Você está na página 1de 11

Yang Li

JunJie Yan1
JinShi Wang
State Key Laboratory of Multiphase Flow in
Power Engineering,
Xian Jiaotong University,
Xian City 710049, P.R. China
e-mail: yanjj@mail.xjtu.edu.cn

GuoXiang Wang
Department of Mechanical Engineering,
University of Akron,
Akron, OH 44325-3903
e-mail: gwang@uakron.edu

A Semi-Empirical Model for


Condensation Heat Transfer
Coefficient of Mixed
Ethanol-Water Vapors
A semi-empirical model describing the heat transfer characteristics of the pseudodropwise condensation of binary vapor on a cooled vertical tube has been formulated. By
ignoring the thin film always present on the condensation surface and the intensification
of mass transfer caused by the Marangoni effect, the heat transfer characteristics of
pseudo-dropwise condensation are tentatively formulated. The model involved an analysis of the diffusion process in the vapor boundary layer along with the heat transfer
process through the condensate drops. This model was applied to the condensation of the
saturated binary vapor of ethanol and water, and was examined using experimental data
at vapor pressure values of 101.33 kPa (provided by Utaka and Wang, 2004, Characteristic Curves and the Promotion Effect of Ethanol Addition on Steam Condensation
Heat Transfer, Int. J. Heat Mass Transfer, 47, pp. 45074516), 84.52 kPa and 47.36
kPa. Calculations using the model show a similar trend to the experimental measurements. With the change of the vapor-to-surface temperature difference, the heat transfer
coefficients revealed nonlinear characteristics, with the peak values under all ethanol
mass fractions of binary vapor. The heat transfer coefficients increased with decreasing
ethanol mass fraction. DOI: 10.1115/1.4003433
Keywords: ethanol-water vapors, Marangoni condensation, pseudo-dropwise, heat
transfer, theoretical model

Introduction

It has been known for many years that the condensation of the
binary vapor of miscible liquids does not always occur in a filmwise manner, even on a hydrophilic surface. Pseudo-dropwise
condensation can occur under certain circumstances during the
condensation of binary vapors. The heat transfer characteristics of
the condensation and the nonfilmwise condensation modes of binary vapors have been widely studied for many years.
The phenomenon of nonfilmwise condensation of a binary vapor on the surface of a vertical cylindrical nickel plate was first
described photographically by Mirkovich and Missen 1. Later, a
number of researchers 216 reported that nonfilmwise condensation modes, such as dropwise, rivulet and rippled, could be realized with a binary vapor, as shown in Table 1. The heat transfer
characteristics of the nonfilmwise condensations are also summarized in Table 1.
The mechanism of the pseudo-dropwise condensation of binary
vapors can be explained, in principle. The pseudo-dropwise condensation of binary vapors is significantly influenced by the Marangoni effect, which describes the effect of surface tension gradients developed in the condensate surface owing to local
perturbations in concentration and temperature. If the change in
surface tension with respect to the film thickness is positive, the
film tends to be unstable 17. Hijikata et al. 18 carried out a
theoretical study concerning the dropwise condensation DWC
mechanisms of ethanol-water vapor mixtures by instability analysis, in which the dropwise condensation was generated by the
instability of the condensate film. They pointed out that any de1
Corresponding author.
Contributed by the Heat Transfer Division of ASME for publication in the JOURNAL OF HEAT TRANSFER. Manuscript received October 4, 2009; final manuscript received: January 5, 2011; published online March 2, 2011. Assoc. Editor: Srinivas
Garimella.

Journal of Heat Transfer

formation in the condensate film would be expected to give a


lower surface temperature at the valley and a higher temperature
at the crest of the condensate drop. They also pointed out that the
change in surface tension with temperature was not always negative. It became positive for certain mixtures due to the dependence
on concentration. Pseudo-dropwise condensation was only realized when the surface tension increased with temperature.
It is generally true that mass transfer resistance is formed because the concentration field at the phase interface and the resistance could obstruct the mass flow in the direction of the phase
interface. As a result, the condensation heat transfer coefficient for
equal vapor-to-surface temperature difference is less than that for
a single vapor. However, the heat transfer of pseudo-dropwise
condensation had already been shown to be significantly enhanced
because of the reduction of heat transfer resistance through the
condensate liquid. It has been widely accepted that, notwithstanding the existence of the vapor phase diffusion resistance, condensation heat transfer could be significantly enhanced by adding
very little of the low-boiling component. This heat transfer enhancement is attributed to the Marangoni effect, which produces a
disturbed condensate film at lower low-boiling component mass
fractions where the diffusion layer resistance remains low. Notably, in the latest research of Utaka and Wang 11, such mixtures
were demonstrated to be extremely effective, especially for low
ethanol mass fractions 0.51%. The effect of adding a small
amount of ethanol was remarkable, and the condensation heat
transfer was enhanced eight times compared with that of steam. In
addition, Murase et al. 13,14 investigated the condensation of
ethanol water mixtures on a horizontal tube. The results showed
that significant heat transfer enhancement could be obtained by
adding a very small amount of ethanol. The trends on the horizontal tube were consistent with the results found earlier by Utaka
and Wang 11. Moreover, on vertical plates and tubes, similar
trends in the condensation heat transfer characteristics of ethanol-

Copyright 2011 by ASME

JUNE 2011, Vol. 133 / 061501-1

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 10/02/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Table 1 Main experimental research on binary vapor condensation


Year

Authors

1961 Mirkovich Missen 1


1992
Hijikata 2
1993
Fujii 3
1995
Utaka 4
1997 Morrison Deans 5
1998 Morrison Philpott 6

Mixture component
mass fraction

Pressure

Test surface
shape

C5H1214%-CH3OH
C5H1219%-CH2Cl2
Ethanol466%-water
Ethanol020%-water
Ethanol0.170.71%-water
Ammonia0.230.88%-water
Methylamine0.034.3%-water

atm

Cylindrical plate

Kim 7 Lefsaker

2-ethyl-1-hexanol1%-water

atm

2001

Utaka 8 Wang

Ethanol0.52%-water

atm

2004 Philpott Deans 9,10


Ammonia0.9%-water
2004
Utaka 11 Wang
Ethanol0.0532%-water
2006
Vemuri Kim 12
2-ethoxyethanol0.11.0%-water
2006 Murase Wang 13,14
Ethanol0.595.5%-water
2008

Yan 15 Yang

Ethanol0.550%-water

2008

Li 16 Yan

Ethanol0.550%-water

Vertical plate
Horizontal shell
and tube
condenser
Vertical plate
Horizontal tube
Horizontal tube

atm
atm
atm
atm
Lower than
atm
Lower than
atm

water vapor have been observed, and a significant heat transfer


enhancement has been observed experimentally under lower pressures than atmospheric pressure 15,16.
As for the diffusion process during the condensation of a binary
vapor, many investigations have focused on the diffusion process
in the vapor boundary layer near the vapor-liquid interface during filmwise condensation. The diffusion process results in mass
transfer resistance and the diminution of heat transfer. A pioneering study was conducted by Colburn and Drew 19, using a formulation semi-empirical in nature, and based only on conservation laws. Later, a theoretical model for filmwise condensation
FWC of a binary vapor on a cooled vertical plate was formulated by Sparrow and Marschall 20. The derivation was based on
the law of conservation of mass and other fundamental physical
principles. Philpott and Deans 9,10 established a onedimensional model based on the binary vapor condensation theory
of Colburn and Drew 19. This model predicated the condensation heat transfer coefficient using Nusselts theory, and highlighted the difference between filmwise and nonfilmwise condensation of binary vapors.
Although a great deal of experimental research has been carried
out into the condensation of ethanol-water vapor mixtures, there
are few theoretical reports into the heat transfer in the pseudodropwise condensation of a binary vapor. The objective of this
research is to establish a simple predictive model for the condensation heat transfer coefficient of a binary vapor by assuming that
the vapor-liquid interfacial temperature is unchanged and uniform.
This predictive model is based on an analysis of the heat transfer
processes in the vapor boundary layer and the condensate drops.
In addition, as a necessary auxiliary for development of the
model, the condensation heat transfer characteristics and modes of
ethanol-water vapor were first investigated experimentally.

Irregular, streaky,
dropwise
Horizontal plate
Dropwise
Horizontal tube Drop, streak, ring, film
Vertical plate
Dropwise, rivulets,
Horizontal tube Turbulent banded film
Horizontal tube
Banded, rippled,
droplet
A bundle of
Dropwise
horizontal tubes

atm
atm
atm
atm
atm

2001

Condensation modes

Vertical plate
Vertical tube

Heat transfer coefficient


compared with steam
N/A
Weakened
Weakened
An increase of 200300%
An increase of 13%
An increase of 30%

An increase of 30%
1%, the mass fraction
of low-boiling
component
Dropwise
An increase of 300450%
An increase of 34%
compared with
Nusselt result 0.9%
Turbulent banded film
Dropwise, rivulets
An increase of 800% 1%
Dropwise
An increase of 47% 0.9%
Ripples, dropwise,
An increase of 300350%
wavy film
compared with Nusselt
result 0.58%
Dropwise, rivulets
An increase of 180% 1%
Dropwise, rivulets

An increase of 800% 1%

generator by gravity. The vapor mixture loop was leak-tight, with


a leak rate less than 0.033 Torr h1, to minimize the detrimental
effect of noncondensable gas in the condensation process. Noncondensable gas was continuously extracted from the outlet of the
auxiliary condenser using a vacuum pump. A shell-tube cooler
was used to cool the inlet of the vacuum pump to maintain a
constant vapor mixture. The condensation chamber was maintained at a constant vapor pressure by the flow adjustment of
cooling water II. When the vapor condition reached the steady
state, the condensation characteristic curves were measured continuously via a quasi-static process and the jet of cooling water
was heated up slowly from close to freezing point almost 2 C to
the saturated temperature of the binary vapor. Simultaneously, the
appearance of the condensate mode was recorded using the charge
coupled device CCD camera.
The ethanol mass fraction of the vapor mixtures was determined using the UNIFAC model, as shown in Fig. 11. A program

Experimental Study

The experimental apparatus is shown in Fig. 1. Vapor mixtures,


generated in the vapor generator, were directed vertically through
the condensation chamber where the jet-water-cooled test tube
was fixed. This resulted in almost all the vapor condensing in the
auxiliary condenser. The condensate from both the condensation
chamber and the auxiliary condenser was returned to the vapor
061501-2 / Vol. 133, JUNE 2011

Fig. 1 A schematic diagram of the experimental apparatus

Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 10/02/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 2 Dimensions of the test tube

was developed to calculate the vapor ethanol mass fraction as a


function of vapor pressure from the liquid ethanol mass fraction.
The dimensions of the copper heat transfer tube are shown in
Fig. 2. The heat transfer test tube was a vertical, half tube with a
semi-annular cross section. The condensation heat transfer flux
was obtained from the temperature gradient at the tube wall.
Detailed descriptions of the experimental apparatus, along with
an uncertainty analysis, were provided in the authors earlier publication 16.
To investigate the condensation characteristics of ethanol-water
vapor mixtures on the vertical tube, tests were performed with a
wide range of ethanol vapor mass fractions We = 0.5%, 1.0%,
2.0%, 5.0%, 10.0%, 20.0% under vapor pressures of 47.36 kPa
and 84.52 kPa and a vapor velocity of 2 m/s, as shown in Fig. 3.
As shown in Fig. 3, similar trends for the condensation heat
transfer coefficients were observed for all of the ethanol mass
fractions. As a function of increasing vapor-to-surface temperature
difference, the condensation heat transfer coefficient revealed
nonlinear characteristics, with peak values. This nonlinear characteristic of the condensation heat transfer coefficient was very similar to that found by Utaka and Wang 11 and Murase et al. 14.
The pseudo-dropwise condensation is quite different from the
condensation of a single vapor on a hydrophobic surface. Pseudodropwise and dropwise condensation differ in two major ways.
First, during the Marangoni condensation, it was confirmed by
Utaka and Nishikawa 21, who used a laser light absorption

Fig. 4 Transformation of the condensate modes

method to measure the film thickness during the condensation of


ethanol-water vapor, that a film of thickness of at least 1 m was
always present on the tube surface. Second, the Marangoni condensation modes always varied with the vapor-to-surface temperature difference. During the condensation process, a variety of condensation modes was observed. These condensation modes were
filmwise condensation FWC, rivulet condensation RC, dropwise condensation DWC, and big block condensation BC with
an irregular border. These condensation modes occurred partially
or totally, and sometimes occurred as a mixed condensation mode,
depending on the ethanol vapor mass fraction. At a vapor pressure
of 84.52 kPa and a velocity of 2 m/s, Fig. 4 shows the transformation of the condensation modes during the condensation process for different ethanol mass fractions.
Moreover, depending on the recorded condensate mode, filmwise condensation only occurred at very small vapor-to-surface
temperature differences T 3.5 K or relatively large vapor-tosurface temperature differences with low ethanol mass fractions.
At a vapor pressure of 84.52 kPa and a velocity of 2 m/s, Fig. 5
shows the transformation of the condensation modes with respect
to ethanol mass concentration and vapor-to-surface temperature
difference. This shows that dropwise condensation could be realized over a wide range of vapor mass fractions and vapor-tosurface temperature differences. This transformation of pseudodropwise condensation agreed with the results found by Utaka on
the vertical plate under atmospheric vapor pressure 11.

Fig. 3 Condensation heat transfer coefficients

Journal of Heat Transfer

Model Development

The physical situation chosen for study is gravity-flow pseudodropwise condensation on a cooled, isothermal vertical tube, as
shown in Fig. 6. The tube is situated in the condensation chamber
JUNE 2011, Vol. 133 / 061501-3

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 10/02/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 5 Transformation of the condensation modes with respect to ethanol mass concentration and vapor-to-surface temperature difference

that is filled with a binary mixture of vapor. The condensate,


formed adjacent to the tube surface, is a binary mixture of the
liquid phases of the condensing vapors. The condensate flows
downward along the tube surface under the action of gravity. The
bulk vapor is assumed to be in the saturated state corresponding to
the ambient ethanol mass fraction We and pressure p. The
specification of the ambient ethanol mass fraction We and pressure p, taken in conjunction with the vapor-liquid phase equilibrium diagram, fixes the ambient temperature T of the bulk vapor. The plate wall temperature Twall is also specified, along with
T and p.
As mentioned in Sec. 1, there are two main heat transfer resistances in the process of binary vapor condensation: the resistance
due to the vapor boundary layer and the resistance due to the
condensate liquid. Because of the concentration gradient existing
in the vapor boundary layer, the heat transfer is affected by the
mass transfer. The transport process in the two regions should be
coupled by conditions of continuity, conservation, and equilibrium
at the interface. However, because of the complex process of the
Marangoni condensation of the binary vapor, the interfacial conditions are complex. The interface is not smooth, since the crest of
the condensate drop has a higher temperature than the valley.
Moreover, the mass transfer through the vapor-liquid interface
was intensified because of the Marangoni effect, and this intensification was hard to analyze qualitatively and quantitatively.

Fig. 6 Schematic diagram for pseudo-dropwise condensation


of binary vapor

061501-4 / Vol. 133, JUNE 2011

Therefore, the interfacial conditions of pseudo-dropwise condensation cannot be solved by use of theoretical analysis at this time.
In this research, based on the fact that the pseudo-dropwise
condensation resulted from the deformation of the condensation
film, an assumption was made that the interfacial temperature of
the condensate drops was uniform and equaled the interfacial temperature of the smooth condensate film of the binary vapor under
the same ambient conditions Twall, T, and p. In other words,
dropwise condensation of binary vapor was simply caused by the
interfacial deformation of the condensate film in the filmwise condensation of binary vapor. Therefore, the assumed interfacial temperature Ti can be taken as the interfacial temperature in the
dropwise condensation of a single vapor. Similarly, during the
dropwise condensation analysis for a single vapor, the vaporliquid interfacial temperature of the condensate drop is also assumed to be uniform 22. Although not strictly realized in practice, this simplifying assumption has been used in the dropwise
condensation analysis for a single vapor and has yielded satisfactory heat transfer predictions 22. Based on this hypothesis, if the
assumed interfacial temperature was calculated, the condensation
heat transfer coefficient could be solved by using the dropwise
condensation method of the pure vapor. The simplifying assumptions of this model are as follows:

The bulk vapor is saturated.


The interface between the condensate liquid and the vapor is
saturated.
The shear stress on the condensate liquid caused by the vapor is negligible.
The pressure drop at the interface is negligible.
The condensate drop radius is small compared with the tube
diameter.
The contact angle of the condensate drop is 90 deg.
The latent heat liberated by the condensing binary vapor is
transferred by heat conduction across the condensate drop
along with the vapor-liquid interfacial resistance. The convective heat transfer in the drops is negligible.
The intensification of mass transfer through the vapor-liquid
interface caused by the Marangoni effect is neglected.
The ethanol mass fraction in the condensate liquid is uniform.
The liquid film of thickness of at least 1 m that always
exists on the condensation surface is ignored.

First, the heat resistance of condensate drops is discussed below.


3.1 Pseudo-Dropwise Condensate Drops. Although the
pseudo-dropwise condensation of a binary vapor was different
from the dropwise condensation of a single vapor, the Marangoni
condensation still had the majority of the dropwise condensation
characteristics of the single vapor. Despite the changing modes of
the Marangoni condensation, the pseudo-dropwise condensation
occurred at a wide range of vapor-to-surface temperature difference for various ethanol mass fractions, as indicated in Fig. 5.
3.1.1 Size Distribution of Drops in Dropwise Condensation. If
the pseudo-dropwise condensation was treated approximately as
the dropwise condensation of a single vapor, the heat transfer
process through the condensate drops could be solved by using the
dropwise condensation method. To illustrate the appropriateness
of this approximate treatment, the condensate drop size distribution of the pseudo-dropwise condensation was investigated for
various operating conditions in the pseudo-dropwise condensation
area, as shown in Fig. 7. The drop size distribution is the fraction
of the surface area covered by drops having a base radius greater
than r. The drop size distribution of the pseudo-dropwise condensation was counted from the condensation mode images recorded
by a CCD camera during the experimental process, as shown in
Fig. 4. With the change of condensation modes, the radius of the
Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 10/02/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

surface is lower than that of a planar surface at the same


vapor pressure. The temperature differences for the drop
curvature of a condensate drop are 23
2Ti
rh fgL

Tc =

where Ti is the assumed interfacial temperature, to be


discussed in Sec. 3.2, is the surface tension, h fg is the
latent heat of vaporization, L is density of the condensate, and r is the drop radius.
The temperature decrease due to the vapor-liquid interfacial resistance for a hemispherical drop is 24
Ti =

Fig.
7 Drop
condensation

size

distribution

of

pseudo-dropwise

r = 2A/P

where A is the area of the condensate block and P is the perimeter


of the condensate block.
Figure 7 shows a comparison between the drop size distribution
of the pseudo-dropwise condensation and that of the dropwise
condensation of steam, as suggested by Rose 22. The trend of
the counted drop size distribution was apparently different. The
counted drop size distribution was convex, and the Roses drop
size distribution was sunken. One reason for these differences
could be that condensate drops less than 0.1 mm could not be
discriminated by the artificial perception of the condensation
mode images.
However, Fig. 7 does indicate that the size distribution of the
pseudo-dropwise condensation follows fairly close to that given
by Rose distribution. Therefore, within the pseudo-dropwise area
see Fig. 5, the heat transfer characteristics of the pseudodropwise condensation could be tentatively formulated, based on
this approximate treatment.
It should be noted that drops grow during condensation and the
drops on the condensation surface are swept periodically when the
drops reach the critical size. During each refresh cycle, the drop
size distribution is therefore continuously changing, and condensation exhibits a strong dynamics feature. Over a long period,
however, a quasi-steady state is reached and the above measurement of drop size distribution can be thought as representative of
average of such a quasi-steady behavior. In addition, for different
vapor-to-surface temperature differences, the drop size distribution might be different. The drop size distribution of pseudodropwise condensation needs to be explored further.
3.1.2 Heat Transfer Through a Single Drop. By assuming that
the shape of the drop was hemispherical and ignoring the bottom
film always present on the condensation surface during the Marangoni condensation, all the resistances to heat transfer through
a single drop can be expressed as temperature differences, namely,
the temperature differences of the drop curvature of the condensate drop, the vapor-liquid interfacial temperature differences for a
hemispherical drop, and the temperature differences of heat conduction through a hemispherical drop.
a

The equilibrium saturation temperature of a curved liquid

Journal of Heat Transfer

qb
2 r 2h i

where qb is the heat flux through a drop and hi is the


interfacial heat transfer coefficient 25.
The temperature differences of the heat conduction
through a hemispherical drop are 26
Td =

largest drop became bigger and bigger, and then grew into a big
liquid block with an irregular border or rivulet, at which point the
radius of the condensate drop was replaced by an equivalent radius, as shown in Fig. 4. The equivalent radius of the liquid block
is defined as

q br
4 r 2 L

where L is the thermal conductivity of the condensate.


According to the total temperature difference between the vapor
and the condensation surface T = Tc + Ti + Td, determined
by combining Eqs. 24, the heat flux through the base of a
hemispherical drop with radius r can be obtained as

2Ti
rLh fg
2
r
+
L hi

4r2 T
qb =

where T is the interface-to-surface temperature difference


T = Ti Twall.
The drop size distribution of a single vapor is used as the
pseudo-dropwise condensation 22.
Nrdr =

1 1
r
3r2 rmax rmax

2/3

dr

Combining Eqs. 5 and 6, the average heat flux for the surface
was obtained by integration over all drop radii
q=

rmax

qb r2 Nrdr

rmin

The condensation heat transfer coefficient for dropwise condensation of the binary vapor can be calculated directly from
h = q/T Tw

It is important to discuss the radius of the smallest and largest


drops first.
3.1.3 Smallest Condensate Drop. For the pseudo-dropwise
condensation, the minimum possible radii of drops that can possibly grow have remained unknown theoretically or experimentally until now. For steam condensation, based on optical and
electron microscope photographs, Graham and Griffith gave
nucleation site densities of 2 108 cm2 and 6 108 cm2 for
condensation where the minimum radius was 0.07 m 27. Tanasawa indicated a site density exceeding 1010 cm2 when the
minimum radius was 0.01 m 25. Based on thermodynamical
analysis, Rose suggested that the radius of the smallest possible
condensation drop was 22
rmin =

2Ts
h fgLT

JUNE 2011, Vol. 133 / 061501-5

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 10/02/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Table 2 Radius of smallest drop under different experimental conditions


47.36 kPa

U=2 m/s

84.52 kPa

U=2 m/s

101.33 kPa

U=2 m/s

We
%

Smallest drop radius


m

We
%

Smallest drop radius


m

We
%

Smallest drop radius


m

0.5
1
2
5
10
20

0.095
0.02
0.14
0.195
0.35
0.35

0.5
1
2
5
10
20

0.12
0.09
0.1
0.165
0.29
0.45

0.5
1
3
6
12
21
32

0.07
0.055
0.05
0.035
0.075
0.15
0.28

where Ts is the saturated temperature of single vapor.


According to Eq. 9, at a pressure of 101.33 kPa and with the
vapor-to-surface temperature difference of 5 K, the radius of
smallest drop was 0.001 m. Rose also pointed out that agreement with the experimental data was best when using a radius for
the smallest drop of 10 rmin 25.
The nucleation mechanisms of pseudo-dropwise condensation
are different from pure vapor condensation. For the pure vapor
condensation, liquid drops nucleate on a solid surface and the
smallest drop size is limited only by thermodynamic requirement.
In the case of a binary mixture, as analyzed by Hijikata et al. 18,
drops are formed due to instability of the condensate film on the
solid surface. The size of the smallest drops should be limited by
the wavelength of the unstable perturbations at the vapor-liquid
interface for pseudo-dropwise condensation of a binary mixture,
not the thermodynamic limit of nucleation. Since there is no existing theory that can be used to determine the radius of the smallest drops, rmin is then left in the model as free parameter that may
be estimated from the comparison between the model predictions
and experimental measured heat transfer data. By using three radii
of minimum drops for the calculation of heat transfer through
drops and the lower limit of integration in Eq. 7, the radius of
the smallest drop that best agreed with the experimental data was
determined. Detailed discussion refers to Sec. 4.3. The determinate radii of the smallest drop were shown in Table 2.
3.1.4 Largest Condensate Drop. For the condensation of a
single species vapor, the radius of the largest drop is taken as 22

rmax = 0.4

Lg

1/2

10

At a vapor pressure of 84.52 kPa and a velocity of 2 m/s, the


comparison of the radius of the largest drop calculated by Eq. 10
with that observed experimentally is shown in Fig. 8. It can be see
that, with increasing vapor-to-surface temperature difference, the
radius of the largest drop calculated by Eq. 10 was approximately constant. With a low vapor-to-surface temperature, the results of Eq. 10 are approximately equal to the largest radius
observed in the experiment. However, with a high vapor-tosurface temperature, the results of Eq. 10 deviated significantly
from the observed largest radius. One reason for this could be the
difference between the pseudo-dropwise and dropwise condensation of a single vapor. The pseudo-dropwise condensation is
caused by the surface tension difference and the radius of the
largest drop became bigger and bigger with increasing vapor-tosurface temperature difference for various ethanol mass fractions.
Taking into account the surface tension difference and the observed radius of the largest drop, the radius of the largest drop is
suggested as

rmax
= KPrL/PrL0n

d/dTiTi Twall
Lg

3.2 Vapor Boundary Layer. The heat transfer resistance of


the diffusion boundary layer is important to binary vapor condensation. The assumed interfacial temperature Ti can be obtained
by the filmwise condensation theory of a binary vapor 20. The
Sparrow model is based on an analysis of the transport processes

1/2

11

where d / dTi expresses the strength of the variation of surface


tension with temperature, which describes the driving force for
061501-6 / Vol. 133, JUNE 2011

pure surface tension driven convection. Ti is the assumed interfacial temperature, which will be discussed in Sec. 3.2. PrL is the
Prandtl number of the condensate liquid and PrL0 is evaluated at
the onset of condensation the vapor-to-surface temperature was
0.1 K. K is an adjustable constant, which is used to consider the
effect of the velocity of vapors in future research. In the present
model, due to vapor velocity was not taken into account, K was
chosen as 1. n is the correction factor. Depending on the maximum vapor-to-surface temperature difference and the corresponding equivalent largest radius of the big condensate block, the correction factor n in Eq. 11 could be determined. Table 3 shows
the values of n for different experimental conditions. It should be
noted that, for the experiment of Utaka and Wang 11, the equivalent largest radius of the big condensate block were not available.
The equivalent largest radius of the big condensate block can be
obtained with this approximate approach as follows: with the increase in the vapor to surface temperature difference, when the
aspects of the condensate shift again to smooth film, the equivalent largest radius of the smooth film is regarded as 35 mm, half of
the test block width.
Figure 9 shows the comparison between the largest drop calculated by Eq. 11 and the observed radius in the experiment for a
vapor pressure of 84.52 kPa and a velocity of 2 m/s. It was found
that the calculated radius of the largest drop was in good agreement with the experimental observation, as shown in Fig. 10. The
maximum error was about 60%. In addition, about 80% of the
error was lower than 30%.

Fig. 8 Comparison of Eq. 10 with the observed radius of the


largest drop

Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 10/02/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Table 3 Value of the n for different experimental conditions


P = 47.36 kPa

U = 2 m s1

P = 84.52 kPa

U = 2 m1

P = 101.33 kPa

U = 2 m1

We
%

We
%

We%

0.5
1.0
2.0
5.0
10.0
20.0

8.88
7.62
5.85
3.55
2.25
2.95

0.5
1.0
2.0
5.0
10.0
20.0

11.52
9.86
7.55
4.52
3.20
3.64

0.5
1.0
3.0
6.0
12.0
21.0
32.0

11.56
9.97
7.77
4.72
3.48
3.90

in the condensate film and in the vapor boundary layer, along with
conditions of continuity, conservation, and equilibrium at the
liquid-vapor interface. More detailed discussion of the calculation
of the interfacial temperature Ti can be found in the paper by
Sparrow and Marschall 20.
The liquid-vapor phase equilibrium diagram for water and ethanol is the prime resource for the calculation of the assumed interfacial temperature. Because of the lack of water-ethanol phase
equilibrium data for the pressures studied 47.36 kPa, 84.52 kPa,
and 101.33 kPa, the phase equilibrium data used to construct the
curves were calculated by the UNIFAC model 28. Figure 11

shows the liquid-vapor phase equilibrium relations for water and


ethanol at pressures of 47.36 kPa, 84.52 kPa, and 101.33 kPa. The
diagram was used for finding the temperature in the saturated
vapor, saturated liquid, and saturated interface, corresponding to a
given mass fraction of the ethanol.
The properties of the vapor and the liquid are to be evaluated at
appropriate reference states, taking into account that both the vapor and the liquid are binary mixtures. As a necessary prelude to
the solution, the reference state with the relevant thermodynamic
and transport properties will be described in the Appendix.

Results and Discussion

4.1 Assumed Interfacial Temperature. According to the interfacial temperature approach for filmwise condensation of binary vapor determined by Sparrow at a vapor pressure of 84.52
kPa, the assumed interfacial temperature is shown in Fig. 12. The
interfacial temperature decreased and then remained unchanged
with increasing vapor-to-surface temperature difference. When the
interfacial temperature was constant, which should be the lowest
interfacial temperature, the vapor-to-interface temperature difference coincided with the bubble and dew point temperature difference. In the low ethanol mass fraction range, because of the
bubble and dew point temperature difference was very small, the
decreasing region of the interfacial temperature was very small.

Fig. 9 Comparison of Eq. 11 and observed radius of the largest drop

Fig. 10 The error of Eq. 11 and observed radius of the largest


drop

Journal of Heat Transfer

4.2 Theoretical Condensation Heat Transfer Coefficient.


The theoretical results are presented in Figs. 1315 for vapor pressures of 101.33 kPa, 84.52 kPa and 47.36 kPa, respectively.
As shown in Figs. 1315, a similar trend was found between
the theoretical results and the corresponding experimental data
11,16. Similar to the trend of the experimental condensation heat

Fig. 11 Liquid-vapor phase equilibrium diagrams for water


and ethanol

JUNE 2011, Vol. 133 / 061501-7

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 10/02/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 12 Assumed interfacial temperature

transfer coefficient, the theoretical condensation heat transfer coefficient revealed nonlinear characteristics and had peak values
with respect to the vapor-to-surface temperature difference for all
studied vapor ethanol mass fractions. For low ethanol mass fractions We = 0.5% 3%, the theoretical heat transfer coefficient
increased immediately and rapidly, reached its peak value, and
then decreased with increasing vapor-to-surface temperature difference. For high ethanol mass fractions We = 5% 32%, with
increasing vapor-to-surface temperature difference, the theoretical
heat transfer coefficient increased slowly, and then began to reach
its peak value relatively fast and then decreased. This trend indicated that the diffusion resistance in the vapor boundary layer
played a significant role before the vapor-to-surface temperature
reached the steep increasing point. Furthermore, Figs. 1315 also
indicate that the theoretical heat transfer coefficient decreased
with the increase of the ethanol mass fractions. In addition, with
the change of vapor-to-surface temperature difference, the ampli-

Fig. 15 Model condensation heat transfer coefficient p


= 47.36 kPa

tude of variation of the theoretical heat transfer coefficient became


greater for low ethanol mass fractions than those for high ethanol
mass fractions.
As for the steep increasing points of the heat transfer coefficients, because of the very small temperature difference between
the bubble and dew point, there were no apparent steep increase
points of the theoretical and experimental coefficients for low ethanol mass fractions. However, for high ethanol mass fractions,
similar to the experimental results, there were obvious steep increasing points. The vapor-to-surface temperature differences of
the steep increasing points correspond to the lowest interfacial
temperature. The lowest interfacial temperature, subjected to the
ambient temperature T and pressure p, was essentially constant
and independent of Twall, and occurred when the vapor-to-surface
temperature difference reached the steep increasing point of the
heat transfer coefficient, as shown in Fig. 12.
Both the theoretical and experimental coefficients had obvious
peak values. With increasing vapor-to-surface temperature difference, the theoretical and experimental coefficient increased,
reached its peak value, and then decreased. To illustrate the reason
why the heat transfer coefficients had a peak value, the heat transfer resistance through the vapor boundary layer and the condensate drops is discussed for a vapor pressure of 84.52 kPa and an
ethanol mass fraction of 10%, as shown in Fig. 16. Rv f represent
the thermal resistance of the vapor boundary layer and the Rc
represent the thermal resistance of condensate drops. Rv f and Rc
can be calculated from
Rv f =

T Ti
q

12

Fig. 13 Model condensation heat transfer coefficient p


= 101.33 kPa

Fig. 14 Model condensation heat transfer coefficient p


= 84.52 kPa

061501-8 / Vol. 133, JUNE 2011

Fig. 16 Thermal resistance analysis of the Marangoni


condensation

Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 10/02/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 17 Sensitivity analysis of the heat transfer calculation to


the radius of the smallest drop We = 0.5%

Rc =

Ti Tw
q

13

Within area I, Rv f was relatively large and far greater than Rc, and
Rv f played important roles in the heat transfer. The heat transfer
coefficient was relatively small. Within area II, Rv f decreased
quickly and Rc moved up slowly. Both Rv f and Rc were small at
this time, and the heat transfer coefficient reached its peak value.
For area III, Rv f decreased to its lowest value and was almost
constant, while Rc continued to increase. Rc was the dominant
resistance, and greater than Rv f . The heat transfer coefficient then
decreased.
4.3 Sensitivity Analysis of the Smallest Radius of a Drop.
The sensitivity of the heat transfer calculation to the smallest radius of drop can be assessed by evaluating the integral in Eq. 7.
At vapor pressures of 101.33 kPa, 84.52 kPa, and 47.36 kPa, and
vapor ethanol mass fractions of 0.5% and 10%, the three values of
the radius of smallest drop were selected to calculate the heat
transfer coefficient. According to the results, for a vapor ethanol
mass fraction of 0.5%, when the peak value differences between
the calculation and experiment were smallest, the radius of the
smallest drops ranged from 0.07 m to 0.12 m, as shown in
Fig. 17.
For a vapor ethanol mass fraction of 0.5%, the radius of smallest drop was almost at the same value no matter for which experimental system or how great is the vapor pressure.
For a vapor ethanol mass fraction of 10%, when the peak value
differences between the calculation and experiment were smallest,
the radius of the smallest drop ranged from 0.08 m to 0.4 m,
as shown in Fig. 18.
Under different experimental conditions, Table 2 shows the radius of the smallest drop obtained by sensitivity analysis. With
Journal of Heat Transfer

Fig. 18 Sensitivity analysis of the heat transfer calculation to


the radius of the smallest drop We = 10%

increasing ethanol mass fraction of the vapor, the radius of smallest drop became bigger and bigger. With decreasing vapor pressure, the radius of the smallest drop tended to increase.
However, the radius of the smallest drop might change for a
different vapor-to-surface temperature difference. In the present
model, the radius of the smallest drop was estimated at the vaporto-surface temperature difference where the heat transfer coefficient reached the peak value. Moreover, the estimated radius of
the smallest drop was considered invariable throughout the calculation. This is the main reason for the significant discrepancies
between the predictions and experiments under several ranges of
the vapor-to-surface temperature difference, as shown in Figs. 17
and 18. The smallest drop of the pseudo-dropwise condensation
needs further investigation by a microcosmic method.
Because of the simplification of the theoretical model for the
dropwise condensation for ethanol-water vapor, it was not surprising that differences were seen between the calculation and experiment. The theoretical model found it hard to predict the value of
the heat transfer coefficients. The comparison between the calculation and experiment indicated that the present theoretical model
led to an appropriate description of the heat transfer process in the
vapor boundary layer and condensate drops during the Marangoni
condensation of a binary vapor.

Conclusion

A theoretical model has been developed to predict the pseudodropwise condensation of a binary vapor. Given the primary assumption of an unchanged and uniform interfacial temperature
during the transformation from filmwise to dropwise condensation, this model combined a theoretical analysis of the diffusion
process in the vapor boundary layer along with a theoretical calculation of the dropwise condensation for a single vapor. The
JUNE 2011, Vol. 133 / 061501-9

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 10/02/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

model took into account both the heat transfer resistance of the
vapor boundary layer and heat transfer resistance of the condensate liquid.
The model predicted well the general trend of the experimental
data. Similar to the experimental condensation curves of the heat
transfer coefficients, the theoretical heat transfer coefficients revealed nonlinear characteristics and had peak values with respect
to the vapor-to-surface temperature difference for all ethanol mass
fractions and all calculation conditions. Moreover, with increasing
ethanol mass fraction, the theoretical heat transfer coefficient decreased. The discrepancies between the calculation and the experiment were chiefly due to simplification of the predictive model.
Compared with the experiment, the theoretical model gives a
more fundamental description of the mechanism for pseudodropwise condensation of a binary vapor.
Moreover, this model ignores several important factors regarding pseudo-dropwise condensation of binary vapors, such as vapor
velocity, the liquid film always being present on the condensation
surface and the intensification of mass transfer through the vaporliquid interface caused by the Marangoni effect. The more precise
model deserves further investigation. The new model should be
concerned with the mechanism of pseudo-dropwise condensation,
namely, a stability analysis for the condensation of liquid films of
binary vapors.

Acknowledgment
The authors gratefully acknowledge financial support from the
National Natural Science Foundation of China Grant Nos.
50476048 and 50521604. The authors also thank Professor
Yoshio Utaka, Division of Systems Research, Faculty of Engineering, Yokohama National University, Japan, for helpful assistance.

TV = T + Ti/2

A6

The viscosity of the single vapor was calculated by the method of


Chung et al. 33, and the viscosity of the binary vapor was obtained by using Wilkes rule 34.
xeVeV
xwVwV
+
xeV + xwV12 xwV + xeV21

A7

1 + eV/wV1/2M w/M e1/42


81 + M e/M w1/2

A8

V =
where

12 =

21 = 12wV/eVM e/M w

A9

where eV and wV are the viscosities of the pure component


vapors.
The latent heat h fg was expressed as 20
h fg = h fgeWeV + h fgwWwV

A10

where h fge and h fgw are the viscosities of the pure component
vapors.
The latent heat was evaluated at the interfacial temperature
Ti .
The density of the binary vapor was determined by the Gibbs
Dalton law from the pure component densities.
The mass diffusion coefficient was evaluated using Hirschfelders equation 35.
DAB =

1 1/2
0.001858TV 3/2M 1
e + Mw
2
pAB
D

A11

The mixture Schmidt number Sc= V / V / DAB followed directly


from the viscosity, density, and mass diffusion coefficient of the
binary vapor.

References
Appendix: Property Data
All condensate liquid property data were evaluated at the liquid

temperature TL and mass fraction WeL


. The reference liquid temperature was estimated using Eq. A1 20, while the ethanol
mass fraction of the condensate liquid was taken to be equal to the
interface liquid ethanol mass fraction Eq. A2 20.
TL = Tw + 1/3Ti Tw

A1

= WeiL
WeL

A2

The viscosity of the condensate liquid L was estimated by a


method suggested by Fujii 29.
The thermal conductivity of the mixture was evaluated from a
rule given by Lucas 30.
2

2 1/2
L = WeL
eL + 1 WeL
wL

A3

where eL and wL are the thermal conductivities of the pure component liquids.
The mixture density L corresponding to given mass fractions
WeL and WwL was calculated from the expression 31

L = WeL/eL + WwL/wL1

A4

where eL and wL are the densities of the pure component liquids.


For aqueous solutions, such as ethanol-water, the surface tension could be calculated by the method of Tamura et al. 32. The
surface tension was evaluated at the interfacial temperature Ti .
All vapor boundary layer properties were evaluated at the reference temperature TV and mass fraction WeV of vapor boundary layer. The reference state for the vapor was defined in accordance with Eqs. A5 and A6 20.
WeV = We + Wei/2
061501-10 / Vol. 133, JUNE 2011

A5

1 Mirkovich, V. V., and Missen, R. W., 1961, Study of Condensation of Binary


Vapors of Miscible Liquids, Can. J. Chem. Eng., 39, pp. 8687.
2 Hijikata, K., Nakabeppu, O., and Fukasaku, Y., 1992, Condensation Characteristics of a Water-Ethanol Binary Vapor Mixture, Proceedings of the 29th
Japan Heat Transfer Symposium, pp. 742743.
3 Fujii, T., Osa, N., and Koyama, S., 1993, Free Convection Condensation of
Binary Vapor Mixtures on a Smooth Horizontal Tube: Condensing Mode and
Heat Transfer Coefficient of Condensate, Proceedings of U.S. Eng. Found.
Conference, pp. 171182.
4 Utaka, Y., 1995, Measurement of Condensation Characteristic Curves for
Binary Mixture of Steam and Ethanol Vapor, Heat Transfer-Jpn. Res., 241,
pp. 5767.
5 Morrison, J. N. A., and Deans, J., 1997, Augmentation of Steam Condensation Heat Transfer by Addition of Ammonia, Int. J. Heat Mass Transfer, 40,
pp. 765772.
6 Morrison, J. N. A., Philpott, C., and Deans, J., 1998, Augmentation of Steam
Condensation Heat Transfer by Addition of Methylamine, Int. J. Heat Mass
Transfer, 41, pp. 36793683.
7 Kim, K. J., Lefsaker, A. M., Razani, A., and Stone, A., 2001, The Effective
Use of Heat Transfer Additives for Steam Condensation, Appl. Therm. Eng.,
21, pp. 18631874.
8 Utaka, Y., and Wang, S. X., 2001, Effect of Ethanol Mass Fraction on Condensation Heat Transfer Characteristics for Water-Ethanol Binary Vapor Mixture, Transactions of the Japan Society of Refrigerating and Air Conditioning
Engineers, 181, pp. 127134.
9 Philpott, C., and Deans, J., 2004, The Enhancement of Steam Condensation
Heat Transfer in a Horizontal Shell and Tube Condenser by Addition of Ammonia, Int. J. Heat Mass Transfer, 47, pp. 36833693.
10 Philpott, C., and Deans, J., 2004, The Condensation of Ammonia-Water Mixtures in a Horizontal Shell and Tube Condenser, ASME J. Heat Transfer,
126, pp. 527534.
11 Utaka, Y., and Wang, S. X., 2004, Characteristic Curves and the Promotion
Effect of Ethanol Addition on Steam Condensation Heat Transfer, Int. J. Heat
Mass Transfer, 47, pp. 45074516.
12 Vemuri, S., Kim, K. J., and Kang, Y. T., 2006, A Study on Effective Use of
Heat Transfer Additives in the Process of Steam Condensation, Int. J. Refrig.,
29, pp. 724734.
13 Murase, T., Wang, H. S., and Rose, J. W., 2006, Marangoni Condensation of
Steam-Ethanol Mixtures on a Horizontal Tube, The 17th International Symposium on Transport Phenomena, Sept. 48, Toyama, Japan.
14 Murase, T., Wang, H. S., and Rose, J. W., 2007, Marangoni Condensation of

Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 10/02/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

15
16
17
18
19
20
21
22
23
24

SteamEthanol Mixtures on a Horizontal Tube, Int. J. Heat Mass Transfer,


50, pp. 37743779.
Yan, J. J., Yang, Y., Hu, S. H., Zhen, K., and Liu, J. P., 2007, Effects of Vapor
Pressure/Velocity and Concentration on Condensation Heat Transfer for
SteamEthanol Vapor Mixture, Heat Mass Transfer, 441, pp. 5160.
Li, Y., Yan, J. J., Qiao, L., and Hu, S. H., 2008, Experimental Study on the
Condensation of Ethanol-Water Mixtures on Vertical Tube, Heat Mass Transfer, 445, pp. 607616.
Ford, J. D., and Missen, R. W., 1968, On the Conditions for Stability of
Falling Films Subject to Surface Tension Disturbances; the Condensation of
Binary Vapors, Can. J. Chem. Eng., 46, pp. 309312.
Hijikata, K., Fukasaku, Y., and Nakabeppu, O., 1996, Theoretical and Experimental Studies on the Pseudo-Dropwise Condensation of a Binary Vapor Mixture, ASME J. Heat Transfer, 118, pp. 140147.
Colburn, A. P., and Drew, T. B., 1937, The Condensation of Mixed Vapors,
Trans. Am. Inst. Chem. Eng., 33, pp. 197215.
Sparrow, E. M., and Marschal, E., 1969, Binary, Gravity-Flow Film Condensation, ASME J. Heat Transfer, 91, pp. 205211.
Utaka, Y., and Nishikawa, T., 2003, Measurement of Condensate Film Thickness for Solutal Marangoni Condensation Applying Laser Extinction Method,
J. Enhanced Heat Transfer, 102, pp. 119130.
Rose, J. W., 2002, Dropwise Condensation Theory and Experiment: A Review, Proc. Inst. Mech. Eng., Part A, 2162, pp. 115128.
LeFevre, E. J., and Rose, J. W., 1966, A Theory of Heat Transfer by Dropwise
Condensation, Proceedings of the Third International Heat Transfer Conference, AICHE, New York, Vol. 2, pp. 362373.
Wen, H. W., and Maa, J. R., 1976, On the Heat Transfer in Dropwise Con-

Journal of Heat Transfer

densation, Chem. Eng. J., 12, pp. 225231.


25 Rose, J. W., 1981, Dropwise Condensation Theory, Int. J. Heat Mass Transfer, 24, pp. 191194.
26 Fatica, N., and Katz, D. L., 1949, Dropwise Condensation, Chem. Eng.
Prog., 4511, pp. 661674.
27 Graham, C., and Griffith, P., 1973, Drop Size Distributions and Heat Transfer
in Dropwise Condensation, Int. J. Heat Mass Transfer, 16, pp. 337346.
28 Fredenslund, A., Jones, R. L., and Prausnitz, J. M., 1975, Group Contribution
Estimation of Activity Coefficients in Nonideal Liquid Mixtures, AIChE J.,
216, pp. 10861099.
29 Fujii, T., 1991, Theory of Laminar Film Condensation, Springer-Verlag, New
York, pp. 190207.
30 Lucas, K., 1993, Methods for Calculating the Properties of Materials, VDI
Heat Atlas, VDI Verlag, Dsseldorf.
31 Minkowycz, W. J., and Sparrow, E. M., 1966, Condensation Heat Transfer in
the Presence of Noncondensables, Interfacial Resistance, Superheating, Variable Properties, and Diffusion, Int. J. Heat Mass Transfer, 9, pp. 11251144.
32 Tamura, M., Kurata, M., and Odani, H., 1955, Practical Method for Estimating Surface Tension of Solutions, Bull. Chem. Soc. Jpn., 28, pp. 8388.
33 Chung, T. H., Mohammad, A., Lloyd, L. L., and Kenneth, E. S., 1988, Generalized Multiparameter Correlation for Nonpolar and Polar Fluid Transport
Properties, Ind. Eng. Chem. Res., 274, pp. 671679.
34 Wilke, C. R., 1950, A Viscosity Equation for Gas Mixtures, J. Chem. Phys.,
18, pp. 517519.
35 Hirschfelder, J. O., Curtiss, C. F., and Bird, R. B., 1964, Molecular Theory of
Gases and Liquids, Wiley, New York.

JUNE 2011, Vol. 133 / 061501-11

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 10/02/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Você também pode gostar