Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Originals Heat and Mass Transfer 34 (1998) 271±277 Ó Springer-Verlag 1998

A two-phase boundary-layer model for laminar mixed-convection


condensation with a noncondensable gas on inclined plates
Y. S. Chin, S. J. Ormiston, H. M. Soliman

271
Abstract A ®nite-volume-based numerical model for Red ®lm Reynolds number = 4C=lL
mixed-convection laminar ®lm condensation from a Rex local Reynolds number = qL x u1 =lL
¯owing mixture of a vapor and a heavier noncondensable T temperature, K
gas on inclined isothermal ¯at plates is presented. The full T T ÿ Tw †= T1 ÿ Tw †
boundary layer equations for the liquid ®lm and the va- u velocity component in the x
por-gas mixtures (including liquid inertia and energy direction, m/s
convection terms) are solved implicitly with appropriate u u=u1
liquid-mixture interface conditions. Results were obtained v velocity component in the y
for three mixtures, covering wide ranges of liquid Prandtl direction, m/s
number and free-stream gas concentration in the forced- v v‰g x cos aŠÿ1=2
convection, mixed-convection and free-convection ¯ow W noncondensable gas mass fraction
regimes. The effects of liquid inertia were found to be x co-ordinate direction along the plate, m
signi®cant only for low-Prandtl-number ¯uids and lower x g x cos a=u21
gas concentrations. The effects of liquid energy convection y co-ordinate direction normal to the plate, m
were found to be signi®cant only for high-Prandtl-number y y‰g cos a= u1 mL †Š1=2
¯uids and to be most signi®cant for mixed-convection
condensation.
Greek symbols
a angle of plate inclination, measured from the
List of symbols vertical, radians
CP speci®c heat, J/kg  K C condensate ¯ow rate per unit width of ®lm,
D diffusion coef®cient, m2 /s kg/m  s
err1 ; err2 convergence checks de®ned in Eq. (16) d liquid ®lm thickness, m
g gravitational acceleration, m=s2 d d‰g cos a= u1 mL †Š1=2
hf g latent heat of vaporization, J/kg l dynamic viscosity, N  s=m2
k thermal conductivity, W/m á K m kinematic viscosity, m2 =s
m_ 00 mass ¯ux at the liquid-mixture interface, q density, kg=m3
kg/s  m2
Nux local Nusselt number = qx x=‰kL T1 ÿ Tw †Š Subscripts
Pr Prandtl number = l CP =k g noncondensable gas
qx local wall heat ¯ux, W=m2 i at the liquid-mixture interface
k iteration number
L liquid
NC numerical solution neglecting the energy
Received on 3 March 1998 convection terms in the liquid energy equation
NI numerical solution neglecting the inertia
Y. S. Chin
Safety Thermalhydraulics Branch terms in the liquid momentum equation
AECL Research v vapor
Pinawa, Manitoba, Canada R0E 1L0 w at the wall
x local value at axial location x
S. J. Ormiston 1 at the free stream
H. M. Soliman
Department of Mechanical and Industrial Engineering
University of Manitoba 1
Winnipeg, Manitoba, Canada R3T 5V6 Introduction
The process of condensation of vapors on the exterior of
Correspondence to: S. J. Ormiston
surfaces with and without the presence of a noncondens-
The ®nancial assistance provided by the Natural Sciences and able gas has several applications in the power and process
Engineering Research Council of Canada is gratefully acknowl- industries (e.g., shell-and-tube heat exchangers and con-
edged. tainment buildings of nuclear reactors during loss-of-
coolant accidents). Due to its importance, this problem has y
received considerable attention since the pioneering work
of Nusselt [1]. For pure vapors, several investigations were
undertaken in order to study the effects of the simplifying
rnal
assumptions in Nusselt's theory (such as inertia and Exte
flow
convection in the liquid ®lm and shear stress at the vapor-
liquid interface) on free-convective condensation. Results T∞
of these investigations (e.g., [2]) indicate that Nusselt's

Liqu
W∞
α
theory is valid over a wide range of conditions and mod-

i
u∞

d-m
i®cations are required mainly for cases of extreme (very g

Wall

ix
small or very large) Prandtl numbers. Similarly, it was

ture
272
shown that inertia and convection in the liquid ®lm can

inte
have signi®cant effects on forced-convective condensation

rfac
only at extreme values of Prandtl number [3, 4]. A review

e
q Wall
of the correlations for laminar ®lm condensation of pure
vapors has been prepared by Rose [5].
The presence of a noncondensable gas in the vapor
mainstream causes a signi®cant decrease in heat transfer
and liquid ®lm thickness due to the associated high con- Vapor-gas mixture
centration of gas at the interface. Therefore, it was con- boundary layer
x
cluded that the effects of inertia and convection in the Liquid film
liquid ®lm are even less signi®cant for condensation of
vapor-gas mixtures than for pure vapors. Most previous Fig. 1. Physical domain for laminar ®lm condensation on an
solutions ignored these terms for free-convective con- isothermal inclined plate from a vapor-noncondensable gas
densation (e.g., [6, 7]), forced-convective condensation mixture
(e.g., [8, 9]), and even for the case of liquid metals [10].
While these assumptions may be justi®ed for many ap-
plications, it is important to assess their in¯uence for a
wide range of ¯ow conditions in the free, forced and mixed o o
q uL † ‡ q vL † ˆ 0 1†
convection modes. ox L oy L
The objective of this investigation is to develop a so-
o o
lution technique for the full two-phase boundary-layer qL uL uL † ‡ q vL uL †
equations for mixed-convective condensation of a vapor- ox oy L
 
gas mixture on isothermal inclined plates. A numerical, o ouL
®nite-control-volume approach applied to an adaptive ˆ l ‡ g qL ÿ q1 † cos a 2†
oy L oy
expanding grid will be used in generating the solutions for  
three vapor-gas mixtures covering a wide range of Prandtl o ÿ  o ÿ  o oTL
qL uL CPL TL ‡ qL vL CPL TL ˆ kL
numbers. This numerical model can then be used in ox oy oy oy
assessing the in¯uences of liquid inertia and convection.
Another contribution or the present study is that the 3†
numerical technique developed here is extendible to other
applications such as condensation over cylinders and 2.2
inside tubes. Mixture layer equations
The vapor-gas mixture layer is assumed to be a homoge-
2 neous ideal gas mixture. At the free stream and at the
Mathematical model liquid-mixture interface, saturation conditions are as-
The governing equations are written for conservation of sumed so that the temperature at both locations is the
mass, momentum, and energy for both the liquid ®lm and equilibrium saturation temperature corresponding to the
the vapor-gas layer. It has been assumed that the ¯ow is a local partial pressure of the vapor. Equations (4), (5), and
steady, incompressible, laminar, Newtonian boundary- (6) are the vapor-gas mixture layer conservation of mass,
layer ¯ow on an isothermal ¯at plate inclined from the momentum, and energy, respectively. Equation (7) is the
vertical by an angle a, as shown in Fig. 1. Thermodynamic equation for conservation of mass of the noncondensable
and transport properties are functions of the temperature gas.
and the pressure. The mathematical model consists of
seven partial differential equations, three for the liquid o o
layer, and four for the mixture layer. qu† ‡ qv† ˆ 0 4†
ox oy
 
2.1 o o o ou
Liquid layer equations qu u† ‡ qv u† ˆ l ‡ g q ÿ q1 † cos a
ox oy oy oy
In the liquid ®lm layer, the equations for conservation of
mass, momentum, and energy are: 5†
o o
quCP T† ‡ qvCP T† y
ox  oy
  
o oT o oW
ˆ k ‡ qD CPg ÿ CPv † T 6† rnal
oy oy oy oy Exte
  f w
lo
o o o oW
qu W† ‡ qv W† ˆ qD 7† T∞
ox oy oy oy W∞
In solving Eqs. (1) to (7), the thermophysical property u∞
variation was taken into consideration by evaluating the α
properties at the local conditions. The properties varied as
functions of temperature and mixture composition, as 273
g
described in [11±13].

2.3
Boundary conditions
The boundary conditions applied to Eqs. (1) to (7) are:
 At the plate surface y ˆ 0†
x
uL ˆ 0; vL ˆ 0; and TL ˆ Tw 8† Liquid film
 At the free stream y ! 1† Liquid-mixture interface
u ˆ u1 ; T ˆ T1 ; and W ˆ W1 9† Vapor and gas mixture
boundary layer
 At the liquid-mixture interface y ˆ d†
uL ˆ u 10† Fig. 2. Computational grid
ouL ou
lL ˆl 11† The width of the grid in the y direction in the mixture
oy oy
layer was set to be a constant multiple of d. The mixture-
TL ˆ Ti 12† layer grid multiple was set by trial and error so that is was
dd dd larger than the thicknesses of the mixture hydrodynamic,
m_ 00 ˆ qL uL ÿ qL vL ˆ qu ÿ qv 13† thermal, and gas concentration boundary layers for the
dx dx
Z d  entire length of the plate for which calculations were
oTL oT d
kL ˆk ‡ hf g qL uL dy 14† performed. In addition, it was ensured that the solution
oy oy dx 0 ®elds were consistent with the free stream boundary
conditions at the outer edge of the mixture layer grid. The
oW
m_ 00 W ‡ q D ˆ0 15† value of the mixture grid multiple was highly dependent
oy on the gas-vapor combination and the free stream condi-
Equations (13), (14), and (15) impose conservation of tions. Values in the range between 1.08 (for one case of
mass, conservation of energy, and the gas impermeability sodium-argon) and 20,000 (for one case of glycerine-bro-
condition at the interface, respectively. mine) were used in this work.
In the y direction, uniform grid spacing was used in the
3 liquid layer and an exponentially-expanding grid was used
Solution method in the mixture layer with a higher concentration of control
volumes near the liquid-mixture interface. In the x direc-
3.1 tion, the grid expanded exponentially so that control vol-
Grid system umes were concentrated near the leading edge of the plate.
A numerical solution of the governing equations for the
two-phase boundary layer was obtained using a ®nite 3.2
volume method [14]. The solution domain was divided Discretization
into non-orthogonal control volumes in both the liquid In the discretization procedure, upwind differencing was
®lm and the mixture layer, as illustrated in Fig. 2. Because used in the x direction and the exponential differencing
of the parabolic nature of the governing equations a scheme [14] was used in the y direction. The boundary
marching procedure was used in the solution. The solution conditions were applied using nodes on the domain
grid had the following characteristics: the same number of boundaries and at the interface. A segregated approach
control volumes was maintained in the liquid ®lm and the was used that solved fully implicit equations for
mixture layer for all stations along the plate and the liquid- uL ; u; TL ; T, and W using a TDMA. The algebraic
mixture interface was always located at the boundary be- equations for uL and u were solved simultaneously be-
tween two control volumes. The grid therefore depended cause the continuity of velocity and shear conditions at the
dynamically on the ®lm thickness calculated from the interface were implemented implicitly in the two equations.
solution of the governing equations. Discretization was performed on the equation set (Eqs. (1)
to (7)) after a transformation that normalized the y co- 4
ordinate by d. The details of the transformed equation set Results and discussion
and the discretization procedure are given in [11]. For the system of partial differential equations considered
in this investigation, a transformation is possible from the
physical coordinates x and y to the dimensionless coor-
3.3
dinates x and y . This transformation follows the work of
Solution algorithm
Lucas [19] and is extremely useful as it allows for a concise
The solution algorithm for a single axial station began by
presentation of the results. In the transformation of the
using the solution ®elds from the previous station as a ®rst
equations, the effects of variable properties were included
estimate of the new solution ®elds and a guess for the new
and it can be shown that the seven nondimensionalized
value of d based on linear extrapolation. The initial guess
dependent variables uL ; vL Re1=2    1=2 
x ; TL ; u ; v Rex ; T ,
274 at the start of the plate was based on the original Nusselt  
and W) are functions x and y for a given vapor-gas
analysis. The discretized equations were solved and up-
combination with speci®ed values of T1 ; TW ; and W1 .
dated values for the ®elds u; uL ; T; TL ; W; and d† were
The validity of the transformation was con®rmed by
obtained following the procedure described in [11]. Iter-
producing results with different combinations of x; g; a,
ation continued at each x station until all solution ®elds
and u1 , all for the same x , and observing that the results
met a prescribed convergence criterion. In the conver-
would collapse when using the dimensionless groups from
gence tests, the change in the ®eld values between the
the transformation [11]. All dimensionless groups are
current iteration (k) and the previous iteration (k-1) was
de®ned in the Nomenclature.
normalized by either the ®eld value or the range of the
Based on the above form of dependence, it can be easily
®eld, as given by Eq. (16). In that equation, / represents a 
shown that the quantities Nux =Re1=2 x ; d , and Wi are de-
®eld variable (e.g., u; T, or W). Convergence was declared 
pendent only on x for a given vapor-gas combination and
when either err1 or err2 was smaller than 10ÿ6 throughout
speci®ed T1 ; Tw , and W1 . It is also important to point out
the calculation domain, where
that the value of x has an implication on the ¯ow regime.
At the limit x ˆ 0, gravity has no effect on the ¯ow and
/k ÿ /kÿ1 /k ÿ /kÿ1
err1 ˆ and err2 ˆ 16† consequently, there is a regime of forced-convective con-
/k range of /k densation from a moving mixture. At the limit x ˆ 1,
there is a regime of free-convective condensation from a
3.4 quiescent mixture driven by the body force. For ®nite
Grid independence tests value of x , a regime of mixed convection prevails.
Grid independence tests were performed using changes in Results for three vapor-gas mixtures were obtained in
the total heat transfer rate to the plate as the criterion. The order to demonstrate the capability of this model over a
number of control volumes across the liquid layer, the wide range of PrL . These mixtures are steam-air at
mixture layer, and along the plate was doubled and the T1 ˆ 400 K and Tw ˆ 380 K PrL  2†, sodium-argon at
change in the total heat transfer rate was examined. Grid T1 ˆ 1100 K and Tw ˆ 1000 K PrL  0:006†, and glycer-
independence was declared when further doubling of the ine-bromine at T1 ˆ 400 K and Tw ˆ 380 K PrL  1000†.
grid produced a change in total heat transfer rate of less More results are available in [11]. For each mixture, three
than 0.5% . These tests were performed separately on each separate solutions were obtained; one corresponding to
of the three vapor-gas combinations studied. The resulting the complete set of equations, the second corresponding to
grids consisted typically of 40 control volumes in the no inertia terms in the liquid momentum equation (i.e.,
transverse direction of the liquid ®lm region, 400 control parabolic velocity pro®le) and the third corresponding to
volumes in the transverse direction of the mixture region, no convection terms in the liquid energy equation (i.e.,
and 400 control volumes in the axial direction (along the linear temperature pro®le). Therefore, information was
plate length). generated to quantify the in¯uence of these terms over
wide ranges of x and W1 . In the following sections, re-
sults from the complete model are presented ®rst and then
3.5 the effects of liquid inertia and convection terms are ex-
Comparison with previous work plored.
Results from the present analysis were compared to pre- Before looking at the heat-transfer results, it is useful to
vious work for both pure vapors (in the quiescent [2], consider the trends in some other ¯ow parameters that have
forced-convection [3, 15, 16], and mixed-convection [16, strong in¯uences on heat transfer. Figure 3 shows the re-
17] regimes) and vapor-gas mixtures (in the quiescent [7], sults for Wi with the limits for exclusively forced convection
forced-convection [18], and mixed-convection [9, 10] re- and exclusively free convection included. These limits were
gimes). Details of these comparisons are given in tabular generated from the present model by setting a ˆ 90 with a
and graphical form in [11±13] and cannot be presented ®nite u1 for the forced-convection limit and a ˆ u1 ˆ 0
here due to space limitations. In general, good agreement for the free-convection limit. It is clear from Fig. 3 that the
was obtained with previous studies for the range of PrL value of Wi approaches the forced- convection limit at small
typical of steam. Deviations, sometimes fairly signi®cant, values of x and appears to approach the free-convection
were noted for very low and very high PrL , typical of liquid limit at large values of x . At both limits, the value of Wi is
metals and viscous ¯uids, respectively. independent of x . Also, an increase in W1 causes a de-
1.0 10
0
W∞ = 0.5 W∞ = 0
0.8 10-5 10-2

10-1 10-4 10-1

δ* = δ [g cos α/(u∞ vL)]1/2


1
0.6 10-3

10-2
Wi

0.4
10-1 0
Glycerine-bromine
0.2 10-3 0.5

0 10-2 275
Steam-air
Forced convection Free convection Glycerine-bromine
10-5 10-4 10-3 10-2 10-1 1 10 102 Sodium-argon
x* 10-3
10-5 10-4 10-3 10-2 10-1 1 10 102
0.8 x*

W∞ = 10-4 10-3 10-5 Fig. 4. Variation of d with x and W1 for three vapor-gas
0.6 10-1 mixtures, each with the same conditions as in Fig. 3

0.4 10-2 10
Wi

Steam-air
0.2 W∞ = 0 10-3
Sodium-argon
10-3
Nux /Rex1/2

10-2
0 Steam-air
1
Forced convection Free convection 10-1

10-5 10-4 10-3 10-2 10-1 1 10 102


x*
Forced convection

Fig. 3. Variation of Wi with x and W1 for a glycerine-bromine 10-1
mixture with T1 ˆ 450 K and Tw ˆ 350 K, a steam-air mixture 10-5 10-4 10-3 10-2 10-1 1 10 102
x*
with T1 ˆ 400 K and Tw ˆ 380 K, and a sodium-argon mixture
with T1 ˆ 1100 K and Tw ˆ 1000 K

W∞ = 0
crease in the value of x at which the free-convective limit is Glycerine-bromine
10-3
10 Sodium-argon
approached, while a decrease in W1 causes an increase in 10-2
the value of x at which the forced-convective limit is ap-
proached. These results suggesting that cases of ®nite x can 10-1
be approximated with reasonable accuracy by pure free or
forced convection have signi®cant practical signi®cance 1
Nux /Rex1/2

0.5
because of the simpli®ed nature of the solution at both 0
limits. Unfortunately, a rigorous criterion cannot be given 10-5
for the value of x at which these limits are approached
because the phenomena depend on many factors such as 10-1 10-4
W1 ; T1 ; Tw , and the properties of the vapor-gas combi-
nation.
Another feature of the results in Fig. 3 is that at any Forced convection 10-3
x ; Wi increases monotonically with W1 . The value of Wi
10-2
at the free-convective limit is always higher than Wi at the 10-5 10-4 10-3 10-2 10-1 1 10 102
forced-convective limit for any given W1 . Also, it is x*
noteworthy that a small trace of argon in the free stream of
sodium has resulted in a large gas concentration at the Fig. 5. Effect of W1 on heat transfer for three vapor-gas mix-
tures, each with the same conditions as in Fig. 3
interface. This is of course related to the ratio of molecular
weights (which impacts on the diffusion coef®cient), as
well as the relationship between saturation pressure and of W1 on d is most signi®cant near the free-convective
saturation temperature for sodium. limit, while at the forced-convective limit, this in¯uence
Figure 4 shows the results for the dimensionless ®lm decreases (to a negligible amount in some cases).
thickness d . For the same values of W1 ; T1 , and Tw , the The heat-transfer results for the three mixtures are
®lm thickness increases as the ¯ow goes from the forced- shown in Fig. 5 in terms of Nux =Re1=2
x versus x . The
convective limit to the free-convective limit. The in¯uence forced-convective limit was calculated separately and
marked in the ®gure. Considering the results of glycerine- 1.2
bromine ®rst, we note that the value of Nux =Re1=2 x at a Forced convection
given W1 starts from the forced-convective limit, in- Free convection W∞ = 0
creases gradually with x up to a point where the rate of 1.1
increase of Nux =Re1=2
x with x starts decreasing (and 10-5
sometimes a plateau is reached) over a ®nite range of x ,

Nux,NI/Nux
beyond which the slope increases again as the free-con- 10-4
vective limit is approached. This behavior is clearly evi- 1.0
10-3
dent for cases of high W1 (e.g., W1 ˆ 0:1 and 0.5), while
for pure vapors W1 ˆ 0†, the slope increases continu- Sodium-argon
ously from zero at the forced-convective limit to a maxi-
276 0.9
mum at the free-convective limit. A possible way of T∞ = 1100 K
qualitatively explaining this behavior is through a closer Tw = 1000 K
examination of the Wi -results. The free-convective com-
ponent of heat transfer is driven by the density difference 0.8 -3
10 10-2 10-1 1 10 102
qi ÿ q1 †, which is related to Wi ÿ W1 †. Therefore, for x*
®xed W1 , the behavior of the free- convective component
of heat transfer may follow the behavior of Wi in terms of Fig. 7. Effect of the liquid inertia terms on heat transfer for a
sodium-argon mixture
the rate of change with x . The results for glycerine-bro-
mine clearly support this possible explanation. Similar
trends were also found in the steam-air results while not as frequently neglected in previous analyses. A complete set
clearly visible as those of glycerine-bromine. The results of solutions was obtained without the inertia terms in the
for sodium-argon appear to be in the forced-convective liquid momentum equation (i.e., parabolic velocity pro-
zone over most of the x -range covered in Fig. 5. In all the ®le). All other terms in the governing equations and
results, the effect of W1 follows the expected qualitative boundary conditions were retained, including continuity
trend of decreasing heat transfer with increasing W1 at of interfacial shear. As expected, the calculations produced
any given x . insigni®cant effects for steam-air and glycerine-bromine at
All present results apply only to laminar ®lm conden- the presently tested conditions of T1 and Tw . The results
sation. A well-accepted condition for the existence of for sodium-argon are shown in Fig. 7 in terms of the ratio
laminar ¯ow with a smooth interface is Red < 30 [20]. of Nusselt number with the inertia terms excluded, Nux;NI ,
Results are presented in Fig. 6 in terms of Red =Re1=2
x ver- to Nusselt number from the complete model, Nux . The
sus x , from which the location x at which transition from forced- and free-convective limits are included in the
smooth laminar ¯ow occurs can be determined. These graph. Clearly the effect of the inertia terms is most pro-
results illustrate that for any vapor-gas mixture, the lam- nounced for the pure vapor and it decreases rapidly with
inar region expands signi®cantly as W1 increases due to W1 . The ratio of Nux;NI =Nux at a given W1 starts from a
the corresponding reduction in heat transfer. As well, there value below 1.0 at the forced-convective limit and in-
is a de®nite expansion in the laminar region with an creases continuously to a value above 1.0 at the free-con-
increase in PrL . vective limit. This interesting behavior was examined and
The present modeling approach has made it possible to found to be related to the ¯ow situation in the liquid ®lm.
quantify the effects of some of the terms that have been The condition Nux;NI =Nux † < 1 was found to be typically
associated with positive interfacial shear (i.e., the mixture
102
0 = W∞
Sodium-argon Glycerine-bromine
10 10-5
0.98 0.5
10 -4 T∞ = 450 K 10-1
10-3 Tw = 350 K
1 10-2
W∞ = 0
Reδ /Rex1/2

0
Nux,NC /Nux

Steam-air
-1 10-2 0.94
10
10-1 0
-2 Glycerine-bromine
10
0.90 10-3

10-3
0.5 10-2 Forced convection
-1 -3 Free convection
10 10
10-4 -3 0.86
10 10-2 10-1 1 10 102 10-6 10-4 10-2 1 102
x* x*

Fig. 6. Effect of W1 on Red for three vapor-gas mixtures, each Fig. 8. Effect of the liquid convection terms on heat transfer for a
with the same conditions as in Fig. 3 glycerine-bromine mixture
is dragging the liquid ®lm) and Nux;NI =Nux † > 1 was While this in¯uence may be small for free and forced
typically associated with negative interfacial shear. The convection, the case of mixed convection may experience
interfacial shear from the two solutions (with and without signi®cant effects particularly in the presence of a non-
inertia) was found to reach zero at nearly the same value of condensable gas.
x , which is also close to the location of Nux;NI =Nux † ˆ 1.
The second effect that was examined is that of the
convection terms in the liquid energy equation. This effect References
was signi®cant only for glycerine-bromine and the results 1. Nusselt W (1916) Die ober¯achenkondensation des was-
serdampfes. Zeitschr. Ver Deutsch. Ing. 60: 541±546; 596±575
for this mixture are shown in Fig. 8 in terms of the ratio of 2. Koh JCY; Sparrow EM; Hartnett JP (1961) The two phase
Nusselt number with no convection terms, Nux;NC , to boundary layer in laminar ®lm condensation. Int J Heat Mass
Nusselt number from the complete solution, Nux . The Transfer 2: 69±82 277
ratio of Nux;NC =Nux was found to be less than 1.0 for the 3. Koh JCY (1962) Film Condensation in a forced convection
whole range tested in this investigation. For W1 ˆ 0, the boundary layer ¯ow. Int J Heat Mass Transfer 5: 941±954
Nusselt-number ratio decreased continuously from the 4. Gaddis ES (1979) Solution of the two phase boundary-layer
forced-convection limit to the free-convection limit. equations for laminar ®lm condensation of vapour ¯owing
perpendicular to a horizontal cylinder. Int J Heat Mass
However, for W1 > 0, this ratio appears to reach a min- Transfer 22: 371±382
imum somewhere in the mixed-convection zone and this 5. Rose JW (1988) Fundamentals of condensation heat transfer:
minimum value can be signi®cantly lower than the values laminar ®lm condensation. JSME 31: 357±375
at the two limits. For example, the maximum error in 6. Sparrow EM; Lin SH (1964) Condensation heat transfer in
Nusselt number from neglecting the convection terms can the presence of a noncondensable gas. J Heat Transfer 86:
be as high as 12% for W1 ˆ 0:1, while the errors at the 430±436
forced and free convective limits are 5% and 2% , re- 7. Minkowycz WJ; Sparrow EM (1966) Condensation heat
transfer in the presence of noncondensables, interfacial re-
spectively. The results in Fig. 8 demonstrate that judging sistance, superheating, variable properties, and diffusion. Int
the importance of the energy convection terms from their J Heat Mass Transfer 9: 1125±1144
in¯uence at the two limits can lead to signi®cant errors. 8. Sparrow EM; Minkowycz WJ; Saddy M (1967) Forced con-
vection condensation in the presence of noncondensables
5 and interfacial resistance. Int J Heat Mass Transfer 10: 1829±
Concluding remarks 1845
A numerical, ®nite-control-volume approach has been 9. Denny VE; Mills AF; Jusionis VJ (1971) Laminar ®lm con-
densation from a steam-air mixture undergoing forced ¯ow
utilized for solving the full set of two-phase boundary- down a vertical surface. J Heat Transfer 93: 297±304
layer equations for laminar ®lm condensation (free-, 10. Turner RH; Mills AF; Denny VE (1973) The effect of non-
forced-, or mixed-convection regimes) on isothermal in- condensable gas on laminar ®lm condensation of liquid
clined plates in the presence of a noncondensable gas. metals. J Heat Transfer 95: 6-11
Since the model accounts for liquid inertia and convection 11. Chin YS (1995) Numerical solution of the complete two-
as well as interfacial shear, there is no restriction on its phase model for laminar ®lm condensation with a noncon-
applicability in terms of ¯uid properties (high or low PrL ) densable gas. M.Sc. Thesis, University of Manitoba, Winni-
peg, Manitoba
or the type of noncondensable gas (heavier or lighter than 12. Chin YS; Ormiston SJ; Soliman HM (1994) Effects of Liquid
the vapor). As well, the present solution technique can be Inertia and Energy Convection on Laminar Film Condensa-
extended to other geometries (inside or outside of tubes), tion With a Noncondensable Gas. Heat Transfer in Thin
different free-stream conditions (constant or varying u1 Films. New York: ASME. HTD-Vol 283: 25±32
and W1 ) and different wall conditions (constant or 13. Chin YS; Ormiston SJ; Soliman HM (1994) Numerical So-
varying wall temperature). lution of the Complete Two-Phase Model for Laminar Film
Solutions were obtained for three vapor-gas mixtures Condensation with a Noncondensable Gas. Proceedings of
the 10th International Heat Transfer Conference, Brighton,
encompassing the range 0:006  PrL  1000. The numer- UK 3: 287±292
ical predictions were validated against previous results and 14. Patankar SV (1980) Numerical Heat Transfer and Fluid Flow.
areas of agreement or deviation were identi®ed. Results of Washington: Hemisphere Publishing Corporation
the complete model were presented in dimensionless form 15. Cess RD (1960) Laminar-Film Condensation of a Flat Plate in
for different values of W1 over a wide range of x , en- the Absence of a Body Force. Z Angew Math Phys 11: 426±433
compassing the whole range of ¯ow regimes from pre- 16. Fujii T; Uehara H (1972) Laminar Filmwise Condensation on
dominantly forced convection to predominantly free a Vertical Surface. Int J Heat Mass Transfer 15: 217±233
17. Denny VE; Mills AF; Gardiner JR (1970) Nonsimilar Solu-
convection. Interesting trends were noted in the heat- tions for Laminar Film Condensation of Liquid Metals.
transfer results in the mixed-convection region and these Proceedings of the Fourth International Heat Transfer Con-
trends were found to be consistent with the results of the ference, Paris, France. Paper Cs 2.1
interfacial gas concentration. 18. Rose JW (1980) "Approximate Equations for Forced- Con-
Effects of liquid inertia were found to be signi®cant vection Condensation in the Presence of a Non-condensing
only for low PrL and these effects decrease with the Gas on a Flat Plate and Horizontal Tube. Int J Heat Mass
presence of noncondensable gases. An interesting result is Transfer 23: 539±546
19. Lucas K (1976) Combined body force and forced convection
that the inertia effects appear to diminish at some location in laminar ®lm condensation of mixed vapours-integral and
in the mixed-convection region close to the location of ®nite difference treatment. Int J Heat Mass Transfer 19:
zero interfacial shear stress. The in¯uence of liquid con- 1273±1280
vection was found to be signi®cant only for high PrL . 20. Mills AF (1992) Heat Transfer. Boston: Irwin. pg. 607

You might also like