Sunteți pe pagina 1din 12

International Journal of Heat and Mass Transfer 107 (2017) 307318

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


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

Film condensation of steam flowing on a hydrophobic surface


D. Del Col a,, R. Parin a, A. Bisetto b, S. Bortolin a, A. Martucci a
a
b

Dipartimento di Ingegneria Industriale, Universit degli Studi di Padova, Via Venezia 1, 35131 Padova, Italy
Riello S.p.A., Via Mussa 20, 35017 Piombino Dese (PD), Italy

a r t i c l e

i n f o

Article history:
Received 29 December 2015
Received in revised form 27 October 2016
Accepted 27 October 2016

Keywords:
Flow condensation
Hydrophobic surface
Steam
Hydrophilic surface
Aluminum substrate

a b s t r a c t
Nano-engineered surfaces have recently been studied as a promising solution for several heat transfer
applications. In particular, the modification of surface wetting properties for condensation heat transfer
is an extremely interesting field of research. In the present work, an aluminum substrate has been modified to obtain a hydrophobic surface and the influence of the wetting properties during condensation of
pure steam on a vertical surface is investigated. Condensation tests, at heat flux between 250 and
500 kW m2 and vapor velocity between 2.2 m s1 and 6.4 m s1, have been performed on multiple samples, both hydrophilic (advancing contact angle <90) and hydrophobic samples, with advancing contact
angle 140. The condensation mode during the present test runs is purely filmwise, even on the
hydrophobic surfaces, due to the complete flooding of the surface. The results of filmwise condensation
on the hydrophobic surfaces display higher heat transfer coefficients compared to the untreated hydrophilic plate.
2016 Elsevier Ltd. All rights reserved.

1. Introduction
Owing to their water repellency properties, hydrophobic and
superhydrophobic surfaces have been studied as a promising solution to several challenges, such as drag reduction, anti-icing and
enhancement of two-phase heat transfer performance [13]. These
surfaces are also widely studied to promote dropwise condensation and thus enhance condensation heat transfer. In fact, condensation is present in many industrial processes, such as power
industry. With the motivation for improved energy efficiency and
miniaturization of heat exchangers, an extensive research has been
undertaken in the area of enhanced condensation heat transfer
from the report by Schmidt et al. [4] where the dropwise condensation was studied for the first time and this research nowadays is
powered by the new potential of surfaces with modified
wettability.
However film condensation, not dropwise, is the most common
condensation mode in condensers for industry application and
thus it is of great importance to understand how hydrophobicity
may affect heat transfer during filmwise condensation.
Surface wettability is defined using the contact angles of a
water drop sitting over it. For a static drop, equilibrium contact
angle is taken into account, while, for moving drops, the advancing
and receding contact angles are taken as the reference. The differ Corresponding author.
E-mail address: davide.delcol@unipd.it (D. Del Col).
http://dx.doi.org/10.1016/j.ijheatmasstransfer.2016.10.092
0017-9310/ 2016 Elsevier Ltd. All rights reserved.

ence among advancing and receding angles gives the contact angle
hysteresis. Surfaces presenting contact angles lower than 90 are
called hydrophilic, while those having h > 90 are named
hydrophobic. Moreover, if the surface presents extremely high contact angles, i.e. greater than 150, and extremely low contact angle
hysteresis, i.e. lower than 10, it is named superhydrophobic.
Hydrophobic surfaces are produced by lowering surface free
energy of the substrate, for example by coating with a thin layer
of a material having low surface energy, such as organic
substances, polymers, and noble metals [59]. The use of organic
substances as low-surface energy promoters requires strong,
long-term adhesion forces between the coating and the metal substrate. Usually, the thicker is the coating, the better its resistance to
corrosion/erosion. However, due to very low thermal conductivity,
thick organic coatings add a heat transfer resistance that may deteriorate the condensation performance. Moreover, the coating
material, if inadvertently removed from the condenser surface,
may contaminate the system.
The correlation between surface wetting properties and condensation heat transfer enhancement is still not well established.
As an example, Xuehu et al. [10] reported heat transfer measurements on a superhydrophobic nanostructured copper sample and
compared them to those obtained on a mirror polished hydrophobic specimen. They found that the nanostructured substrate does
not improve the condensation heat transfer performance as
expected from the higher contact angle, but better results were
achieved with the hydrophobic substrate. Flow condensation tests

308

D. Del Col et al. / International Journal of Heat and Mass Transfer 107 (2017) 307318

Nomenclature
(dp/dz)F
A
c
Dh
g
G
h
Hlv
HTC
L
_
m
p
Q
q
S
T
T0
T00
uA
uB
uC
uM
uTOT
z
z1
z2

frictional pressure gradient, Pa m1


heat transfer area, m2
specific heat capacity, J kg1 K1
hydraulic diameter, m
gravitational acceleration, m s2
specific mass flow rate, kg m2 s1
specific enthalpy, J kg1
latent heat of vaporization, J kg1
heat transfer coefficient, W m2 K1
length, m
mass flow rate, kg s1
pressure, Pa
heat flow rate, W
heat flux, W m2
cross section area, m2
temperature, K
specimen temperature at z1, K
specimen temperature at z2, K
type-A uncertainty, %
type-B uncertainty, %
combined uncertainty, %
expanded uncertainty, %
global uncertainty, %
orthogonal axis of the sample, m
position (1.5 mm) along z, m
position (3.25 mm) along z, m

Greek symbols
b
slip length, m
C
condensate mass flow rate per unit width, kg m1 s1

of saturated vapor on superhydrophobic nanotextured copper surfaces presented by Torresin et al. [11] show that condensing drops
form and penetrate into the surface texture, with a reduction of
their mobility. However, the authors found that the vapor shearinduced roll-off of droplets compensates for the reduced drops
mobility and enhances the overall thermal transport with respect
to non-treated copper surfaces.
Works cited above refer to dropwise condensation mode and
from its discovery several authors [1214] managed to promote
it over hydrophobic or superhydrophobic surfaces. Instead, very
limited studies have addressed the phenomenon of film condensation over a hydrophobic surface, which is the actual subject of the
present investigation.
1.1. Literature review
Some works in the literature studied vapor condensation inside
micro and minichannel with different surface wettability. In Fang
et al. [15] flow condensation inside hydrophilic, hydrophobic and
semi-hydrophobic microchannel is presented. The authors show
some images of the condensation flow patterns, where the wall
is wetted by drops or liquid film depending on the surface wettability and on the position along the microchannel. In Ref. [8] vapor
condensation is studied in minichannels. Hence, the heat transfer
coefficient does not depend on vapor quality, in fact for each mass
flow rate, the heat transfer coefficients at low vapor quality
(x = 0.2) and at high vapor quality (x = 0.9) are comparable. It is
interesting to notice that at very low vapor quality, high HTCs
are measured on hydrophobic surfaces during vapor condensation;
this may occur in filmwise mode although this is not clearly stated
in the paper.

d
Dh
DT
DTml
h
k

l
q
si

liquid film thickness, m


contact angle hysteresis,
temperature difference, K
mean logarithmic temperature difference, K
contact angle,
thermal conductivity, W m1 K1
dynamic viscosity, Pa s
density, kg m3
interfacial shear stress, Pa

Subscripts
adv
advancing
BC
boiling chamber
CALC
calculated
cool
cooling side
EXP
experimental
HPHIL
hydrophilic
HY
hydrophobic
IN
inlet
l
liquid
m
mean
Nu
Nusselt component
OUT
outlet
rec
receding
SAT
saturation
SS
shear stress component
V
vapor
WALL
surface

Some papers in the literature investigate the effect of the


hydrophobicity during single phase flow and heat transfer
[1,16,17]. In these works the heat transfer coefficients measured
over hydrophilic and hydrophobic surfaces were compared, showing a relationship between the surface wetting properties and the
heat transfer and flow parameters. Two main results have been
achieved over the hydrophobic surfaces: a drag reduction and a
heat transfer decrease. For example in Ref. [16] the hydrophilic
microchannel consistently shows a higher heat transfer coefficient
than that of the hydrophobic microchannel although this increase
is only about 8%; in Ref. [17] the heat transfer coefficients of superhydrophobic tubes display a small decrease as compared to
untreated tubes, depending on the tube diameter. In the same
work [17], the drag reduction ranged from 8.3% to 17.8% when
moving from untreated to superhydrophobic tubes. The explanation of these results has been attributed to the existence of a slip
velocity at the wall, which was, moreover, measured during those
experimental tests. The most validated theory about the slip
boundary explains that air cavities can be entrapped in between
the surface roughness and, since the dynamic viscosity of air is
remarkably smaller than that of liquid, large velocity gradients
near the hydrophobic wall are achieved and, thus, the frictional
flow resistance is reduced [1,18,19].
Normally, studies on film condensation heat transfer essentially
comply with no-slip boundary conditions at the solid surface. In
fact, for macroscopic flows of simple fluids, the slip length is usually so small (b is equal to few nanometers) that it can be
neglected, and the no-slip boundary condition can be used without
loss of accuracy [20]. However, some literature shows the existence of a slip velocity when a liquid flows on hydrophobic or
superhydrophobic treated surfaces [1,1619,2124]. Recent exper-

309

D. Del Col et al. / International Journal of Heat and Mass Transfer 107 (2017) 307318

imental and computation work indicates that slip boundary


increases with an increase of Reynolds number and a reduction
of the viscous sub-layer thickness, but a more detailed understanding of the physical nature of this scaling has yet to be presented
[1,19,23].
Pati et al. [25] theoretically studied how the slip velocity can
influence the condensation heat exchange over vertical surfaces.
The authors studied the relationship between slippage phenomenon and two-phase heat transfer performance by introducing
a nonzero slip length, defined as the distance behind the solid
interface at which the liquid velocity extrapolates to zero, into
the classical condensation theory. In [25] the authors demonstrated that when promoting a slip length b 0 the condensate
film thickness over the surface reduces, in comparison to classical
theory (b = 0), the more the higher is the slip length. This is because
an increase of b is associated with an augmentation of the slip
velocity at the surface, resulting in a thinner condensate layer.
The reduction of the liquid film thickness is associated with a
reduced thermal resistance of the condensate layer, thus it leads
to an enhancement of the two-phase heat transfer coefficient. Similar results are shown in Ref. [26], where variation of the velocity
along the liquid layer, of the condensate film thickness and of
the condensation mass flow rate per unit width are reported for
different slip lengths. The authors showed that a non-zero slip
length reflects on a non-zero slip velocity at the solid-liquid interface, leading to reduced condensate film thickness and enhanced
condensation mass flow rate per unit width.
However, to the best of the present authors knowledge, no
experimental analysis about film condensation over hydrophobic
surfaces has been presented in the literature and the effect of
hydrophobicity on the liquid film and the heat transfer coefficient
has never been experimentally investigated.
The present paper presents an investigation of vapor condensation on hydrophobic aluminum substrates being aluminum a
material of great interest for many industrial fields, especially in
the heat transfer area.
First, a technique for achieving hydrophobic properties over
aluminum substrates is introduced, and surface properties characterization, by means of contact angles analysis, Scanning Electron
Microscopy (SEM) and Atomic Force Microscopy (AFM), is
presented.
Then, the experimental investigation of film condensation over
hydrophobic surfaces is addressed. Heat transfer data are presented and compared against values obtained during condensation
over standard hydrophilic aluminum substrates.
2. Surface wetting properties modification
The effect of the wetting properties of the metallic surface on
the heat transfer coefficient during pure steam condensation has
been investigated by modifying the wettability of the aluminum
substrates through a proper chemical process, in order to obtain
hydrophobic characteristics. In this Section, the method used to
impart low wetting properties and the surface characterization
are presented.

2.1. Methods
A high purity (AW 1050, minimum Al quantity 99.50%) aluminum smooth plate was used for specimen fabrication. Hexane
(anhydrous, P99%) and 1H,1H,2H,2H-Perfluorooctyltriethoxysi
lane (98%) (FOTS) were purchased from Sigma Aldrich.
Five samples have been prepared for the condensation tests.
Their main characteristics are summarized in Table 1. Two tested
surfaces were hydrophilic, while three samples were hydrophobic.
The HPHIL #1 sample has not undergone any treatment: it was
tested in the apparatus as it was provided. For the HPHIL #2 surface, the as-received aluminum substrate has been sanded using
emery paper #1200 and then washed in distilled water. The three
hydrophobic surfaces (HY #1, HY #2 and HY #3) have all been
obtained with the same procedure: the samples were sanded using
emery paper #1200 and then were immersed for 15 min into Isopropanol (IPA), continuously stirring the solution through a sonication probe, and dried into a nitrogen stream, for cleaning purposes.
Then a low surface tension coating was deposited on the clean substrate. The process consisted in spin-coating a HexaneFOTS mixture (5% by volume) onto the sample at 800 rpm for 30 s. After
spin coating, the sample was baked at 150 C for 30 min for final
solvent evaporation and film stabilization.
2.2. Characterization
The morphology of the surface has been investigated by Scanning Electron Microscopy (SEM). In Fig. 1 it is possible to see the
morphology of the aluminum substrate before and after the deposition of the hydrophobic coating. The deposited coating is
homogenous without any crack. A film thickness of 100 nm has
been measured by a profilometer on a scratch made just after
the film deposition. Surface wetting properties and morphological
characteristics were analyzed by means of contact angles analysis.
Contact angles are measured using the standard sessile drop
method, recording a water drop (backlight illuminated by a LED
light) expanding and contracting quasi-statically over the horizontally oriented surface of interest. Advancing (hadv) and receding
(hrec) contact angles are evaluated by post-processing the videos
and fitting with a circle the drop profile near the contact point.
Contact angle hysteresis (Dh) is calculated as the difference
between hadv and hrec. The values of the contact angles reported
in this paper are the average between at least five measurements
taken at different positions over the surface; in addition the standard deviation is reported together with the corresponding mean
value.
Fig. 2 shows the advancing and receding contact angles measured over the untreated and the hydrophobic substrates.
Prior to any treatment, the as-received aluminum sample presents an advancing contact angle equal to 84.8 4.4, a receding
contact angle equal to 17.4 4.6 and a contact angle hysteresis
equal to 67.3 4. After the use of emery paper, the advancing
and the receding contact angle decreases and they are
hadv 31:6  4:4 and hrec 6:4  1:2. These surfaces, according to
the usual classification, can be named as hydrophilic and they

Table 1
Advancing and receding contact angles and contact angle hysteresis of the tested aluminum surfaces.
Mean value

HPHIL #1
HPHIL #2
HY #1
HY #2
HY #3

Standard deviation

hadv []

hrec []

Dh []

hadv []

hrec []

Dh []

84.8
31.6
143.5
127.0
133.1

17.4
6.4
43
24.8
37.9

67.3
25.2
100.5
102.2
95.2

4.4
4.4
2.6
4.0
4.2

4.6
1.2
4.8
4.4
3.9

4
4.6
3.4
6.9
5.7

310

D. Del Col et al. / International Journal of Heat and Mass Transfer 107 (2017) 307318

AFM analysis has also been done over a 2  2 cm sample


obtained with the same treatment of the HY #1, HY #2 and HY
#3 in order to measure the surface roughness; an image of AFM
on the hydrophobic surface is shown in Fig. 3. The analysis has
been done on areas with dimensions of 50  50 lm, where a mean
surface roughness was measured equal to 507 nm, with a minimum value of 187 nm and a maximum value of 917 nm. The surface roughness measured with the AFM may provide an
explanation of the high advancing contact angle measured over
the hydrophobic substrates.

(a)

2 m

3. Experimental setup
3.1. Two-phase flow loop

(b)

2 m
Fig. 1. SEM images of: a) flat polished aluminum and b) hydrophobic surface.

are listed as HPHIL #1 and HPHIL #2 in Table 1. After the treatment, the functionalized hydrophobic sample is characterized by
a mean advancing contact angle equal to 135, a mean receding
contact angle equal to 35 and a mean Dh = 99. Such surfaces
are usually considered as hydrophobic and they are listed as HY
#1, HY #2 and HY #3 in Table 1.

Advancing Contact Angle

Receding Contact Angle

HYDROPHOBIC

UNTREATED

Sample

A schematic diagram of the new two-phase flow loop for condensation measurements is shown in Fig. 4.
The system consists of four main components: the boiling
chamber, the test section, the cooling water loop and the postcondenser. Steam is generated in a cylindrical stainless steel boiling chamber which is connected to the test section by a stainless
steel vapor line. The water vaporization is promoted inside the
chamber by means of four electrical heaters with a total power
of 4 kW. The electrical power supplied to the heaters is measured
using a power analyzer NORMA 4000. The pipe connections
between the boiler and the test section are thermally insulated
and heated by means of a resistance wire installed around the pipe,
to avoid formation of condensate before the entrance of the test
section (wall temperature is checked through a T-type thermocouple). The steam enters the test section in saturated conditions with
vapor quality equal to one. In the test section, steam is partially
condensed over the test aluminum surface and the latent heat is
removed by the cold water coming from a thermostatic bath. The
coolant inlet temperature is measured by a T-type thermocouple,
while the coolant temperature difference between the inlet and
the outlet is measured by means of a three-junction copperconstantan thermopile. The coolant mass flow rate is measured

Fig. 2. Examples of advancing and receding contact angles for the hydrophilic and the hydrophobic specimens.

D. Del Col et al. / International Journal of Heat and Mass Transfer 107 (2017) 307318

3.7 m

0
0

50 m
Fig. 3. AFM image of a hydrophobic sample (scan area 50 x 50 lm).

311

by means of a Coriolis effect mass flow meter. The vapor pressure


and temperature are gauged at the inlet of the measuring section
by means of a differential pressure transducer (coupled with an
absolute one for ambient pressure evaluation) and a T-type thermocouple, respectively. Downstream the test section, the twophase mixture passes through a secondary water condenser where
the condensation is completed and the liquid subcooled. The subcooled liquid returns to the boiler driven by the difference between
liquid and vapor density and it closes the loop. The cooling water
temperature is measured at the inlet and the outlet of the postcondenser by means of T-type thermocouples while its mass flow
rate is measured using a magnetic flow meter.
To regulate the system pressure, a hydraulic accumulator is
installed in the liquid line downstream the post-condenser. A needle valve, placed before the boiling chamber, is used to regulate the
liquid flow in the main loop and it allows to achieve stable conditions during the tests. In between the hydraulic accumulator and
the needle valve, a mechanical filter and a liquid indicator are
installed. Before entering the boiling chamber the temperature of
the subcooled liquid is measured by means of a T-type thermocouple. All the components of the test rig, with the exception of the
test section, are made of stainless steel in order to avoid contami-

Fig. 4. Schematic of the experimental thermosyphon loop for condensation tests. P = Pressure transducer, T = Thermocouple, dT = Thermopile, CFM = Coriolis mass Flow
Meter, MFM = Magnetic mass Flow Meter, MF = Mechanical Filter.

312

D. Del Col et al. / International Journal of Heat and Mass Transfer 107 (2017) 307318

nation of the fluid. The boiling chamber and the stainless steel lines
are thermally insulated to prevent heat losses to the ambient.
Since even a few concentration of non-condensable gases (NCG)
in the vapor could lead to a huge decrease in the thermal performance of the condensation process, several actions have been
undertaken to avoid it. First of all, the test rig is always operated
in overpressure to prevent air infiltration inside the test loop. Furthermore, before each test run, the whole system is vacuumed;
then the test rig is charged with deionized (DI) water coming from
the supplying tank and pumped inside the test rig by a centrifugal
pump. When the pressure inside the setup becomes higher than
the ambient pressure, some water is released from the top valve
while the refill pump is continuing to run, in order to get rid of
the non-condensable gases still present inside the setup. After
one minute of free water discharging, the top valve and the filling
line are closed. Subsequently the boiler is started: when the water
reaches the saturation temperature and boils for several minutes,
the vapor is released from the top of the system and from a discharge valve located in the upper part of the post-condenser. This
procedure is repeated several times in order to get rid of the gases
dissolved in the water. In all these operations, particular care is
observed to keep always the test rig pressurized. The hydraulic
accumulator is used to maintain the system in overpressure overnight, thereby avoiding the need for DI water refilling every day.

at 1.5 mm and two at 3.25 mm depth below the front surface of the
sample.
The sample is vertically mounted and is located in between the
two main parts of the test section: one contains the steam channel
(in contact with the frontal face of the specimen) and the other
contains the cooling channel (in contact with the back face of the
specimen). The cooling path consists of a 20 mm  5 mm rectangular PEEK channel and it is used to remove the heat from the condensation process. The cooling water flows in countercurrent to
the steam direction inside the test section. In between the vapor
block and the cooling system block, two frames are located for
thermocouples and gaskets positioning.

3.2. Test section

The test section (Fig. 5) is designed for the measurement of the


heat transfer coefficient over a metallic surface and for the simultaneous visualization of the condensation process. It consists of
several pieces realized both in PEEK and glass. A rectangular channel 160 mm long (cross section 30 mm  5 mm) was grooved into
a PEEK block and the vapor, coming from the boiling chamber,
flows inside it. The PEEK channel is fitted with a rectangular aluminum plate (vertically oriented), over which condensation occurs.
One side of the channel is covered by a double glass (with an air
chamber for thermal insulation) to allow the visualization of the
process. In addition, the frontal glass is heated up by means of an
electrical heater to avoid vapor condensation over it. The other side
of the channel, opposite to the glass, is machined for accommodating the metallic substrate.
The metallic specimen has 10 mm thickness and the condensation surface is 50 mm high and 20 mm wide. The specimen is fitted
with four T-type thermocouples located at the inlet and at the outlet of the condensing surface, inside four 0.7 mm holes, two located

where A is the heat transfer area of the specimen (50  20 mm).


Since the wall temperatures are measured with the thermocouples placed in the aluminum (1.5 mm and 3.25 mm deep), the surface temperatures can be obtained from the measured values with
the hypothesis of one-dimensional temperature distribution, since
the PEEK thermal conductivity is much lower in comparison to the
aluminum sample. At the inlet and outlet of the aluminum sample
the two surface temperatures are calculated as:

(b)

VAPOR
OUTLET

WATER
INLET

where ccool is the specific heat capacity of the coolant evaluated at


the mean temperature between the inlet and the outlet.
Hence, the condensation heat flux q is given by

Q cool
A

T WALL;IN T 0IN T 0IN  T 00IN

z1
z2  z1

T WALL;OUT T 0OUT T 0OUT  T 00OUT

z1
z2  z1

where T0 are the inlet and outlet temperatures measured in the aluminum sample at z1 = 1.5 mm and T00 are the inlet and outlet temperatures measured in the aluminum sample at z2 = 3.25 mm
from the condensing surface.

(c)

T''IN

T''OUT

T'IN

WATER

WATER
OUTLET

_ cool ccool DT cool


Q cool m

METALLIC
SAMPLE

GLASS

(a)

_ cool is directly measured, as


Since the coolant mass flow rate m
well as the inlet coolant temperature TIN-cool and the coolant temperature difference DTcool, the heat flow rate extracted from the
condensing vapor can be obtained as

VAPOR

VAPOR
INLET

3.3. Data reduction

T'OUT

Fig. 5. Test section: a) Sketch of the assembled component. b) View of a cross section. c) Metallic specimen.

313

D. Del Col et al. / International Journal of Heat and Mass Transfer 107 (2017) 307318

Finally, the condensation heat transfer coefficient is evaluated


as:

q
DT ml

where DTml is the mean logarithmic temperature difference


between the surface and the steam. The steam saturation temperature is obtained from the pressure measurement at the inlet of the
test section.
_ steam is obtained from Eq. (6), where
The steam mass flow rate m
QBC is the heat flow rate given to the boiling chamber (measured
from the electrical power supplied to the heaters), besides hV and
hIN,BC are the enthalpy values of the saturated steam and of the subcooled liquid at the entrance of the boiling chamber, respectively.

_ steam
m

Q BC
hv  hIN;BC

All the thermodynamic and transport properties of the steam are


evaluated by means of NIST Refprop Version 9.1 [27] from temperature and pressure measurements.
Once the mass flow rate is known, the mass velocity of the
vapor flowing inside the test section can be calculated as

Gsteam

_ steam
m
S

where S is the cross section area of the channel. By acting on the


power supplied to the boiling chamber, it is possible to regulate
the mass flow rate of the steam inside the system and to perform
tests at varying mass velocity.
3.4. Uncertainty analysis
The experimental uncertainties are calculated following the
general rules reported in ISO Guide to the Expression of Uncertainty in Measurement [28]. For each parameter, the combined
uncertainty uc is calculated considering Type A and Type B
components.
Type A uncertainty is equal to the standard deviation of the
measured value while Type B uncertainty is related to the instrument properties. Table 2 reports Type B uncertainty of the most
relevant measured parameters. The uncertainty of the thermocouples position is assumed to be negligible.
The global uncertainty of the measured data xi is evaluated as

q
uTOT xi u2A xi u2B xi

4. Experimental analysis
4.1. Results over untreated hydrophilic surface: validation of
experimental procedure vs. literature
The condensation heat transfer was first studied on an aluminum hydrophilic surface (HPHIL #1) presenting advancing and
receding contact angles equal to hadv 84:8  4:4 , and
hrec = 17.4 4.6, respectively.
Tests were performed at G = 1.5 kg m2 s1, TSAT = 107 C and
_ cool = 0.11 kg s1, while the coolant inlet temperature was varied,
m
leading to changes in the temperature difference between saturation and wall. Fig. 6 reports the experimental values of the HTC
(Section 3.3) measured on the untreated hydrophilic surface. As
expected for film condensation, the mean heat transfer coefficient
decreases when the mean logarithmic temperature difference
increases.
These measurements have also been compared with correlations available in the literature for filmwise condensation to verify
the present experimental technique. When applying a model for
condensation driven only by gravity, one would expect to underestimate the experimental data because the effect of velocity would
be neglected. Therefore a more complete model, which combines
gravity and shear stress components, has been adopted [29]. The
global heat transfer coefficient can be then calculated by Eq. (10):

HTC CALC

q
HTC 2Nu HTC 2ss

10

where
"

HTC Nu 1:15 0:026

Hlv ll
kl T SAT  T wall

0:5

"

0:79 0:943

ql ql  qv gHlv k3l
ll T SAT  T wall L

#0:25

11

20

where uA(xi) and uB(xi) are respectively the Type A and Type B
uncertainties of the i-th parameter xi.
The combined standard uncertainty of the derived data
y f x1 ; x2 ;    ; xn is evaluated as

v
u n  2
uX df
uC y t
u2TOT xi
dxi
i1

The expanded uncertainty uM(xi) is finally obtained by multiplying


uC(xi) by a coverage factor k 2. In this analysis the uncertainties
of the thermodynamic properties evaluated using NIST Refprop Version 9.1 [27] are neglected.

15

HTC [kW m-2 K-1]

HTC

The results of the uncertainty analysis lead to a mean expanded


uncertainty of the logarithmic temperature difference between the
steam and the wall equal to 1.6%, a mean expanded uncertainty of
the heat flux removed in the measuring section equal to 7.8% and
a mean expanded uncertainty of the heat transfer coefficient equal
to 11.1%.

10

Experimental HTC
Serie3
Calc. HTC (gravity + SS)

UNTREATED
(HYDROPHILIC)

Serie2
Calc. HTC (gravity)
0

Table 2
Type B uncertainty of the measured parameters.
Temperature
Cooling water flow rate
Differential pressure
Absolute pressure
Electrical power

0.05 K


_
0.15% 0:1
_  100 % of the flow rate m
m
0.04%
0.1%
0.1%

10

15

20

25

30

35

Tml [C]
Fig. 6. Calculated and experimental heat transfer coefficient versus mean logarithmic temperature difference between saturated vapor and wall during condensation
on the untreated hydrophilic sample (HPHIL #1) at G = 1.5 kg m2 s1. The values
Calc. HTC (gravity) are obtained by means of Eq. (11). The values Calc. HTC
(gravity + SS) are determined using Eq. (10).

D. Del Col et al. / International Journal of Heat and Mass Transfer 107 (2017) 307318

Eq. (11) is referred to the gravity controlled condensation heat


transfer coefficient: it includes the Nusselt equation [30] together
with terms which account for wave formation and inertia effects.
Since waves formation was observed on the condensate film during
the experiments, the heat transfer enhancement has to be taken
into account introducing the correction factor 1.15 [31]. The inertia
effects, which are neglected in the Nusselts theory, are included by
the terms in between the square brackets, as suggested in Ref. [32].
The variation of the physical properties of the condensate with temperature must also be considered and this is done here by calculating all the physical properties of the condensate at the mean
temperature:

T m 0:75 T WALL 0:25 T SAT

deviation between experiments and calculations is equal to 4.5%,


far below the experimental uncertainty.

4.2. Investigation over hydrophobic surfaces


Condensation tests over the hydrophobic aluminum specimens
(HY #1) are hereinafter presented and compared against the values
obtained during condensation over the untreated hydrophilic sample (HPHIL #1). Condensation heat transfer measurements were
performed at about 106 C saturation temperature, which corresponds to a saturation pressure pSAT  1.25 bar. Experimental tests
were performed by varying the inlet temperature and the mass
flow rate of the cooling water, resulting in variable mean logarithmic temperature difference between the steam and the wall and
thus variable heat flux. Tests were performed varying the vapor
mass velocity by acting on the power of the boiling chamber. During tests, the heat flux was varied between 250 and 500 kW m2
and the steam velocity was varied between 2.2 and 6.4 m s1.
Thanks to visualization, it was possible to verify that filmwise
condensation occurred in all the tests, without presence of drops.
Fig. 7a reports an image of the steam condensing on the surface
taken with the high speed video camera Photron FASTCAM Mini
UX100 high speed camera (CMOS sensor with Bayer color filter
array) at 1280 by 606 pixel resolution and 2000 frames per second.
As it can be seen in Fig. 7a, there is no presence of drops on the condensing surface and the condensation mode is completely filmwise; no solid-liquid-vapor contact line can be detected. Other
images like the one in Fig. 7a have been taken during the tests,
and they all confirm that, at present operating conditions, filmwise
condensation always occurred. The reason why dropwise condensation is not observed at these operating conditions must be
related to the flooding mechanism over the surface. When a
hydrophobic-rough surface is subject to a sufficiently high condensation heat flux, thus high temperature difference between the
wall and the saturated vapor, the surface may be in a flooded
state and single droplets could not be observed. As outlined in
[35], hydrophobic and rough surfaces may be optimal to promote
dropwise condensation at low heat flux while at higher heat flux
smooth condensing surfaces with low contact angle hysteresis
have a better performance, avoiding the flooding mechanism
which would occur on a superhydrophobic surface. Enright et al.
[13,36] noted that 80 kW m2 is a limiting heat flux value in the
case of superhydrophobic surfaces. It is interesting to note that this

12

while the latent heat Hlv and the density of the saturated steam qv
are calculated at TSAT [33].
The second term under the square root in Eq. (10) is due to the
vapor velocity and it is expressed as follows

s
k2l si ql
HTC ss
2 Cl l l

13

The shear stress correlation used in the present work (Eq. (13)) is
the one proposed in [33] for the shear-controlled condensation process with laminar flow in the condensate film (the maximum Reynolds number of the condensate film is about 130 for the present
data).
As reported in Ref. [30], the condensate mass flow rate per unit
width in Eq. (13) can be evaluated from the liquid film thickness as

ql ql  qv gd3
3ll

14

and the shear stress at the liquid-vapor interface can be obtained


from the frictional pressure gradient:

si 

 
Dh dp
4 dz F

15

The model by Friedel [34] has been used in the present work to
compute the frictional pressure gradient in Eq. (15).
The calculated values of the heat transfer coefficients are presented in Fig. 6. As expected, if the vapor velocity is neglected
(HTC calculated using Eq. (11)), the calculated HTCs underpredict
the experimental values, whereas when using Eq. (10) the mean

(a)

(b)

160

35

HTC

140
120

TIN

Temperature [C]

100

30
25

TIN

20

80
15

60
TOUT

40
20

TOUT

10
5

Tin,ref

HTC [kW m-2 K-1]

314

2000

Time [s]

4000

6000

Fig. 7. a) Image of the filmwise condensation process over the sample HY #1. b) Wall temperatures measured by thermocouples within the specimen T 0IN ; T 00IN ; T 0OUT ; T 00OUT
plotted versus time (T0 and T00 are measured at 1.5 mm and 3.25 mm depth, respectively) during condensation over the HY #1 sample. The inlet water temperature (Tin,ref) and
the measured heat transfer coefficient (HTC) are also reported. Data refer to G = 5.1 kg m2 s1.

315

D. Del Col et al. / International Journal of Heat and Mass Transfer 107 (2017) 307318

limiting value is considerably lower than the experimental heat


flux in the present tests, being the minimum value around
250 kW m2. Therefore, the roughness of the present hydrophobic
surfaces, as measured by means of AFM, and the heat flux
exchanged during the tests suggest that the hydrophobic surfaces
must be in a flooded state, yielding a filmwise condensation mode.
The occurrence of a film condensation process can be also
checked by looking at the four temperatures measured by thermocouples within the specimen wall (T 0IN ; T 00IN ; T 0OUT ; T 00OUT , Section 3.3)
which are plotted vs. time in Fig. 7b. The inlet water temperature
(Tin,ref) and the measured HTC are also plotted in the same
graph, which refers to a test run at vapor mass flux equal to
5.1 kg m1 s1. A confirmation of the film condensation is provided
by the comparison between the wall temperatures at the inlet and
the wall temperatures at the outlet: during filmwise condensation,
one can expect lower wall temperatures at the outlet as compared
to the inlet, because of the liquid film thickness, while, under dropwise condensation, typically no difference between inlet and outlet
is expected. Instead, Fig. 7b displays a clear difference between T0 IN
and T0 OUT and between T00 IN and T00 OUT.
When the inlet water temperature Tin,ref decreases, the global
temperature difference between steam and cooling water
increases and the wall temperatures at inlet and outlet diverge

600

more, meaning that the thermal resistance between saturated


vapor and aluminum surface grows along the flow of the steam,
i.e. a condensate film is present and its thickness increases
between inlet and outlet when increasing the difference between
saturation and surface temperature.
The heat flux and heat transfer coefficient values measured during filmwise condensation over the hydrophobic treated aluminum
substrate are reported in Fig. 8, at three values of mass velocity
(1.53.15.1 kg m2s1). Trends are those expected during filmwise condensation. Heat flux increases both when increasing
steam mass velocity and saturation-to-wall temperature difference. Instead, the condensation heat transfer coefficient increases
with Gsteam, but it tends to decrease almost linearly when increasing DTml, with higher slope at higher steam velocity.
Fig. 9 shows a comparison between heat transfer coefficients
measured on the untreated and the hydrophobic surfaces, for different saturation-to-wall temperature difference and steam mass
velocity. The Figure shows that heat transfer coefficients on the
hydrophobic surface are higher by 1045% than those on the
untreated surface, for the present conditions. Two aspects can be
seen:
the ratio HTC HY =HTC HPHIL increases with the increase of Gsteam at
a fixed DT ml . The increase of the vapor mass flow rate entails the
increase of the condensate velocity;

30
25

400

HTC [kW m-2 K-1]

q [kW m-2]

500

300
200

-2 -1

G=1.5 kg m s

G=3.1 kg m-2 s-1

100

10

15

20

25

10
-2 -1

G=5.1 kg m s
0
0

15

G=1.5 kg m s

HYDROPHOBIC

-2 -1

20

30

Tml [K]

-2 -1

G=5.1 kg m s

HY

HY

HY

HPHIL

HPHIL

HPHIL

10

20
Tml [K]

(a)

30

40

(a)

35

1.5

30

1.4

25

HTCHY / HTCHPHIL [/]

HTC [kW m-2 K-1]

-2 -1

G=3.1 kg m s

20
15
-2 -1

G=1.5 kg m s

10

G=3.1 kg m-2 s-1


5

G=5.1 kg m-2 s-1

HYDROPHOBIC

10

15

20

25

1.2
-2s-1 s-1
1.5 kg
GG==1.5
kgmm-2
-2s-1 s-1
GG==3.1
kgmm-2
3.1 kg
-2s-1
5.1 kg
GG==5.1
kgmm-2
s-1

1.1
1

0
0

1.3

30

Tml [K]

(b)
Fig. 8. Heat flux (a) and heat transfer coefficient (b) versus mean logarithmic
temperature difference between saturated steam and surface during condensation
over the hydrophobic sample (HY #1). Data refer to different steam mass velocities
G (1.53.15.1 kg m2 s1).

10

20

30

Tml [K]

(b)
Fig. 9. Comparison between the heat transfer coefficient on the hydrophobic (HY
#1) and on the untreated (HPHIL #1) sample (a), and ratio between the two values
(b), as a function of wall subcooling degree, for different steam mass velocities G
[kg m2 s1].

316

D. Del Col et al. / International Journal of Heat and Mass Transfer 107 (2017) 307318

the ratio HTC HY =HTC HPHIL decreases with the increase of DT ml at a


fixed Gsteam. The increase of the mean logarithmic temperature
difference between the steam and the surface leads to the augmentation of the condensate film thickness.
4.3. Repeatability and discussion of results
To the best of the present authors knowledge, the results
shown in Fig. 9 are the first ones of this kind, i.e. the heat transfer
enhancement due to the surface hydrophobicity during filmwise
condensation has never been experimentally proved in the literature before. Several tests have been made to verify those results
and check the reliability of the present data. As previously
reported, beside HPHIL #1 and HY #1, more samples have been
prepared with the aim of collecting more data points during condensation over hydrophilic and hydrophobic surfaces. The test
samples and their values of contact angle measurements are listed
in Table 1. The experimental tests have been carried out keeping
the same vapor mass flow rate, saturation temperature and cooling
water mass flow rate as adopted in the tests of Fig. 9, while the
inlet cooling water temperature is varied to vary the mean logarithmic temperature difference between steam and wall. In
Fig. 10 the main results are summarized, in particular the HTCs
measured on the hydrophilic samples (HPHIL #1 and HPHIL #2)
and on the hydrophobic ones (HY #1, HY #2 and HY #3) are displayed. Tests on both hydrophilic and hydrophobic surfaces present a good repeatability, meaning that the phenomenon has
always the same behavior independently from the specimen and
it is related to the surface wettability. All the results plotted in
Fig. 10 are obtained on new samples to avoid the effect of surface
degradation. In fact, it was seen that the degradation of the
hydrophobic layer can severely affect the present results. In particular, after testing the surface for a certain period, which can vary
from one hour to two hours, all the hydrophobic samples (HY #1,
HY #2 and HY #3) displayed lower heat transfer coefficients, in
agreement with those measured on the hydrophilic samples. The
contact angles were then checked after tests, finding
hadv 39:6  12:6 and hrec  0 and thus proving the surface
degradation from the initial hydrophobic characteristics.

In the light of these results, the enhancement of condensation


heat transfer over hydrophobic surfaces cannot be explained with
dropwise condensation on some part of the heat transfer area
because discrete droplets were not observed neither a solidliquid-vapor contact line was observed.
No physical explanation of the present results can be proved at
the moment. Nevertheless, since our surfaces are completely
flooded, some discussion can be done by looking at previous experiments during liquid flow on hydrophobic surfaces. As reported in
Section 1.1, the drag reduction of liquid flow over hydrophobic surfaces could be explained by supposing a slip velocity at the wall. In
the present tests, the presence of a slip mechanism cannot be
demonstrated, however it may be interesting to notice that the
assumption of a slip velocity of the condensate layer at the wall
implies some heat transfer enhancement. The ratio of HTCHY to
HTCHPHIL reflects how the hydrophobicity affects the condensation
performance as compared to an untreated (hydrophilic) specimen.
The present results suggest that the measured heat transfer
enhancement must be related to two main parameters: vapor
velocity and heat flux. As shown in Fig. 9, the ratio of heat transfer
coefficient (HTCHY/HTCHPHIL) at constant DTml increases with the
steam velocity, which means that for the same temperature difference between saturation and wall the relative performance of the
hydrophobic surface is better at higher steam velocity. Limiting
this discussion at qualitative level, it can be said that a higher heat
transfer coefficient must be due to a lower film thickness. Studies
in the literature [19] suggest that when a slip velocity occurs, it is
proportional to the fluid velocity which means that by increasing
the liquid-vapor interfacial velocity we get a higher slip velocity
and thus a lower liquid film thickness. The results in Fig. 9 also
show that the ratio of heat transfer coefficient (HTCHY/HTCHPHIL)
decreases when the saturation-to-wall temperature difference
increases, i.e. when the film thickness increases. This suggests that,
if a slip mechanism occurs, its effect diminishes when the film is
thicker. Interesting to say, as pointed out in Section 1.1, these
two effects of steam velocity and heat flux on the heat transfer
enhancement are in agreement with the literature on slippery
mechanism [1,17,19]. In the present study no measurement of
the condensate film velocity is performed. However, a qualitative
analysis compared with the literature shows that assuming a slip
velocity at the wall goes in the direction of explaining the measured heat transfer enhancement.

40

4.4. Comparison with condensation model

HYDROPHOBIC
SAMPLES

35

HTCEXP [kW m-2 K-1]

30
25
20

HPHIL #1
HPHIL #2

15

HY1
10

HY2

HYDROPHILIC
SAMPLES

HY3

5
0
0

10

15

20

25

Tml [C]
Fig. 10. Heat transfer coefficient versus temperature difference between the
saturated steam and the surface during condensation over hydrophilic (HPHIL #1
and HPHIL #2) and hydrophobic (HY #1, HY #2 and HY #3) samples. Data refer to
G = 5.1 kg m2 s1.

The present experimental data are compared against a model


available in the open literature for filmwise condensation. In the
present data, both the gravitational effect and the shear stress
effect affect the condensation process, thus both need to be
accounted for in the calculations. The calculations start from the
classical theories of gravity [30] and shear controlled condensation
heat transfer [33]. Thus, the model applied here is the one already
mentioned in Section 4.1 (Eq. (10)).
The results of this comparison are reported in Fig. 11, both for
the hydrophilic surface (HPHIL #1 in Fig. 11a) and the hydrophobic
sample (HY #1 in Fig 11b). The deviation between the calculated
and the experimental values increases when increasing the steam
mass flow rate: this is typical in the case of shear driven condensation process, because of the higher uncertainty in the estimation of
the shear stress effect on the heat transfer. However, the model
clearly underestimates the experimental data measured on the
hydrophobic surface (Fig. 11b), by 20 to 30%, while the agreement
is within 10% for the heat transfer coefficients measured over the
untreated hydrophilic sample (Fig. 11a). Therefore, the classic filmwise condensation theory is not able to describe the present
phenomenon.

D. Del Col et al. / International Journal of Heat and Mass Transfer 107 (2017) 307318

20
-2 -1

G=1.5 kg m s

+10%

-2 -1

Serie3 kg m s
G=3.1

HTCCALC [kW m -2 K-1]

-10%

-2 -1

15

Serie9
G=5.1 kg m s

10

HYDROPHILIC
0
0

10
HTCEXP [kW m -2 K-1]

15

20

(a)
30
-2 -1

+10%

G=1.5 kg m s
25

-2 -1
-2 -1

HTCCALC [kW m -2 K-1]

3. During steam condensation over the hydrophobic treated substrates, enhanced heat transfer has been measured although
condensation always occurred in filmwise mode caused by
the flooding of the surfaces, due to the high heat flux and the
surface roughness.
4. The heat transfer enhancement with respect to the hydrophilic
samples increases when increasing vapor mass velocity and
reduces when increasing saturation-to-wall temperature difference. In the tested conditions, the heat transfer coefficient is
higher by 1045% on the hydrophobic surface as compared to
the hydrophilic sample.
5. Several samples, both hydrophilic and hydrophobic, have been
tested and the repeatability of the present results was confirmed. On the hydrophobic samples, after a certain period, a
change in the surface wettability was found, with decrease of
the heat transfer coefficients to the values of the hydrophilic
surface. Contact angle measurements have been done after tests
on the hydrophobic surfaces, demonstrating the degradation of
the fluorosilane layer.
6. The condensation heat transfer enhancement cannot be
explained using the classical model of filmwise condensation.
Besides, no droplets were observed on the surface neither a
solid-liquid-vapor contact line was observed.
7. Available literature on liquid flow over hydrophobic surfaces
suggests an explanation of the present results by assuming
some slip of the condensate at the wall. This assumption is
not experimentally proved in the present study but it goes in
the direction of explaining the measured heat transfer enhancement because it implies a thinner condensate film.

-10%

Serie3
G=3.1 kg m s

Serie9
G=5.1 kg m s

20

317

Acknowledgments
The European Space Agency is greatly acknowledged for supporting this work through the MAP Condensation program.

15

References

10

HYDROPHOBIC
0
0

10
15
20
HTCEXP [kW m -2 K-1]

25

30

(b)
Fig. 11. Comparison between the experimental data acquired during condensation
over the hydrophilic (HPHIL #1) sample (a) and over the hydrophobic (HY #1)
sample (b). The values are calculated with the filmwise condensation model (Eq.
(10)).

5. Conclusions
1. In this paper, a method to impart hydrophobic properties over
aluminum substrates is presented. This consists in forming onto
the metal a low surface energy film, by spin coating a
fluorosilane-Hexane solution over it.
2. Tests of pure steam flow condensation over hydrophobic treated surfaces hadv 140 are performed at heat flux between
250 and 500 kW m2 and steam velocity between 2.2 and
6.4 m s1. The present results are compared with those found
when testing hydrophilic untreated specimens hadv < 90 .

[1] J.P. Rothstein, Slip on superhydrophobic surfaces, Annu. Rev. Fluid Mech. 42
(2010) 89109.
[2] C. Antonini, M. Innocenti, T. Horn, M. Marengo, A. Amirfazli, Understanding the
effect of superhydrophobic coatings on energy reduction in anti-icing systems,
Cold Reg. Sci. Technol. 67 (2011) 5867.
[3] A.R. Betz, J. Jenkins, C.J. Kim, D. Attinger, Boiling heat transfer on
superhydrophilic, superhydrophobic, and superbiphilic surfaces, Int. J. Heat
Mass Transfer 57 (2013) 733741.
[4] E. Schmidt, W. Schurig, W. Sellschopp, Versuche ber die kondensation von
wasserdampf in film- und tropfenform, Forsch. Ingenieurwes. 12 (1930) 53
63.
[5] S. Vemuri, K.J. Kim, An experimental and theoretical study on the concept of
dropwise condensation, Int. J. Heat Mass Transfer 49 (2006) 649657.
[6] S. Vemuri, K.J. Kim, B.D. Wood, S. Govindaraju, T.W. Bell, Long term testing for
dropwise condensation using self-assembled monolayer coatings of noctadecyl mercaptan, Appl. Therm. Eng. 26 (2006) 421429.
[7] R. Jafari, M. Farzaneh, Fabrication of superhydrophobic nanostructured surface
on aluminum alloy, Appl. Phys. A Mater. Sci. Process. 102 (2011) 195199.
[8] M.M. Derby, A. Chatterjee, Y. Peles, M.K. Jensen, Flow condensation heat
transfer enhancement in a mini-channel with hydrophobic and hydrophilic
patterns, Int. J. Heat Mass Transfer 68 (2014) 151160.
[9] R.A. Erb, Dropwise condensation on gold improving heat transfer in steam
condensers, Gold Bull. 6 (1973) 26.
[10] L. Zhong, M. Xuehu, W. Sifang, W. Mingzhe, L. Xiaonan, Effects of surface free
energy and nanostructures on dropwise condensation, Chem. Eng. J. 156
(2010) 546552.
[11] D. Torresin, M.K. Tiwari, D. Del Col, D. Poulikakos, Flow condensation on
copper-based nanotextured superhydrophobic surfaces, Langmuir 29 (2013)
840848.
[12] J.W. Rose, Dropwise condensation theory and experiment: a review, Proc. Inst.
Mech. Eng. Part A J. Power Energy 216 (2002) 115128.
[13] R. Enright, N. Miljkovic, J.L. Alvarado, K. Kim, J.W. Rose, Dropwise condensation
on micro- and nanostructured surfaces, Nanoscale Microscale Thermophys.
Eng. 18 (2014) 223250.
[14] N. Miljkovic, E.N. Wang, Condensation heat transfer on superhydrophobic
surfaces, MRS Bull. 38 (2013) 397406.

318

D. Del Col et al. / International Journal of Heat and Mass Transfer 107 (2017) 307318

[15] C. Fang, J.E. Steinbrenner, F.M. Wang, K.E. Goodson, Impact of wall
hydrophobicity on condensation flow and heat transfer in silicon
microchannels, J. Micromechan. Microeng. 20 (2010) 45018.
[16] S.S. Hsieh, C.Y. Lin, Convective heat transfer in liquid microchannels with
hydrophobic and hydrophilic surfaces, Int. J. Heat Mass Transfer 52 (2009)
260270.
[17] F.Y. Lv, P. Zhang, Drag reduction and heat transfer characteristics of water flow
through the tubes with superhydrophobic surfaces, Energy Convers. Manage.
113 (2016) 165176.
[18] D.C. Tretheway, C.D. Meinhart, A generating mechanism for apparent fluid slip
in hydrophobic microchannels, SOCAR Proc. 2010 (2010) 6068.
[19] S. Granick, Y. Zhu, H. Lee, Slippery questions about complex fluids flowing past
solids, Nat. Mater. 2 (2003) 221227.
[20] G.K. Batchelor, An Introduction to Fluid Dynamics, Cambridge Univ. Press,
1985.
[21] D.C. Tretheway, C.D. Meinhart, Apparent fluid slip at hydrophobic
microchannel walls, Phys. Fluids 14 (2002) 812.
[22] R.S. Voronov, D.V. Papavassiliou, L.L. Lee, Slip length and contact angle over
hydrophobic surfaces, Chem. Phys. Lett. 441 (2007) 273276.
[23] C.H. Choi, K.J.A. Westin, K.S. Breuer, Apparent slip flows in hydrophilic and
hydrophobic microchannels, Phys. Fluids 15 (2003) 28972902.
[24] B. Bhushan, Y. Wang, A. Maali, Boundary slip study on hydrophilic,
hydrophobic, and superhydrophobic surfaces with dynamic atomic force
microscopy, Langmuir 25 (2009) 81178121.
[25] S. Pati, S.K. Som, S. Chakraborty, Slip-driven alteration in film condensation
over vertical surfaces, Int. Commun. Heat Mass Transfer 46 (2013) 3741.

[26] J.A. Al-Jarrah, A.F. Khadrawi, M. A. AL-Nimr, Film condensation on a vertical


microchannel, Int. Commun. Heat Mass Transfer 35 (2008) 11721176.
[27] E.W. Lemmon, M.L. Huber, M.O. McLinden, NIST Standard Reference Database
23: Reference Fluid Thermodynamic and Transport Properties-REFPROP,
Version 9.1., 2013.
[28] ISO Guide to the Expression of Uncertainty in Measurement, 1995.
[29] A. Bisetto, S. Bortolin, D. Del Col, Experimental analysis of steam condensation
over conventional and superhydrophilic vertical surfaces, Exp. Therm. Fluid
Sci. 68 (2015) 216227.
[30] W. Nusselt, Die Oberflchenkondensation des Wasserdampfes, Zeitschrift VDI.
60 (27) (1916), 541546 and 569575.
[31] H.D. Baehr, K. Stephan, Waerme- und Stoffuebertragung, Springer-Verlag,
2004.
[32] C.A. Depew, R.L. Reisbig, Vapor condensation on a horizontal tube using teflon
to promote dropwise condensation, I&EC Process Des. Dev. 3 (1964) 365369.
[33] G.F. Hewitt, G.L. Shires, T.R. Bott, Process Heat Transfer, 1994.
[34] L. Friedel, Improved friction pressure drop correlations for horizontal and
vertical two-phase pipe flow, in: Eur. Two-Phase Flow Gr. Meet. Ispra, 1979.
[35] N. Miljkovic, R. Enright, Y. Nam, K. Lopez, N. Dou, J. Sack, E.N. Wang, Jumpingdroplet-enhanced condensation on scalable superhydrophobic nanostructured
surfaces, Nano Lett. 13 (2013) 179187.
[36] R. Enright, N. Miljkovic, A. Al-Obeidi, C.V. Thompson, E.N. Wang, Condensation
on superhydrophobic surfaces: the role of local energy barriers and structure
length scale, Langmuir 28 (2012) 1442414432.

S-ar putea să vă placă și