Taketani, Y. Mori, K. Hijikata, Augmentation of [23] H. Sadek, A.J. Robinson, J.S. Cotton, C.Y. Ching, M. Shoukri,
condensation heat transfer by applying non-uniform electric fields, Proc. Int. Electrohydrodynamic enhancement of in-tube convective condensation heat
Heat Transfer Conf. 5 (1982) 189–194. transfer, Int. J. Heat Mass Transfer 49 (9–10) (2006) 1647–1657.
[17] IEEE-DEIS-EHD Technical Committee, Recommended international standards [24] S. Laohalertdecha, S. Wongwises, Effect of EHD on heat transfer enhancement
for dimensionless parameters used in electrohydrodynamics, IEEE Trans. during two-phase condensation of R-134a at high mass flux in a horizontal
Dielect. Electr. Insul. 10 (1) (2003) 3–6. smooth tube, Heat Mass Transfer 43 (9) (2007) 871–880.
[18] P.H.G. Allen, T.G. Karayiannis, Electrohydrodynamic enhancement of heat [25] J. El-Hajal, J.R. Thome, A. Cavallini, Condensation in horizontal tubes, part 1:
transfer and fluid flow, Heat Recov. Syst. CHP 15 (5) (1995) 389–423. two-phase flow pattern map, Int. J. Heat Mass Transfer 46 (18) (2003) 3349–
[19] J. Seyed-Yagoobi, J.E. Bryan, Enhancement of heat transfer and mass transport 3363.
in single-phase and two-phase flows with electrohydrodynamics, Adv. Heat [26] S.J. Kline, F.A. McClintock, Describing uncertainties in single-sample
Transfer 33 (1999) 95–186. experiments, Mech. Eng. 75 (1) (1953) 3–8.
[20] A. Singh, M.M. Ohadi, S. Dessiatoun, EHD enhancement of in-tube [27] J.R. Thome, J. El-Hajal, A. Cavallini, Condensation in horizontal tubes, part 2:
condensation heat transfer of alternate refrigerant R-134a in smooth and new heat transfer model based on flow regimes, Int. J. Heat Mass Transfer 46
microfin tubes, ASHRAE Trans. 103 (1) (1997) 813–823. (18) (2003) 3365–3387.
[21] A. Gidwani, M. Molki, M.M. Ohadi, EHD-enhanced condensation of alternative [28] P. Griffith, Dropwise Condensation, Heat Exchanger Design Handbook,
refrigerants in smooth and corrugated tubes, HVAC&R Res. 8 (3) (2002) 219– Hemisphere Publishing, New York, 1983 (Chapter 2).
238.

You might also like