Você está na página 1de 22

9.1.

Basic Processes of Condensation


By analogy with the process of evaporation, liquid may form in one of three ways
corresponding to the existence of an unstable, metastable or stable equilibrium state. Let
us briefly look at each one of these to understand the condensation process. In practical
engineering design of heat exchange equipment the stable condensation situation needs
to exist.
Consider a liquid drop of radius, r*, in equilibrium with its surrounding vapor at a
system temperature, , and pressure, . The vapor pressure, under equilibrium
conditions, is higher than the vapor pressure, , for a planar surface, and this difference
is given by

where R is the vapor gas constant and is the surface tension between the liquid and
vapor. With the local condition of mechanical equilibrium for the liquid droplet and its
pressure, ,

one can find the liquid pressure in the droplet as

One can also use the Clausius-Clapeyron relation to calculate for this simple situation
the saturation temperature,
, of the droplet above the vapor temperature for
maintaining this equilibrium

Analogous to boiling, the rate of nucleation of these liquid droplets depends on whether
one considers homogeneous nucleation or heterogeneous nucleation processes. For
homogeneous nucleation the rate expression, dn/dt, is quite similar to that for boiling,

where r* can be found either from Eq.2 or 3. The term,


molecules per unit volume and is a collision frequency
by

, is the number of vapor


for vapor collisions given

where m is the mass of one vapor molecule. One should note that in a similar fashion to
boiling this nucleation rate is altered if it occurs on solid surfaces since the work
required to form a critical size nuclei (r*) is reduced due to wetting of the solid surface.
Now in reference to the more stable situations of a vapor condensing on a planar surface
covered by its own liquid one must consider the local mass transfer situation. Consider a
pure saturated vapor at a pressure, , and a temperature, , condensing on its own
liquid phase whose surface temperature is . The phenomenon of such an interface
mass transfer can be viewed from the standpoint of kinetic theory as a difference
between two quantities; a rate of arrival of molecules from the vapor space towards the
interface and a rate of departure of molecules from the surface of the liquid into the
vapor space. When condensation takes place the arrival rate exceeds the departure rate.
During evaporation the reverse occurs, and during an equilibrium the two rates are equal
and there is no net mass transfer.
From kinetic theory it can be shown that, in a stationary container of molecules, the
mass rate of flow (of molecules) passing in either direction (to right or left) through an
imagined plane is given by

where
= flux of molecules (mass per unit time per unit area) M =molecular weight R
= universal gas constant P and T = pressure and temperature related by the saturation
line. Equation 7 is the starting point for many theories of interfacial phase change.
In general it can be stated that the net molecular flux through an interface is the
difference between these fluxes in the directions from gas to liquid and vice-versa,

Since the condition close to the surface is not one of static thermal equilibrium, for any
significant rate of evaporation or condensation, it is really not meaningful to make use
of the thermostatic pressure and temperature on each side of the interface. Rather there
is a concentration and therefore, a temperature difference,
, across this interface
which drives the mass transfer. Strictly speaking one should solve the Boltzman
transport equation with appropriate boundary conditions and asymptotes which are
conditions of thermal equilibrium at several mean free path distances from the interface.
However, some considerable success for engineering purposes has been achieved by
using simplified kinetic theory techniques and applying correction factors to the
resulting predictions of this mass transfer and the associated temperature difference. In
most practical situations the energy removal rate from this interface controls the
condensation rate. Only in the presence of noncondensable gases (continuum) or at low
pressure (non-continuum) is this temperature difference,
, important to consider. We
will investigate the case of the presence of noncondensable gases during condensation;
one can get a physical feeling of the magnitude of this temperature difference.

9.2. Theoretical Developments of


Condensation
Stationary Pure Vapor
For filmwise condensation of a "stationary" saturated vapor, Nusselt (1916) presented
the first analytical solution for heat transfer on a plane surface (Fig.9.1 ) with the
following assumptions (Collier, 1981):
1. the flow of condensate in the film is laminar,
2. the fluid properties are constant,
3. subcooling of the condensate may be neglected,
4. momentum convective changes through the film are negligible,
5. the vapor is stationary and exerts no drag on the downward motion of the
condensate
6. heat transfer through the film is by conduction only.
The mean value of the heat transfer coefficient over the whole surface was given
by

One should note here that for a tube, L is replaced by the tube diameter, D, and
0.943 becomes 0.725. This model had been extended to include the effects of
Nusselt's assumptions. In particular, Bromley (1952) considered the effects of
subcooling within the liquid film and Rohsenow (1956) also allowed for a nonlinear distribution of temperature through the film due to energy convection. The
results indicated that the latent heat of vaporization, , in Eq. (9) should be
replaced by

However, it should be noted that in most engineering applications, the value of


is small (typically less than 0.001) and can be neglected.
Sparrow and Gregg (1959) removed assumption (4) and included inertia forces
and a convection term within the condensate film by using a boundary layer
treatment for the condensate film. For common fluids with Prandtl numbers
around and greater than unity, inertia effects are negligible for values of
less
than 2.0 For liquid metals with very low Prandtl numbers, however, the heat
transfer coefficient falls below the Nusselt prediction with increasing

when

is greater than 0.001. Poots and Miles (1967) have looked at the effect of
variable physical properties (assumption 2) on vertical plates. More recently
Koh et al. (1961) and Chen (1961) included the influence of the drag exerted by
the vapor on the liquid film. Both results show that the interfacial shear stress
can reduce heat transfer due to the effect of "hold up" of the condensate film for
low values of Pr, but this effect is small and steadily decreases with increasing
Pr for Prandtl number greater than unity.
As a conclusion for pure steam-water condensation (Pr ;SPMgt; 1), Nusselt's
assumptions can be accepted for a stationary vapor without noncondensable gas
in practical engineering situations.
Moving Pure Vapor
The effects of vapor velocity and its associated drag on the condensate film have
been found to be significant in many practical problems. For the case of vapor
flow parallel to a horizontal flat plate, Cess (1960) presented uniform property
boundary layer solutions, obtained by means of similarity transformations by
neglecting the inertia and energy convection effects within the condensate film
and assuming that the interfacial velocity was negligible in comparison with the
free stream vapor velocity. Shekriladze and Gomelauri (1966) simplified the
problem and also considered the case of an isothermal vertical plate with similar
assumptions (1973). Mayhew et al. (1966, 1987) attempted to expand Nusselt's
simple approach to take account of vapor friction as well as momentum drag.
South and Denny (1972) proposed an interpolation formula for the interfacial
shear stress in a simplified manner as Mayhew. However, such an interpolation
formula only led to a small difference in the heat transfer rate.
Jacobs (1966) used an integral method to solve the boundary layer by matching
the mass flux, shear stress, temperature and velocity at the interface. The inertia
and convection terms in the boundary layer equations of the liquid film were
neglected. The variation of the physical properties and the thermal resistance at
the vapor-liquid interface were also neglected. Since Jacobs used an incorrect
boundary condition for the vapor boundary layer, Fujii and Uehara (1972)
solved the same problem with the correct boundary condition. In addition, the
velocity profile in the vapor layer was taken as a quadratic. They presented the
numerical results and their approximate expressions for the cases of free
convection, forced convection, and mixed convection. The results show good
agreement with numerical calculations and with Cess' approximate solution
(1960).
The current recommendation in this area is the model developed by this latter
work as a best estimate. One should be cautious as the Nusselt number increases
because this implies a higher vapor and film flow with accompanying film
turbulence, not accounted for in these models. For design purposes the
recommendation is to use Nusselt laminar film model (Equ. 9.9), since it will
predict a slightly lower heat transfer coefficient, with a thicker condensate film.
Stationary Vapor with a Noncondensable Gas

A noncondensable gas can exist in a condensing environment and leads to a


significant reduction in heat transfer during condensation. A gas-vapor boundary
layer (e.g., air-steam) forms next to the condensate layer and the partial
pressures of gas and vapor vary through the boundary layer as shown in Fig. 9.2.
The buildup of noncondensable gas near the condensate film inhibits the
diffusion of the vapor from the bulk mixture to the liquid film and reduces the
rate of mass and energy transfer. Therefore, it is necessary to solve
simultaneously the conservation equations of mass, momentum and energy for
both the condensate film and the vapor-gas boundary layer together with the
conservation of specied for the vapor-gas layer. At the interface, a continuity
condition of mass, momentum and energy has to be satisfied.
For a stagnant vapor-gas mixture, Sparrow and Eckert (1961) and Sparrow and
Lin (1964) solved the mass, momentum and energy equations for laminar film
condensation on an isothermal vertical plate by using a similarity
transformation. Sparrow and Eckert (1961) considered the notion of the vaporgas mixture from the downward motion of the condensate film, whereas
Sparrow and Lin (1964) included free convection arising from density
differences associated with composition differences. These analyses indicated
that the condensing rate is dependent on the bulk gas mass fraction, the vaporgas mixture Schmidt number,
and
. The numerical calculations
show that the effect of the noncondensable gas increases with increasing
Schmidt number and increasing value of
. Since free convection arises
from both the temperature and the concentration difference in a boundary layer,
it is important when the vapor and the noncondensing gas have significantly
different relative molecular weights and that its importance increases with
increasing bulk gas mass fraction and increasing values of
. Minkowycz
and Sparrow (1965, 1966) also included free convection arising from
temperature differences. In addition, the effect of interfacial resistance,
superheating, thermal diffusion and property variation in the condensate film
and in the vapor-gas mixture were considered and concluded to be less important
except for superheating.
To reduce the computation time, Rose (1969) presented an approximate integral
boundary layer solution assuming uniform properties except for density in the
buoyancy term. Plausible velocity and concentration profiles for the vapor-gas
boundary layer were used and it was assumed that these two layers had equal
thickness. The results showed quite good agreement with those of Minkowycz
and Sparrow and this is recommended for use.
Moving Vapor with a Noncondensable Gas
For a laminar vapor-gas mixture case, Sparrow, et al. (1967) solved the
conservation equations for the liquid film and the vapor-gas boundary layer
neglecting inertia and convection in the liquid layer and assuming the streamwise velocity component at the interface to be zero in the computation of the
velocity field in the vapor-gas boundary layer. Also a reference temperature was

used for the evaluation of properties. The results showed that the effect of
noncondensable gas for the moving vapor-gas mixture case is much less than for
the corresponding stationary vapor-gas mixture. A moving vapor-gas mixture is
considered to have a "sweeping" effect, thereby resulting in a lower gas
concentration at the interface (compared to the corresponding stationary vaporgas mixture case). Also, the ratio of the heat flux with a noncondensable gas to
that without a noncondensable gas was calculated to be independent on the bulk
velocity. The computed results reveal that interfacial resistance has a negligible
effect on the heat transfer and that superheating has much less of an effect than
in the corresponding free convection case.
Koh (1962) and Fujii et al. (1977) solved this problem without the simplifying
assumptions used by Rose (1969) except for uniform properties and showed
good agreement with the approximate analysis. Instead of solving a complete set
of the conservation equations, Rose (1980) used the experimental heat transfer
result for flow over the flat plate with suction (1979). Denny et al. (1971, 1972)
also considered the case of downward vapor-gas mixture flow parallel to a
vertical flat plate. They presented a numerical solution of similar mass,
momentum and energy equations for a vapor-gas mixture by means of a forward
marching technique. Interfacial boundary conditions at each step were extracted
from a locally valid Nusselt type analysis of the condensate film. Local variable
properties in the condensate film were evaluated by means of the reference
temperature concept, while those in the vapor-gas layer were treated exactly.
Asano et al. (1978) treated the condensate film as in the Nusselt analysis but
assumed the interfacial shear stress was the same as that for single-phase flow
over an impermeable plate.
The analytical model described above was solved using only a laminar vaporgas (or pure vapor) boundary layer except for Mayhew (1966). Whitley (1976)
proposed a simple model, which uses the analogy between heat and mass
transfer for forced convection condensation of a turbulent mixture boundary
layer by neglecting the interfacial velocity and treating the surface of the
condensate film to be smooth. Kim (1990) improved on Whitley's approach for
forced convection and natural convection applications by extending it to a
wavy/turbulent film. By using well accepted correlations for a flat plate
geometry, the solution procedure is simplified to computing the condensate film
thickness and the local Reynolds and Sherwood numbers in the downstream
direction. This leads to a computationally efficient solution, which can be easily
expanded to include more detailed models of the condensate film. Total heat
flow is controlled by the gas phase heat transfer and the heat flow through the
condensate film. Therefore, the total condensation heat transfer coefficient can
be written as:

Gas phase heat transfer consists of convection heat transfer and the latent heat
released as a result of mass transfer. Radiation heat transfer can be neglected in
the temperature range of interest (30C). Hence hgas is given by

where

is defined as:

The analogy between momentum-heat-mass transfer is used in order to find the


heat and mass transfer coefficients along the plate. The friction factor and
Stanton number are correlated by

For a smooth surface, the local skin friction factor can be correlated with the
local Reynolds number, and a result like Whitley is obtained

By substituting equation 15 into equation 14, the local Nusselt number is


obtained,

Utilizing Reynold's analogy between heat and mass transfer, equation 16 is


modified to obtain the local Sherwood number:

The turbulent Prandtl and Schmidt numbers in these equations can be replaced
with the equations derived by Jischa and Rieke (Kim, 1990). They derived the
following results from transport equations of turbulent kinetic energy, heat flux
and mass flux, as

where coefficients and were fitted to experimental data. The recommended


values are given in the following table. Finally, the local Nusselt and Sherwood
numbers for a smooth surface can be obtained by substitution of equations 18
and 19 in equations 16 and 17,

Heat and mass transfer coefficients can now be solved from 20 and 21,
respectively. The condensation heat transfer coefficient hcond can then be
obtained by substituting the following definition into equation 13,

where G is the mass transfer coefficient and W is the mass fraction.


The boundary layer thickness is reduced due to the apparent suction effect of the
condensation process. This leads to larger temperature and concentration
gradients close to the interface that, consequently, increase heat and mass
transfer rates. The following correction factors were implemented to account for
the suction effect,

where

is defined as:

where

is defined as:

where

is defined as:

The droplets or waves that form on the condensate interface can increase the
shear stress and lead to enhanced turbulent mixing at the interface. The effective
surface area of the interface is also increased due to droplets and waves. The
case where waves and droplets are present was modelled as a rough surface
(Kim, 1990). Kim integrated the non-dimensional temperature profile and

expressed the resulting Stanton number as a function of the turbulent Prandtl


number, friction factor and a roughness parameter,

where the roughness parameter

is based on experimental correlation,

where is the shear stress at the wall.


The Nusselt number can be solved from equation 20,

The Sherwood number can be obtained utilizing the Reynold's analogy and
equations 31 and 32,

In the aforementioned equations, the effect of the condensate interface structure


is included in the surface roughness parameter . In Kim's original model the
interfacial waves were considered to influence the roughness parameter. A
condensate film Reynolds number of 100 was used as a critical threshold value
for wave formation. Kim used Wallis' correlation to account for waviness of the
interface by linking it to the condensate film thickness

The current recommendation in this area would be to use this simple engineering
correlation of Kim as an estimate for most situations. If more exact estimates are
necessary then other more detailed fluid mechanics analyses could be used for
the bulk gas flow.

9.3. Experimental Investigations


Stationary Pure Vapor
A number of earlier experimental results (before 1950) show some difference with the
predictions of the Nusselt theory (McAdams, 1954). The differences can be attributed to
one or more of the following reasons: 1) significant forced-convection effects; 2)
presence of noncondensable gas; 3) waviness and turbulence within the condensate
film; 4) presence of dropwise condensation.
More recently, Mills and Seban (1967) condensed steam on a copper vertical flat plate
and Slegers and Seban (1969) conducted some experiments with n-butyl alcohol. These
tests support the Nusselt theory for pure stationary vapor condensation.
Moving Pure Vapor
Mayhew and Aggarwal (1973) experimented with pure steam condensing on a flat
surface. To avoid air in-leakage, the experiments were carried out at pressures slightly
above atmospheric. Good agreement is obtained between the experimental results and
the calculated values by their own theory. It is very interesting to note that the results for
the counter-current flow cases are always appreciably higher than those predicted by the
author's own model and indeed always higher than the corresponding co-current
velocity vapor values. The reason was investigated and explained as follows in the
original paper;
An obvious explanation was provided by dye-injection tests which showed that, with
counterflow, no laminar film flow could be achieved. The film was torn off the plate
(i.e. flooding occurred at quite moderate values of vapor velocity. Similar observations
with parallel flow confirmed that the film was always both laminar and smooth. From
work with noncondensing films it was expected that rippled flow would be encountered
over part of the surface at the higher velocities used. In fact remarkable smooth films
were observed suggesting that mass transfer, and possibly also surface tension effects on
the non-isothermal film, must have had a stabilizing effect.
More recently Asano et al. (1978) reported their data for the condensation of pure
saturated vapors on a vertical flat copper plate and showed good agreement with the
authors' own model.
Stationary Vapor with a Noncondensable Gas
Perhaps the earliest definitive experiment of the effect noncondensable gas was done by
Othmer (1929), who introduced air mole fractions of up to 11 The experimental heat
transfer coefficient data of Hampson (1951) and Akers et al. (1960) were 20 Al-Diwany
and Rose (1973) reported heat transfer measurements for steam condensing in the
presence at air, argon, neon and helium. The vapor-gas mixture was passed into the
steam chamber via flow straighteners which provided uniform flow of the mixture
towards the condensing surface so as to preclude forced convection effects. The
experimental data for steam-air, steam-argon and steam-neon showed satisfactory

agreement with the predicted theoretical values of Sparrow but for steam-helium
showed a lower value than the theoretical values.
Recently, DeVuono and Christensen (1984) reported their experiment of natural
convection of a steam-air mixture at pressures above atmospheric to 0.7 MPa to
investigate the effect of pressure. The experiments were performed on a horizontal
copper tube with 7.94 cm O.D. by 1.22 m of active condensation length. The tube was
mounted in a cylindrical pressure vessel 1.52 m O.D. by 3.35 m long. Saturated steam
was supplied by an external source and allowed to diffuse to the tube resulting in
steady-state, natural convection conditions. An expression, which is a function of
,
percent noncondensable gas by volume (Y

where

MPa

0.0 < Y < 14.0


C.
Even though this experiment was done over a large range of pressure for a containment
analysis and showed a significant effect of pressure, the pipe geometry and length scale
make it questionable to apply this correlation to a large scale system. Unfortunately, the
experiment results were not compared with any other theoretical model.
Moving Vapor with a Noncondensable Gas
Rauscher, Mills and Denny (1974) performed experiments of filmwise condensation
from steam-air mixtures undergoing forced flow over 0.74 in. O.D. horizontal tube. The
heat transfer coefficient at the stagnation point was reported for bulk air mass fractions
0-7

9.4. Separate Effects and Large Scale


Tests
The previous experiments were separate effect tests in which model development was
an integral part of the research. There have been other data collected in which direct
empirical correlations have resulted or in which analysis is not completed. These
relevant experiments are summarized in Table 9.2. We consider both separate effects
data and large scale tests.

Separate Effects Experiments


Typically, the heat transfer coefficients have been observed to decrease significantly
with increased noncondensable gas mass fraction under various conditions and test
geometries. The degradation of heat transfer is caused by the accumulation of a
noncondensable gas layer near the cold wall through which the vapor must diffuse.
Buoyancy Forces in a Stagnant Gas Mixture
Cho and Stein (1988) investigated condensation of steam in the presence of air and
helium on a small horizontal plate facing down with stagnant flow conditions. The
results with air were successfully modelled by taking into account the buoyancy forces
caused by different molecular weights of participating gases. Since helium is a lighter
gas than steam, a suppression of natural convection was expected. However, the tests
with a moderate helium content showed higher heat transfer rates than predicted by a
diffusion analysis. A convective heat transfer mechanism caused by fog and mist
formation was hypothesized. Fog that formed near the cold surface was observed to
form localized swirls and generally move in downward direction. These visual
observations seemed to confirm the presence of hypothesized natural circulation. A
similar geometrical arrangement was also used by Kroger and Rohsenow (1968).
Potassium vapor was condensed in the presence of argon and helium. The diffusion
theory successfully predicted the experimental data with helium. In the case of argon,
experimental results indicated a superimposed natural circulation flow. Vapor phase
instabilities and secondary flow cells were also reported by Spencer, Chang and Moy
(1970). They investigated the condensation of Freon-113 in the presence of helium,
nitrogen and carbon-dioxide on a vertical surface under stagnant conditions. Both visual
observations and heat transfer measurements were performed. The results, indicate a
modest effect of noncondensable gas molecular weight. Dehbi (1991) studied the
influence of an air/helium noncondensable mixture on the condensation heat transfer
under stagnant conditions. The condensing surface consisted of a 3500 mm long vertical
copper tube. Helium mass fraction was varied from 1.7 to 8.3 weight percent. Dehbi
reported that the heat transfer rates decreased with a increased helium mass fraction.
When the helium mass fractions were relatively high, sharp stratification patterns were
observed as helium migrated to the top of the test vessel and air/stream mixture stayed
at the bottom. The natural convection patterns in all these data suggest that scale
dependence must be strongly considered.
Forced Flow

As mentioned previously Dallmeyer (1970) studied condensation of


and
on a
vertical plate in the presence of air. The results showed that heat transfer rates increased
with the Reynolds number and the vapor concentration. Dallmeyer performed detailed
measurements of the velocity, temperature and concentration profiles near the wall with
laminar and turbulent flow. Measured profiles illustrated the apparent suction effect of
the condensation that increases the gradients near the wall and thus leads to higher heat
and mass transfer rates in the laminar flow region. Condensation process and, in
particular, high condensate mass fluxes were observed to dampen the turbulence level in
the turbulent region.
Barry (1987) performed condensation experiments with the mixture of steam and air.
His apparatus consisted of a horizontal plate facing upwards. The velocity and mass
ratio range was chosen so that it covered the conditions that are likely to exist in a
containment during an accident in a developing parallel flow situation. Barry's results
show expected the effects of velocity and the mass ratio as mentioned previously.
Kutsuna, Inoue and Nakanishi (1987) studied condensation of steam on a horizontal
plate (facing up) in the presence of air. They also reported increased heat transfer rates
due to forced convection. Their results, indicate the expected effects of noncondensable
gas concentration and velocity on the heat transfer coefficients. Tests were performed
with higher steam content than the tests by Barry, and consequently, the heat transfer
coefficient were also significantly higher. When the tests are performed with a high
steam content, the heat transfer results become very sensitive to the air content. This
may be the reason why the data scatter is markedly higher than in Barry's experiment.
Unfortunately, the experimental uncertainties were not discussed.
Pressure
Several workers have investigated the effect of the system pressure with stagnant flow
conditions (no forced convection). The heat transfer rates are reported to increase with
system pressure (e.g., Gerstmann, 1964), because the densities of gas components
increase with pressure. Cho and Stein (1988) reported that an increase in the system
pressure (0.31 MPa to 1.24 MPa) also influenced the mode of condensation on a
downward facing surface with helium as the noncondensable gas. Higher pressures led
to mixed mode of condensation (filmwise and dropwise condensation coexisting) with a
downward facing polished surface.
Nuclear reactor safety evaluations have prompted studies of the effect of pressure on
condensation heat transfer under transient conditions (large concentrations of
noncondensable gases). Robinson and Windebank (1988) studied the effect of pressure
in the range of 0.27-0.62 MPa with an air/stream mixture. The noncondensable gas mass
fraction was varied from 24 to 88 percent. The heat transfer rates were measured with a
cooled disk that was placed inside a pressure vessel. The results show that heat transfer
rates increase with pressure and decrease with the mass ratio of noncondensable gas.
Robinson and Windebank noted that the velocity field due to the steam injection might
have had an effect on their results. The magnitude of the induced velocities within the
vessel were stated to be below 2
performed.

, although no detailed measurements were

Similar tests were also conducted by Dehbi. Heat transfer rates were measured at three
different pressures (0.15, 0.275 and 0.45 MPa). The noncondensible gas mass fraction in
the tests ranged from 25 to 90 percent. The experimental apparatus consisted of a three
meter and one half long cooled tube (0.038 m Dia) in a pressure vessel. The motivation
behind using a relatively large vertical dimension was to simulate the length scale of
internal containment structures. Surprisingly narrow pressure vessel was used (L/D =
10). This led to difficulties to establish homogeneous test conditions in the vessel. In the
tests, the mass ratio of air was 8-33 percent greater in the upper part of the vessel than in
the lower part. Secondly, the flow field created by the natural convection may have been
affected by the sidewalls. Therefore, the results by Dehbi have some unspecified
uncertainty. He confirmed the observations of Robinson that heat transfer rate increases
with system pressure.
Condensate Film Structure
Several studies have been done to address the effect of condensate film characteristics
on the heat transfer rates. The condensate film characteristics depend on its flow field
and the nature of the condensing surface, e.g. roughness, wetting and orientation.
Forced flow induces interfacial instabilities that increase the heat transfer rates by
reducing the thickness of gas phase laminar sublayer and enhancing the mixing of both
the liquid (condensate film) and gas phase. Barry (1987) studied the effects of
interfacial structure caused by shear. Since the condensation length was relatively short,
a film injection system was used to produce a condensate film that was sufficiently thick
for measurements. The qualitative results suggested that enhanced mixing, which is
caused by the interfacial film structure, somewhat compensated for the effect of the
noncondensable gas.
The surface finish has a major effect on the mode of condensation for a downward
facing surface and it is the wetting characteristics of the surface that ultimately
determine this. Dropwise condensation is likely to exist on non-wetting surfaces and
filmwise condensation is likely on wetting surfaces. In dropwise condensation mode
with polished metal surfaces, the heat transfer characteristics are likely to change due to
oxidation of the surface or tarnishing. Thus, one cannot precisely know the wetting
characteristics as surface aging occurs. Gerstmann and Griffith (1967) studied the
condensation of pure, stagnant Freon-113 and water vapor at atmospheric pressure. Heat
transfer measurements and visual observations of the interfacial behavior were made.
Gerstmann and Griffith observed several distinct flow regimes in the condensate film
depending on the angle of inclination. Unstable condensate film with pendant drops and
lengthwise ridges existed at the horizontal position. The characteristic length scale of
these formations was on the order of the Taylor wavelength. The ridge formation was
associated with the presence of a noncondensable gas. When the surface was tilted, the
condensate waves developed into "roll waves." The waves were fully developed at
about 20 degrees of inclination. The influence of the condensate film on the heat
transfer rates was successfully analyzed using an assumption of quasi-steady state with
force and energy balance equations. Generally, heat transfer rates were found to
decrease with increasing inclination angle. The presence of lengthwise ridge waves
induced by noncondensable gas was also reported by Spencer et al. (1970). However, no
discussion of the effect of the ridge waves on the heat transfer rates was given.

Integral and Large Scale Experiments


In addition to the heat transfer data, integral and large experiments provide valuable
background information about physical conditions such as gas concentrations,
temperatures, prevailing flow fields and system pressures in a particular circumstance.
The data from the experiments can be used as integral benchmarks for models. In the
subject area of condensation these data usually involve the study of condensation in
large containment structures for nuclear reactor safety.
The first integral experiments were performed by Jubb and Kolflat (1960). The results
of these integral tests were correlated with the experimental parameters. However, some
of the parameters that Jubb and Kolflat used were uniquely dependent on the
experimental apparatus. Therefore, the resulting correlations were not generally
applicable to anything other geometry.
Uchida et al. (1965) and Tagami (1965) performed experiments using a model of a
containment structure (3.4 meters dia and 6.4 meters in height). Three water cooled
cylinders inside the containment structure were used as test surfaces. Uchida measured
the post-blowdown heat transfer rates using a vertical surface with subcooling of C.
The dimensions of the test surface were 0.14 by 0.3 meters (width and height). The
pressure in the tests varied from 0.1 to 0.3 MPa. The heat transfer rates were reported to
decrease with an increasing concentration of noncondensable gas. Contrary to the
findings reported in the separate effects section, Uchida et al., reported that the heat
transfer rates depend only on the mass ratio and not on the molecular weight of the
participating gases, local velocities or pressure. Uchida's and Tagami's results can be
correlated in metric units as,
Uchida:

Tagami:

where W is defined as the mass fraction of noncondensable gas. The geometrical aspects
and the effect of velocity field were ignored. Therefore, caution should be used to
extrapolate results from the correlations for the long sections of structural walls.
Unfortunately, it is quoted and used in safety analyses.
The CVTR test series was conducted using a full scale structure of a decommissioned
nuclear power plant (Schmitt, 1970). The steam was injected through a diffuser (0.25
meters dia and 3 meters in height) located three meters above the operating floor. Three
tests were conducted. The heat transfer rates into the wall were computed from the
measured temperature profiles in the wall using the inverse conduction method. The
velocity field was measured by ultrasonic anemometers. The experimental data from the
CVTR test were used by Kim (1990) as one data set to benchmark his condensation
model. The major conclusion from the analysis of these tests was that local gas
velocities were needed to accurately predict the data.

Historically, integral tests have been used to find simple correlations that would predict
the heat transfer rates. These correlations have gained wide acceptance and are regularly
used in safety analyses. In this light, it is surprising to find out that until recently, most
of the data from these integral tests have been based on a very limited number of
measurements of the prevailing conditions. Therefore, these correlations generally have
a very limited value in making accurate predictions of heat transfer rates through
different geometries.

Table 9.1. Theoretical and Experimental Investigation of Condensation


Presence of
noncondensable gas Pure Vapor Vapor - noncondensable gas
Vapor-gas boundary Stationary Moving vapor(a) Stationary (b) Moving vapor-gas
layer condition vapor vapor-gas
Laminar Turbulent
Interface condition Smooth Smooth Wavy(c) Smooth Smooth(d) Smooth Wavy
Solution of Sparrow & Gregg (1959) Cess (1960) ***** Sparrow & Eckert (1961)
Sparrow et al (1967) Jones & Renz Kim(1990)
conservation equations Koh et al (1961) Sparrow & Lin (1964) Koh (1962)
Chen (1961) Minkowycz & Denny et al (1971)
Sparrow (1966) Asano et al (1978)
Approximate solution Nusselt (1961) Shekriladze & ***** Rose (1969) Rose (1979)
Whitley (1976) Kim(1990)
Gomelauri (1966)
Mayhew et al (1966)
Experimental work Mills&Seban (1967) Mayhew & (1973) Othmer (1929) Mills &
(1971) Dallmeyer (1970) Barry(1987)
Slegers & (1969) Aggarwal Hampson (1951) Denny Huhtiniemi(1993)
Seban Goodykoontz & (1966) Akers et al (1960) Asano et al (1978)
Dorsh (1967) Slegers &Seban (1970)
Jacobs et al (1935) Al-Diwany &
Rose (1973)
DuVuono &
Christensen (1984)
(a) Since there is no heat resistance in a pure vapor boundary layer, the classification of
a laminar or turbulent moving vapor are not needed.
(b) For the case of the presence of noncondensable gas, the natural convection flow will
be generated from the temperature and the concentration difference on the air-vapor
boundary layer.
(c) Even though the wavy interface does not mean the turbulent condensate film, both
phenomena enhance the condensation.
(d) The wavy interface with laminar moving vapor-gas mixture is not considered.
*****The analysis or data is sparse and will be discussed in the text.

9.5. Observations
Condensation phenomena can be classified by the presence of noncondensable gas, the
gas mixture velocity, the flow characterization (laminar or turbulent) of the gas mixture
and the condensate film and the interface condition as shown in Table 9.1, which
presented the summary of the theoretical and experimental investigations discussed.
For all cases with a simple geometry except the turbulent gas mixture boundary layer
and the wavy interface of both the pure vapor and the vapor-air mixture case, it is seen
that numerical solutions of the conservation equations and more approximate analytical
solutions of the conservation equations agree well with the corresponding experimental
work.
As the geometry of the condensing surface becomes more complex more prototypic
experiments must be performed. Examples of these cases are provided in the previous
section for integral containment tests. The presence of a turbulent gas mixture (natural
or forced convection) or a wavy/turbulent film interface complicates the analysis even
for simple geometries. Examples of separate effect tests and correlations under a variety
of conditions were also presented in the previous section. In these situations theoretical
analysis of this turbulent condition is still needed as is consideration of the effect of
geometric scale. This may require multi-dimensional, multi-fluid modelling of the
condensation process both near the wall and gas boundary layers as well as in the bulk
gas mixture. If this approach is taken then one must address the appropriate scaling of
these calculations to produce scaling of these calculations to produce useful
condensation heat transfer design correlations or procedures.

References

W.W. Akers, S.H. Davis, Jr and J.E. Crawford, "Condensation of a Vapor in the
Presence of a Noncondensing Gas," Chemical Engineering Progress Symposium
Series, No 30, Vol 56, pp 139-144, 1960.
H.K. Al-Diwany and J.W. Rose, "Free Convection Film Condensation of Steam
in the Presence of Noncondensing Gases," Int J Heat Mass Transfer, Vol 16, pp
1359-1369 1973.
K. Almenas and U. Lee, "A Statistical Evaluation of the Heat Transfer Data
Obtained in the HDR Containment Tests," University of Maryland, 1984.
K. Asano and Y. Nakano, "Forced Convection Film Condensation of Vapors in
the Presence of Noncondensable Gas on a Small Vertical Flat Plate," J of Chem
Engr of Japan , 1978.

J.J. Barry, "Effects of Interfacial Structure on Film Condensation," PhD Thesis,


University of Wisconsin, 1987.
L.A. Bromley, "Effect of Heat Capacity of Condensate," Ind. Eng. Chem., Vol
44, pp. 2966-2969, 1952.
R.D. Cess, "Laminar Film Condensation on a Flat Plat in the Absence of Body
Force," Zeitschrift fur Angewandte und Physik, 11, pp 426-433, 1960.
M.M. Chen, "An Analytical Study of Laminar Film Condensation Part I-Flat
Plates," J. Heat Transfer, Vol 83, Series C, pp 48-55, 1961.
D.C. Cho and R.P. Stein, "Steam Condensation on the Underside of a Horizontal
Surface," Proceedings of Third International Topical Meeting on Nuclear Power
Plant Thermal Hydr. and Operations, Nov 1988.
J.G. Collier, "Convective Boiling and Condensation," 2nd ed., McGraw-Hill
Book Company, London, 1981.
H. Dallmeyer, "Stoff-und Warmeubertragung beider Kondensation eines
Dumpfes aus einem Gemisch mit einem licht kondensierenden Gass in
Laminarer und Turbulenter Stromungsgrenzschicht," VDI-Forschungsheft 539,
pp 5-24, 1970.
A.A. Dehbi, "Analytical and Experimental Investigation of the Effects of
Noncondensable Gases on Steam Condensation under Turbulent Natural
Convection Conditions," PhD Thesis, Dept of Nuclear Engineering, MIT, Jan
1991.
V.E. Denny and V.J. Jousionis, "Effects of Noncondensable Gas and Forces
Flow on Laminar Film Condensation," Int. J Heat Mass Transfer, Vol 15, pp
315-326, 1972.
V.E. Denny and V. South III, "Effects of Forced Flow, Noncondensable and
Variable Properties on Film Condensation of Pure and Binary Vapors at the
Forward Stagnation Point of a Horizontal Cylinder," Int J Heat Mass Transfer,
Vol 15, pp 2133-2142, 1972.
V.E. Denny, A.F. Mills and V.J. Jusionis, "Laminar Film Condensation from a
Steam-Air Mixture Undergoing Forced Flow Down a Vertical Surface," J Heat
Transfer, Vol 93, pp 297-304, 1971.
A.C. DuVuono and R.N. Christensen, "Experimental Investigation of the
Pressure Effects on Film Condensation of Stem-Air Mixtures at Pressure Above
Atmospheric, Fundamentals of Phase Change; Boiling and Condensation," The
Winter Annual Meeting of ASME, New Orleans, Louisiana, HTD-Vol 38, 1984.
H.W. Emmons and D.C. Leigh, "Tabulation of the Blasius Function with
Blowing and Suction, Fluid Motion Sub-committee," Aeronaut Res Coun,
Report No. FM 1915, 1953.
T. Fujii and H. Uehara, "Laminar Filmwise Condensation on a Vertical Surface,"
Int J Heat Mass Transfer, Vol 15, pp. 217-233, 1972.
T. Fujii, H. Uehara, K. Mihara and Y. Kato, "Forced Convection in the Presence
of Non-condensables-A Theoretical Treatment for Two-Phase Laminar
Boundary Layer" (In Japanese), University of Kyushu Research Institute of
Industrial Science, Report No 66 pp 53-80, 1977.
J. Gerstmann and P. Griffith, "Laminar Film Condensation on the Underside of
Horizontal and Inclined Surfaces," Int J Heat Mass Transfer, Vol 10, pp 567-580,
1967.
J.H. Goodykoontz and R.G. Dorsch, "Local Heat Transfer Coefficients for
Condensation of Steam in Vertical Down Flow Within a 8-May inch Diameter
Tube," NASA TN D-3326, 1966.

J.H. Goodykoontz and R.G. Dorsch, "Local Heat Transfer Coefficients and
Static Pressures for Condensation in High-Velocity Steam Within a Tube,"
NASA TN D-3953, 1967.
H. Hampson, "The Condensation of Steam on a Metal Surface," Proc General
Disc on Heat Transfer, Inst of Mech Engrs and ASME, New York, pp 58-61,
1951.
C.L. Henderson and J.M. Marchello, "Film Condensation in the Presence of a
Noncondensable Gas, Transactions of ASME," J Heat Transfer, Vol 91(3), pp
447-50, 1969.
M. Jacob, S. Erk and H. Erk, "Verbasserte Messungen und Berechnungen des
Warmeuberganges beim Kondensieren Stromenden Dampfes in einem
Vertikalen Rohr," Phys Z 36(3), pp 73-84, 1935.
H.R. Jacobs, "An Integral Treatment of Combined Body Force and Forced
Convection in Laminar Film Condensation," Int. J Heat Mass Transfer, Vol 9, pp
637-648, 1966.
D.H. Jubb, "Condensation in a Reactor Containment Vessel," Nuclear
Engineering, pp 431-434, Dec 1959.
M.H. Kim and M.L. Corradini, "Modelling of Condensation Heat Transfer in a
Reactor Containment," Nuclear Engineering and Design, Vol 118, 1990.
J.C. Koh, E.M. Sparrow and J.O. Hartnett, "The Two-Phase Boundary Layer in
Laminar Film Condensation," Int J Heat Mass Transfer, pp 69-82, 1961.
J.C.Y. Koh, "Laminar Film Condensation of Condensable Gases and Gaseous
Mixture on a Flat Plate," Proc 4th USA Nat Cong Appl Mech, 2, pp 1327-1336,
1962.
A. Kolflat, "Results of 1959 Nuclear Power Plant Containment Tests," SL-1800,
March 1960. D.G. Kroger and W.M. Rohsenow, "Condensation Heat Transfer in
the Presence of Non-condensable Gas," Int J Heat Transfer Conference, Vol 11,
pp 15-26, 1968.
H. Kutsuna, K. Inoue and S. Nakanishi, "Filmwise Condensation of Vapor
Containing Noncondensable Gas in a Horizontal Duct," Int. Symposium on Heat
Transfer, Beijing 1987.
W.C. Lee, "Filmwise Condensation on a Horizontal Tube in the Presence of
Forced Convective and Non-condensing Gas," PhD. Thesis, Univ. of London,
1982.
F. Legay-Desesquelles and B. Prunet-Foch, "Dynamic Behaviour of Boundary
Layer with Condensation along a Flat Plate: Comparison with Suction," Int J
Heat Mass Transfer, Vol 28, No 12, pp 2363-2370, 1985.
Y.R. Mayhew and J.K. Aggarwal, "Laminar Film Condensation with Vapor Drag
on a Flat Surface," Int. J Heat Mass Transfer, Vol 16, pp 1944-1949, 1973.
Y.R. Mayhew, D.J. Griffith and J.W. Phillips, "Effect of Vapor Drag on Laminar
Film Condensation on a Vertical Surface," Proc. Instn Mech Engrs, Vol 180, Part
3J, pp 280-289, 1965-1966.
Y.R. Mayhew, Comments on the Paper, "Theoretical Study of Laminar Film
Condensation of Flowing Vapor," (by I.G. Shekriladze and V.I. Gomelauri), Int.
J Heat Mass Transfer, Vol 10, pp 107-108, 1987.
W.H. McAdams, "Heat Transmission," 3rd Ed, McGraw-Hill, 1954.
A.F. Mills and R.A. Seban, "The Condensation Coefficient of Water," Int. J Heat
Mass Transfer, Vol 10, pp 1815-1827, 1967.

W.J. Minkowycz and E.M. Sparrow, "Condensation Heat Transfer in the


Presence of Noncondensables, Interfacila Rersistance, Variable Properties and
Diffusion," Int J Heat Mass Transfer, Vol 9, pp 1125-1144, 1966.
W.J. Minkoycz, "Laminar Film Condensation of Water Vapor on an Isothermal
Vertical Surface," PhD Thesis, University of Minnesota, 1965.
W.A. Nusselt, "The Surface Condensation of Water Vapor," Zieschrift Ver. Deut.
Ing., Vol. 60, pp. 541-546, 1916.
D.F. Othmer, "The Condensation of Steam, Ind Eng and Chem," Vol 21, No 6,
pp 577-583 1929.
G. Poots and R. Miles, "Effects of Variable Physical Properties on Laminar Film
Condensation of Saturated Steam on a Vertical Flat Plate," Int J Heat Mass
Transfer, Vol 10, pp 1677-1692, 1967.
J.W. Rauscher, A.F. Mills and V.E. Denny, "Experimental Study of Film
Condensation from Steam-Air Mixtures Flowing Downward over a Horizontal
Tube," J. Heat Transfer, Vol 96, pp. 83-88, 1974.
J.A. Robinson and S.R. Windebank, "Measurement of Condensation Heat
Transfer Coefficients in a Steam Chamber Using a Variable Conductance Heat
Pipe," Proc 2nd UK National Conference on Heat Transfer, Vol 1, pp 617-637,
Sept 1988.
W.M. Rohsenow, "Heat Transfer and the Temperature Distribution in Laminar
Film Condensation," Trans. ASME, Vol 78, pp. 1645-1648, 1956.\ J.W. Rose,
"Condensation of a Vapor in the Presence of a Noncondensing Gas," Int J Heat
Mass Transfer, Vol 12, pp 233-237, 1969.
J.W. Rose, "Approximate Equations for Forced Convection Condensation in the
Presence of a Noncondensing Gas on a Flat Plate and Horizontal Tube," Int. J
Heat Mass Transfer, Vol 23, pp 539-546, 1980.
J.W. Rose, "Boundary Layer Flow with Transpiration of an Isothermal Flat
Plate," Int J Heat Mass Transfer, Vol 22, pp 1243-1244, 1979.
R.C. Schmitt, G.F. Bingham and J.A. Norberg, "Simulated Design Basis
Accident Tests of Carolinas Virginia Tube Reactor Containment-Final Report,"
Idaho Nuclear Corp., IN-1403, 1970.
R.C. Schmitt, G.F. Bingham and J.A. Norberg, "Simulated Design Basis
Accident Tests of Carolinas Virginia Tube Reactor Containment-Final Report,"
Idaho Nuclear Corporation, IN-1407, 1970.
I.G. Shekriladze and V.I. Gomelauri, "Theoretical Study of Laminar Film
Condensation of Flowing Vapor," Int. J Heat Mass Transfer, Vol 9, pp. 581-591,
1966.
D.C. Slaughterbeck, "Review of Heat Transfer Coefficients for Condensing
Steam in a Containment Building Following a Loss of Coolant Accident," Idaho
Nuclear Corp., IN-1388, 1970.
L. Sleger and R.A. Seban, "Nusselt Condensation of n-butil Alcohol," Int J Heat
Mass Transfer, Vol 12, pp 237-239, 1969. L. Sleger and R.A. Seban, "Laminar
Film Condensation of Steam Containing Small Concentrations of Air," Int J
Heat Mass Transfer, Vol 13, pp 1941-1947, 1970.
L. Slegers and R.A. Seban, "Laminar Film Condensation of Steam Containing
Small Concentrations of Air," Int J Heat Mass Transfer, Vol 13, pp 1941-1947,
1970.
V. South III and V.E. Denny, "The Vapor Shear Boundary Condition for Laminar
Film Condensation," Trans. ASME, 94, pp 248-249, 1972.

E.M. Sparrow and J.L. Gregg, "A Boundary Layer Treatment of Laminar Film
Condensation," J. Heat Transfer, Vol 21, Series C, pp. 13-18, 1959.
E.M. Sparrow and E.G. Eckert, "Effects of Superheated Vapor and
Noncondensable Gases on Laminar Film Condensation," AIChE J, 7,3, pp 473477, 1961.
E.M. Sparrow and S.H. Lin, "Condensation Heat Transfer in the Presence of a
Noncondensable Gas," J Heat Transfer, Vol 86, pp 430-436, 1964.
E.M. Sparrow, W.J. Minkowycz and M. Saddy, "Forced Convection in the
Presence of Noncondensables and Interfacial Resistance," Int J Heat Mass
Transfer, Vol 10, pp 1829-1845, 1967.
D.L. Spencer, K.I. Chang and H.C. Moy, "Experimental Investigation of
Stability Effects in Laminar Film Condensation on a Vertical Cylinder," 4th
International Heat Transfer Conference, Paris, Vol 6, Paper Cs 2.3, 1970.
T. Tagami, "Interim Report on Safety Assessment and Facilities Establishment
Project for June 1965," No. 1, Unpublished work, 1965.
T. Tagami, "Interim Report on Safety Assessments and Facilities Establishment
Project for June 1965," No. 1, Unpublished work, 1965.
H. Uchida, A. Oyama and Y. Togo, "Evaluation of Post-Accident Cooling
Systems of LWR's," Proc Int Conf Peaceful Uses of Atomic Energy, 13, pp 93102, 1965.
R.H. Whitley, "Condensation Heat Transfer in a Pressurized Water Reactor Dry
Containment Following a Loss of Coolant Accident," MS Thesis, University of
California at Los Angeles, 1976.

Você também pode gostar