You are on page 1of 8

Experimental Thermal and Fluid Science 47 (2013) 9097

Contents lists available at SciVerse ScienceDirect

Experimental Thermal and Fluid Science


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

Condensation on downward-facing surfaces subjected to upstream ow


of airvapor mixture
Fernando F. Czubinski a, Marcia B.H. Mantelli a,, Jlio C. Passos b
a
b

Departamento de Engenharia Mecnica, LEPTEN/Labtucal Centro Tecnolgico, Universidade Federal de Santa Catarina, 88040-900 Florianpolis, SC, Brazil
Departamento de Engenharia Mecnica, LEPTEN/Boiling Centro Tecnolgico, Universidade Federal de Santa Catarina, 88040-900 Florianpolis, SC, Brazil

a r t i c l e

i n f o

Article history:
Received 26 September 2012
Received in revised form 7 January 2013
Accepted 10 January 2013
Available online 4 February 2013
Keywords:
Film condensation
Condensing surfaces
Noncondensable gases
Inclined smooth surfaces
Grooved surfaces

a b s t r a c t
This paper reports experimental results for lm condensation on vertical and downward inclined smooth
and grooved surfaces subjected to ascending streams of a vaporair mixture. An experimental facility was
built in order to evaluate the heat uxes for different surface inclinations, surface sub-cooling temperatures and air to vapor ratios in the mixture. Two experimental procedures were employed to evaluate the
heat absorption rate during the condensation process. The employment of grooved surfaces resulted in a
10% enhancement of the condensation rate in the case of the pure vapor condensation and a negligible
effect when noncondensable gases (NCGs) were present in the mixture.
2013 Elsevier Inc. All rights reserved.

1. Introduction
The condensation heat transfer phenomenon is encountered in
a variety of engineering and industrial applications. Knowledge of
the physical mechanisms which drive condensation is important
for the design of several types of equipment.
The rst lm condensation model was proposed by Nusselt in
1916. It assumes negligible effects of the interfacial shear stress
at the vapor/liquid interface, and equates gravity and viscous
forces for a quiescent pure vapor environment in contact with an
isothermal vertical smooth plate. A linear temperature prole
was considered. Subsequently, a number of researchers improved
Nusselts model by removing some of the original restrictive
assumptions and including effects like sub-cooling, a nonlinear
temperature prole, inertial terms and interfacial stress [14].
The wettability of the condenser surface plays an important
role, as the interaction between the condensate liquid and the condensing surface dictates the mode in which condensation occurs.
The dropwise condensation mode is associated with higher heat
transfer coefcients when compared to lm condensation. Therefore, many researchers have directed their efforts to the development of suitable condensation promoters such as polymeric
lms, a monolayer of organic materials and ion-plating technology.
These techniques provide poor wettability of the substrate and
thus dropwise condensation is maintained for a longer period [5
Corresponding author. Tel./fax: +55 48 3721 9379.
E-mail address: marcia@emc.ufsc.br (M.B.H. Mantelli).
0894-1777/$ - see front matter 2013 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.expthermusci.2013.01.004

8]. However, as is well known, dropwise condensation is often difcult to maintain for all the period.
Wavy condenser surfaces, also known as Gregorig surfaces [9],
are used to produce better heat transfer rates for the lm condensation mode. The enhancement is caused by the variation in the
interface radius of the liquid lm over the substrate, which, together with the surface tension, produces a gradient pressure in
the condensate, driving the lm from the crest to the valley of
the surface grooves, improving the condensation performance.
The tests performed showed that this conguration results in heat
transfer coefcients up to ve times larger than those for smooth
surfaces, for the same projected area [1012].
In some practical operations, noncondensable gases (NCGs) may
be found in the condensing vapor, deteriorating the heat transfer
process. This is caused by the formation of a gas boundary layer
over the condensate liquid lm, reducing the partial vapor pressure
at the condensate interface. Several studies have been conducted
on stagnant and forced convection occurring over smooth plates
or inside and outside tubes in order to investigate the effect of
an NCG. For a quiescent condensing mixture on smooth plates,
the decrease in heat transfer rates could be around 50% with an
air fraction of only 0.5% (by mass) in the airvapor mixture. In
cases where the mixture ows, the reduction in the heat transfer
decreases since the stream disperses the gas boundary layer. This
effect is dependent on the Reynolds number of the mixture. Another effect that could spread the gas boundary layer is the undulation in the liquid lm condensation interface, which occurs in the
wavy and turbulent ow regime [1316].

F.F. Czubinski et al. / Experimental Thermal and Fluid Science 47 (2013) 9097

91

Nomenclature
Alphabetic
NCG
Noncondensable gase
Q
Heat transfer rate (kW)
_
m
mass ow rate (kg s1)
Cp
Specic heat (kJ kg1 K1)
Tout
Outlet cooling water temperature (K)
Inlet cooling water temperature (K)
Tin
hlv
Latent heat (kJ kg1)

h
Mean heat transfer coefcient (kW m2 K1)
g
Gravity (m s2)
L
Surface length (m)
DT
Sub-cooling temperature (K)
A
Area (m2)
T
Temperature (K)
Sc=l/qD Schmidt dimensionless number ()
W
Air mass fraction ()
M
Molecular weight (kg Kmol1)
D
Diffusivity (m2 s1)

As mentioned above, several studies have focused on theoretical


assessments, modes of uxes and ow regimes, and the composition and type of mixture in different geometrical congurations.
In this study, experimental analysis and models presented in the
literature will be employed to investigate the downward condensation of vapor contained in airvapor mixtures, in vertical ascension, reaching the cooled surface of a condensing plate from below.

Pressure (atm)

Greek Letters
Density (kg m3)
h
Surface inclination ()
j
Thermal conductivity (kW m1 K1)
l
Dynamic viscosity (kg m1 s1)

Subscript
w
cond
liq
vap
sat
G
1
0

Cooling water
Condensate
Liquid
Pure water vapor
Saturation
Gas (liquid vapor mixture)
Bulk
Condensate interface

2. Experiment
2.1. Experimental apparatus
Fig. 1 shows a scheme of the experimental facility. It consists of
a test section and auxiliary equipment (steam generator, cooling
water system, NCG supplier and data acquisition system) which
were grouped into three main sections: boiler, vapor supply line

1.1. Literature review


Condensation studies using this particular conguration have
been carried out by Gerstmann and Grifth [17] and Chung et al.
[18,19]. This phenomenon can be observed in condensers designed
for the recovery of water from humid air which is released to the
atmosphere in industrial cooling towers [20].
Gerstmann and Grifth [17] investigated the condensation of
pure Freon-113 on the underside of a smooth plate in a closed
chamber. The heat transfer rates were of the same order of magnitude as those predicted by Nusselt analysis (which considers a quiescent condensing vapor). Chung et al. [18], also used a smooth
plate as the condensing surface and showed that in the case of pure
water vapor the condensation heat transfer agreed with the Nusselt theoretical model, even though the experimental tests were
conducted under a slow ow. With NCGs, when this slow ow
reaches the surface, the NCG boundary layer is spread, so that
the effect of the NCG on the heat transfer rate is not so obvious.
Chung et al. [18] used a deector before the mixture ow hit the
cooled condensing surface within the condensing chamber, to
avoid the impact of the direct ow on the surface. They also studied congurations where the condensing chamber was closed, to
create a quiescent environment.
The main objective of the study reported herein was to investigate experimentally the condensation of the water of an upstream
airvapor mixture ow, which hits directly the downward face of a
cooled surface. It should be noted that the condensation of water
under such conditions is a subject not previously explored in the
literature. The heat transfer rate resulting from the lm condensation process was measured. Models and results available in the literature were used as benchmarks for the analysis of the physical
phenomenon involved. The effects of smooth and grooved surfaces
of electrolytic copper and aluminum alloys were also studied, varying the inclination angles, the sub-cooling temperature and the
amount of air in the airvapor mixture.

Fig. 1. Schematic view of experimental apparatus.

92

F.F. Czubinski et al. / Experimental Thermal and Fluid Science 47 (2013) 9097

and test section. All parts of the equipment were made of galvanized carbon steel plates, of 2 mm thickness, and were thermally
insulated using 50 mm rockwool sheets in order to reduce the heat
loss to the surroundings.
The boiler, which consists of a 600 mm  500 mm  270 mm
rectangular vessel, has electrical heaters inside to supply the controlled power needed to generate the steam. The NCG employed
was atmospheric air, supplied with the aid of an air blower through
a square duct with 50 mm of internal cross section and 300 mm of
length (see Fig. 1). The air was warmed to the same temperature as
the steam using a controlled electrical power source to provide
power for the air heater, so that the resulting mixture is kept in
the dry saturated condition. The air mass ow rate can be calculated by the thermal capacity of the air times the temperature difference before and after the heater measured by thermocouples.
The air-vapor mixture is created by these two streams, which meet
at the vapor supply line. The supply line is a cylindrical vertical
tube with 147 mm of inner-diameter and 1000 mm of length.
The test section, illustrated in Fig. 2, is a box with a square base
with 250 mm of edge and 350 mm of height. One wall of the chamber is made of glass to allow the visualization of the condensing
phenomenon. The mixture formed in the vapor supply line enters
the test section from the bottom and directly reaches the condensing surface. This surface is xed by pins located in the lateral walls,
allowing inclination variations.
Different inclinations of the cooling plate lead to variations in
the cross sectional area of the airvapor stream ow. Therefore,
the ow area beside the plate changes with the inclination, causing
different degrees of connement of the mixture. To keep approximately the same connement (same cross sectional area of the
ow) for all tested cases, plates with different lengths, which vary
according to their inclination, were constructed. Also, one of the
plates was tested with different inclinations to study the connement effect.
Smooth electrolytic copper plates with lengths of 116.2 mm,
142.8 mm and 200 mm were tested at inclinations of 30, 45,
60, respectively. This latter plate was tested at inclinations of
60 and 90 (vertical position).
To collect the condensed water, a drain was installed in the lower part of each surface, as shown in Fig. 2. Seven thermocouples
were inserted within small holes drilled in parallel and 0.1 mm below the condensing surface, in order to measure the surface temperature distribution, as can be seen in Fig. 3a.
The back face of the condensing test plate closes a small heat
exchanger, comprised of a hollow thermal insulated metallic parallelepiped box, where cooling water circulates at controlled rates
and temperatures, allowing different sub-cooling levels of the condensing surface (see Fig. 3b).

Fig. 3. Distribution of thermocouples (a). Surface cooling system (b).

The inuence of the surface material and nishing on the condensation was tested by means of surfaces of the same length
(200 mm) and same inclination (60), made of smooth and grooved
aluminum 5052 alloy and grooved electrolytic copper plates. The
grooves were made in the longitudinal direction of a plate of the
same thickness as the smooth plate being tested, with a distance
of 2 mm between crests and 1.44 mm of depth. Fig. 4 shows the
testing surfaces with their main dimensions.
2.2. Data reduction
The experimental heat transfer rates were measured by two
methods. In the rst method, the heat absorbed by the cooling
water was calculated by multiplying the mass ux rate by the difference between the inlet and outlet water temperatures, using the
expression:

_ w T out  T in
Q w mCp

_ w is the coolant water ow rate veried by a calibrated


where m
rotameter, Cpw is specic heat of the cooling water and Tout, Tin
are the outlet and inlet temperatures of the cooling water,
respectively.
In the second method, the condensate heat transfer rate is calculated as:

_ cond hlv
Q cond m

_ cond is the condensate mass ow rate determined with the


where m
use of a weighing scale and hlv is the latent heat of vaporization.In
order to assist the interpretation of the experimental results for the

Fig. 2. Test section.

F.F. Czubinski et al. / Experimental Thermal and Fluid Science 47 (2013) 9097

93

temperature is in good agreement with the guessed value, the heat


transfer rate can be obtained by using Eq. (3). If there is no agreement, this procedure is repeated until a good agreement is reached.
In the study reported herein, these values were evaluated up to
103.
The thermophysical properties of the condensate liquid boundary layer where evaluated at the temperature suggested by Minkowycz and Sparrow [14]. The properties of the mixture were
assessed on a mass fraction basis considering the initial state of
formation. The diffusion coefcient of water vapor in air was calculated using the following equation given by Marrero and Manson
[21],

DH2 O;air 1:87  1010

Fig. 4. Surfaces tested.

pure vapor, the theoretical heat transfer rate was determined using
the modied Nusselt model [2]:

314

qliq qliq  qv ap gsinhk3liq hlv



4
5
h

0:943
Nusselt
Llliq DT

where qliq, jliq and lliq are the density, thermal conductivity and
viscosity of the condensed liquid density respectively, qvap is the
pure vapor density, jliq is the thermal conductivity, g is the force
of gravity, L is the surface length, h is the surface inclination from
the horizontal and DT is the difference between the temperatures
of the interface and the wall.
Thus, the heat transfer rate is obtained by:

Q Nusselt hNusselt Asurface T sat  T wall

where hNusselt is determined from Eq. (3), Asurface is the surface area
of each condensing plate and Tsat, Twall represent the temperatures
of the pure saturated steam at the interface and the surface temperature, respectively, measured by the thermocouples.
To account for the presence of NCG, the Rose model [10,13] was
used to predict the theoretical heat transfer. In this case, in contrast to the pure vapor, the saturation temperature at the interface
is not previously known, and thus the following equation is used to
estimate this parameter:


lliq qliq W 1 2 20
W0
Sc
21
lG qG w0
W1
! 

8
lG qG
w0
5
w0
2
FX
W 0 28
3
F Sc lliq qliq

10FSc

100 W 1
w0
2
8Sc
21 W 0
W0

where qG and lG are the density and viscosity of the airvapor mixture, respectively, W0 is the air mass fraction at the condensate
interface, W1 is the air mass fraction of the mixture, Sc is the
Schmidt number for the gas mixture, F = jliq (T0 Twall) /lliqhlv,
where T0 is the interface temperature, w0 = W0  W1 and X =
(Mair Mvapor)/[Mair W1 (Mair Mvapor)], where Mair and Mvapor
are the molecular weight of air and water respectively.
The evaluation of the interface temperature therefore begins by
guessing a value for this parameter, which must be between the
temperatures of the surface and the initial mixture formation.
Therefore, F, in Eq. (5) can be computed and this equation is solved
for W0, the noncondensable air mass fraction at the interface. Since
the vapor mass fraction is equal to 1 W0, the partial pressure of
the vapor at the interface is evaluated by assuming equilibrium
at the interface and using Raoult and Daltons Law. With the partial
vapor pressure, the corresponding saturation temperature can be
ascertained from steam tables. If this new value for the interface

T 2:072
1
P1

where T and P are the temperature and pressure of the system.


The tests were carried out at atmospheric pressure with a steam
mass ow rate of 2.4 g/s. Air (NCG) was combined with the steam
resulting in mixtures with mass fractions varying from 0 to 50%.
The surface temperature varied (between 35 and 85 C) for each
condition of NCG and each inclination.
The boiler heat loss through the insulation to the environment
was estimated to be around 1.7% of the power. The heat loss from
the air heating section was estimated to be between 3% and 8%,
depending on the input power.
Uncertainty analysis was performed for the experimental determination of the heat transfer. The temperatures were measured by
calibrated k-type thermocouples, which provide an accuracy of
0.25 C. 5% of uncertainty was evaluated for the air mass ow
measurement. The uncertainties observed for the heat transfer rate
using Eq. (1) varied from 10 to 50%. With Eq. (2), these uncertainties were less than 5%. Large uncertainties were associated with Eq.
(1) because the difference between the inlet and outlet cooling
water temperature is very small, and virtually the data acquisition
system shows almost the same values as temperature readings. On
the other hand, the uncertainties obtained using Eq. (2) are much
smaller because the volume of collected condensate, which was
performed in a large time frame, was much larger that the smallest
reading of the weighting scale. These heat transfer uncertainties
are presented by vertical error bars superposing the data in the
plots shown in the gures of the next section. In some of the data
presented, these uncertainties are very small and difcult to observe in these plots.
One should note that the thermocouples are inserted inside the
cooled surface at 1 mm of distance from the condensation surface,
as already mentioned, causing a maximum wall temperature measurement error of 1 C. The propagated error using Eq. (4) for the
heat transfer calculation is negligible.
3. Results and discussion
Fig. 6 shows the test results for the heat ux transferred during
the condensation of pure vapor on a vertical smooth plate, as a
function of the surface temperature. The solid line represents the
predictions provided by the Nusselt model, Eq. (4), while the solid
and open symbols show the experimental values obtained from
Eqs. (1) and (2), respectively. The heat uxes obtained from Eqs.
(1) and (2) showed a good agreement between them but not with
the Nusselt theory. As the condensing vapor is not stagnant (the inlet and outlet of the test section are fully opened), a large volume of
the vapor stream does not have contact with the condensing surface, owing freely inside the condensing chamber. In order to verify this situation, two modications were made to create a
stagnant environment. Firstly, as performed by Chung et al. [18],
a deector was used to avoid the direct impact of the upstream

94

F.F. Czubinski et al. / Experimental Thermal and Fluid Science 47 (2013) 9097

Fig. 5. Modication in the test section.

400

Eq. (1)
Eq. (2)
Eq. (4) Nusselt

q" [kW m-2]

300

200

100

0
40

60

80

Surface Temperature [C]


Fig. 6. Smooth plate in the vertical position using pure vapor.

ow on the downward face of the condensing plate and the test


section outlet was almost completely closed, as can be seen in
Fig. 5. Table 1 shows the results obtained for one testing case with
this modication, showing a good agreement with the Nusselt theory This means that the experimental apparatus is well designed
and that its testing parameters are well controlled, since the heat
ux results approach the Nusselt predictions as the ow approaches the stagnant conditions, as assumed in the Nusselt model. One should note it was not possible to obtain more data with
this set up because, as the test chamber was full of warm vapor,
more cooling power in the condensate plate would be necessary.
The thermal bath used was not able to cool down the plate to temperature levels tested in the other experiments (from 40 to 80 C),
so that only the plate temperature of around 83 C was tested.
Fig. 7 illustrates the heat transfer results observed in the condensation of vapor from a mixture of airvapor, condensing on a
vertical smooth copper plate. The line represents the theoretical
results obtained by Rose, Eq. (5), the symbols the experimental
measurements and the vertical bars the degree of uncertainty.
As can be seen in Fig. 7, the heat transfer rate decreases systematically as the surface temperature and the presence of NCG increase, both for experimental and theoretical results. Actually the
NCG tends to accumulate at the liquid interface, resulting in a

Table 1
Heat transfer results.
Surface temperature (C)

82.6

Heat ux (kW/m2)
Eq. (4) Nusselt

Eq. (1)

Eq. (2)

138.9

125.4

128.3

reduction in the partial pressure of the steam, decreasing the saturation temperature at which the condensation takes place, and,
therefore, reducing the heat ux. In other words, an additional
thermal resistance is built up, reducing the heat transfer rate. On
the other hand, the heat uxes predicted by the Rose model for
mixtures in quiescent environment are smaller than those observed in streams, where the ow hitting the condensation surface
spreads the NCG boundary layer. In the latter case, the saturation
temperature is not so strongly affected by the accumulation of
the NCG. In other words, Rose considered a stagnant environmental while in this work the mixture was in an upstream ow. For the
case with a 20% air mass fraction, the heat ux decrease, compared
with the pure vapor case, was between 15% and 22%, depending on
the surface temperature. Rose [13] reported that for a quiescent
mixture with a 0.5% air mass fraction the NCG caused a large thermal resistance, with a decrease of more than 50% in the heat ux.
Similar trends were observed for plate inclinations of 60, 45
and 30, as shown in Figs. 810, respectively, for smooth copper
plate
The Nusselt theory shows that the heat transfer for vertical
plates (in a quiescent environment) is larger than for tilted surfaces, as the gravity pulls the liquid lm downward. However, on
comparing Fig. 6 (vertical case) with Figs. 810 (tilted cases), it is
clear that this effect is not observed. It was noted that, as the surface approximates the horizontal position the ascending stream
tends to be more trapped by the condensing surface and the vapor
of the airvapor mixture exchanges more heat with the cooled surface, increasing the production of condensate and, therefore, the
heat transfer coefcient. It is well known that the mean liquid lm
thickness, which acts as thermal resistance, increases as the surface length increases (larger condensing area). As already explained, the length of the tested plates decreases as their
inclinations tend to the horizontal position, and therefore, the lm
thickness and the thermal resistance are expected to decrease.
Actually this is not observed. Taking Fig. 7 (vertical case) and 8
(60 case), although the tested surface is the same (and so the same
mean liquid lm thickness), the heat transfer rate for the inclined
surface is larger than for the vertical case. Therefore, the tests show
that the effect of the liquid lm thickness is less important than the
inuence of inclination. Consequently, the condensation heat
uxes were higher for an inclination of 30 (Fig. 10) than for 45
(Fig. 9) and 60 (Fig. 8).
It is interesting to note that during the experiments and for all
slopes tested, no condensate drops were observed to fall from the
cooling surface. Instead, all of the condensate liquid lm ran down
the downward condensing surface and was collected by the drain.
This means that all the condensate formed over the condensing
surface was collected in the drain, improving the quality of the
experimental data.
Also, one should note that the heat transfer calculated using Eq.
(2) is always larger than that calculated using Eq. (1) for all inclined
testing cases, although most of the data lie within the experimental
uncertainty range. This is not observed for the vertical case. Even
though it is not possible to explain exactly why this happens, it
is believed that this difference is due to the different stream congurations over the condensing plates in both positions. More investigation would be need to a deep understanding of this
phenomenon. In spite of this difference the comparison between
the heat transfer determined by Eqs. (1) and (2) is better than with
the results of the literature theoretical models.
The inuence of other surface materials was also investigated in
this study. Samples of aluminum and copper, in the smooth and
grooved surface congurations, were tested for an inclination of
60. In the following plots the theoretical results and degrees of
uncertainty are not shown, as no new information other than that
previously discussed herein can be obtained.

F.F. Czubinski et al. / Experimental Thermal and Fluid Science 47 (2013) 9097

Eq.1PureVapor
Eq.2PureVapor
Eq.1 - 20% NCG
Eq.2 - 20% NCG
Rose - 20% NCG
Eq. 1 - 30% NCG
Eq. 2 - 30% NCG
Rose - 30% NCG
Eq. 1 - 40% NCG
Eq. 2 - 40% NCG
Rose - 40% NCG
Eq. 1 - 50% NCG
Eq. 2 - 50% NCG
Rose - 50% NCG

30

q" [kW m-2]

95

20

10

0
40

60

80

Surface Temperature [C]


Fig. 7. Smooth plate in the vertical position.

Eq.1PureVapor
Eq.2PureVapor
Eq.1 - 20% NCG
Eq.2 - 20% NCG
Rose - 20% NCG
Eq. 1 - 30% NCG
Eq. 2 - 30% NCG
Rose - 30% NCG
Eq. 1 - 40% NCG
Eq. 2 - 40% NCG
Rose - 40% NCG
Eq. 1 - 50% NCG
Eq. 2 - 50% NCG
Rose - 50% NCG

40

q" [kW m-2]

30

20

10

0
40

60

80

Surface Temperature [C]


Fig. 8. Smooth plate with an inclination of 60.

50

Eq.1PureVapor
Eq.2PureVapor
Eq.1 - 20% NCG
Eq.2 - 20% NCG
Rose - 20% NCG
Eq. 1 - 30% NCG
Eq. 2 - 30% NCG
Rose - 30% NCG
Eq. 1 - 40% NCG
Eq. 2 - 40% NCG
Rose - 40% NCG
Eq. 1 - 50% NCG
Eq. 2 - 50% NCG
Rose - 50% NCG

q" [kW m-2]

40

30

20

10

0
40

60

80

Surface Temperature [C]


Fig. 9. Smooth plate with an inclination of 45.

In Figs. 1113, the heat ux is plotted against surface temperature for the condensing surfaces at 60 degrees of inclination, for
grooved copper, grooved aluminum and smooth aluminum surfaces, respectively. In the case of the grooved surfaces, the heat
uxes were evaluated based on the apparent projected surface
area.
The results for the heat uxes show that the behaviors of the
smooth and grooved surfaces follow the same trend, that is, an in-

crease in the surface temperature and in the amount of NCG leads


to a decrease in the heat ux.
On comparing Figs. 1113 with Fig. 8 (all experimental data for
surfaces with the same inclination) it can be observed that the
highest heat uxes were obtained for the pure vapor with the
grooved copper surface. The heat transfer was enhanced by around
10% compared with the smooth copper surface. This result is much
lower than that observed by Markowitz et al. [11], where an

96

F.F. Czubinski et al. / Experimental Thermal and Fluid Science 47 (2013) 9097

Eq.1PureVapor
Eq.2PureVapor
Eq.1 - 20% NCG
Eq.2 - 20% NCG
Rose - 20% NCG
Eq. 1 - 30% NCG
Eq. 2 - 30% NCG
Rose - 30% NCG
Eq. 1 - 40% NCG
Eq. 2 - 40% NCG
Rose - 40% NCG
Eq. 1 - 50% NCG
Eq. 2 - 50% NCG
Rose - 50% NCG

50

q" [kW m-2]

40

30

20

10

0
40

60

80

Surface Temperature [C]


Fig. 10. Smooth plate with an inclination of 30.

50

Eq. 1PureVapor
Eq. 2PureVapor
Eq.1 - 20% NCG
Eq. 2 - 20% NCG
Eq.1 - 30% NCG
Eq. 2 - 30% NCG
Eq.1 - 40% NCG
Eq. 2 - 40% NCG
Eq.1 - 50% NCG
Eq. 2 - 50% NCG

q" [kW m-2]

40

30

20

10

0
40

60

80

Surface Temperature [C]


Fig. 11. Grooved copper surface at an inclination of 60.

q" [kW m-2]

40

Eq. 1PureVapor
Eq. 2PureVapor
Eq.1 - 20% NCG
Eq. 2 - 20% NCG
Eq.1 - 30% NCG
Eq. 2 - 30% NCG
Eq.1 - 40% NCG
Eq. 2 - 40% NCG
Eq.1 - 50% NCG
Eq. 2 - 50% NCG

30

20

10

0
40

60

80

Surface Temperature [C]


Fig. 12. Grooved aluminum surface at an inclination of 60.

enhancement of around 150% was achieved for ascending Freon113 (comparing grooved with smooth surfaces). In the study reported herein, the vapor ows upwards and the grooves are not
able to hold the vapor for long and, therefore, the improvement observed is due to an increase in the condensing area, while Markowitz et al. [11] attributed the improvement to the surface tension
and the augmentation of the liquid vapor lm curvature.
For different types of surfaces, it is observed that with an increase in the NCG in the airvapor mixture the heat uxes for

the four different surfaces tend to assume the same value. This
can be clearly observed, for instance, in the case of 50% of NCG.
In other words, copper and aluminum with a grooved or smooth
surface showed the same behavior for high amounts of NCG. As explained by Markowitz et al. [11], the NCG tends to accumulate in
the grooves, blocking the access of the condensate liquid to the valley regions, reducing the efciency of the surfaces.
In summary, when airvapor mixture ascending ows reach
condensing surfaces, the enhancement provided by the channels

F.F. Czubinski et al. / Experimental Thermal and Fluid Science 47 (2013) 9097

97

40

Eq. 1PureVapor
Eq. 2PureVapor
Eq.1 - 20% NCG
Eq. 2 - 20% NCG
Eq.1 - 30% NCG
Eq. 2 - 30% NCG
Eq.1 - 40% NCG
Eq. 2 - 40% NCG
Eq.1 - 50% NCG
Eq. 2 - 50% NCG

q" [kW m-2]

30

20

10

0
40

60

80

Surface Temperature [C]


Fig. 13. Smooth aluminum surface with an inclination of 60.

was observed only in the case of the pure vapor. With NCG, the values for the heat transfer rate are similar for different condensing
surfaces.
4. Conclusions
Film condensation of ascending ows of airvapor mixtures
was experimentally investigated on vertical and downward inclined surfaces, in relation to the effects of different inclination angles, sub-cooling levels and the presence of NCG in the upstream
ow, for smooth and grooved surfaces of copper and aluminum.
The test results were evaluated based on the Nusselt and Rose
theories as benchmarks. Although the presence of NCG had an effect on the decrease in the heat transfer rate, the conguration of
the condensation system plays the main role since the mixture
ows upstream freely inside the condensing chamber.
Reverse trends were observed in the heat transfer rates for the
test results and the theory predictions, considering the effect of
inclination. The experimental results show that the heat uxes increase as the plate inclination decreases from the vertical to an
inclination of 30. As the amount of NCG increases, a systematic
reduction of the heat transfer rate is observed. Due to the mixture
ow, the formation of the NCG boundary layer over the condensate
liquid is disturbed and the decrease in the heat transfer rate was
not so strongly affected by the NCG, as observed in the literature
for quiescent mixtures. The heat transfer variation with the condenser surface sub-cooling level follows the same trend as predicted by the theoretical models.
Furthermore, the enhancement of the heat transfer rates for
grooved surfaces was small and observed only in the case of the
pure vapor, especially for the copper samples. The effect of grooves
was negligible for streams with NCGs.
Acknowledgement
The authors would like to acknowledge the nancial support
provided by Petrobras and CNPq for this research.
References
[1] L.A. Bromley, Effects of heat capacity of condensate, Industrial and Engineering
Chemistry 44 (1952) 29662969.
[2] W.M. Rohsenow, Heat transfer and temperature distribution in laminar lm
condensation, Journal of Heat Transfer 78 (1956) 16451648.

[3] E.M. Sparrow, J.L. Gregg, A boundary-layer treatment of laminar lm


condensation, Journal of Heat Transfer 8 (1959) 1318.
[4] M.M. Chen, An analytical study of laminar lm condensation: Part 1 at
plates, Journal of Heat Transfer 81 (1961) 4854.
[5] X.H. Ma, J.B. Chen, D.Q. Xu, J.F. Lin, C.S. Ren, Z.H. Long, Inuence of processing
conditions of polymer lm on dropwise condensation heat transfer,
International Journal of Heat and Mass Transfer 45 (2002) 34053411.
[6] S. Vermuri, 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, Applied Thermal Engineering 26 (2006) 421429.
[7] L. Zhong, X.H. Ma, W. Sifang, W. Mingzhe, L. Xiaonan, Effects of surface free
energy and nanostructures on dropwise condensation, Chemical Engineering
Journal 156 (2010) 546552.
[8] A.B. Kananeh, M.H. Rausch, A. Leipertz, A.P. Froba, Experimental study of
dropwise condensation on plasma-ion implanted stainless steel tube,
International Journal of Heat and Mass Transfer 49 (2006) 50185026.
[9] V.P. Carey, Liquid-Vapor Phase Change Phenomena: An Introduction to the
Thermophysics of Vaporization and Condensation Process in Heat Transfer,
Taylor & Francis, USA, 1992.
[10] L.C. Burmeister, Convective Heat Transfer, second ed., John Wiley & Sons, USA,
1993.
[11] A. Markowitz, B.B. Mikic, A.E. Bergles, Condensation on a downward-facing
horizontal rippled surface, Transaction of ASME, Journal of Heat Transfer 94
(1972) 315320.
[12] M. Izumi, S. Kumagai, R. Shimada, N. Yamakawa, Heat transfer enhancement of
dropwise condensation on a vertical surface round shaped grooves,
Experimental Thermal and Fluid Science 28 (2004) 243248.
[13] J.W. Rose, Condensation of a vapor in the presence of a non-condensable gas,
International Journal of Heat and Mass Transfer 12 (1968) 233237.
[14] W.J. Minkowycz, E.M. Sparrow, Condensation heat transfer in the presence of
noncondensables, interfacial resistance, superheating, variable properties, and
diffusion, International Journal of Heat and Mass Transfer 9 (1966) 1125
1144.
[15] A.M. Zhu, S.C. Wang, J.X. Sun, L.X. Xie, Z. Wang, Effects of high fractional
noncondensable gas on condensation in the dewvaporation desalination
process, Desalination 214 (2007) 128137.
[16] S.K. Park, M.H. Kim, K.J. Yoo, Effects of a wavy interface on steamair
condensation on a vertical surface, International Journal of Multiphase Flow 23
(1997) 10311042.
[17] J. Gerstmann, P. Grifth, Laminar lm condensation on the underside of
horizontal and inclined surfaces, International Journal of Heat and Mass
Transfer 10 (1967) 567580.
[18] B.J. Chung, S. Kim, M.C. Kim, Film condensations of owing mixtures of steam
air an inclined at plate, International Communication of Heat and Mass
Transfer 32 (2005) 233239.
[19] B.J. Chung, S. Kim, Film condensations on horizontal and slightly inclined
upward and downward facing plates, Heat Transfer Engineering 29 (11) (2008)
936941.
[20] R. Zimmermann, M.B.H. Mantelli, T.P. Borges, C.A. Costa, Viability Study Of
Retrieving The Evaporated Water In A Mechanical Draft Cross Flow Cooling
Tower, Proceedings of The 14th International Heat Transfer Conference Ihtc14
August 813, Washington D.C, USA, 2010.
[21] T.R. Marrero, E.A. Mason, Gaseous diffusion coefcients, Journal of Physical
and Chemical Reference Data 1 (1972) 3118.

You might also like