Index

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 51

25

Heat Transfer During Condensation

2.1 Introduction higher than the wall temperature. As the wall is below its
saturation temperature, the vapor condenses, and the con-
Condensation is the phenomenon of vapor changing to densate flows down in a film under the influence of gravity,
liquid. There are two basic modes of condensation, namely, opposed by viscous and buoyancy forces. Considering the
film condensation and dropwise condensation. During film liquid film to be laminar and taking force balance on an
condensation, vapor condenses in the form of a continuous element of thickness (𝛿 − y),
liquid film over the condensing surface. During drop- du
𝜌f g(𝛿 − y)dz = 𝜇f dz + 𝜌g g(𝛿 − y)dz (2.2.1)
wise condensation, vapor condenses in the form of drops dy
covering parts of the surface. Heat transfer coefficients Integrating with the boundary condition u = 0 at z = 0,
during dropwise condensation are much higher than those ( )
(𝜌f − 𝜌g )g y2
during film condensation. Hence dropwise condensation u= 𝛿y − (2.2.2)
is desirable. However, most common fluids condense in 𝜇f 2
the film mode over ordinary metallic surfaces. Most of Integrating over the film thickness,
the condensers used in the industry today operate with 𝛿
film condensation. The mechanism of film condensation 𝜌f (𝜌f − 𝜚g )g𝛿 3
Γ = 𝜌f udy = (2.2.3)
is well understood, and the rate of heat transfer during ∫ 3𝜇f
0
film condensation can be predicted with reasonable accu-
racy in most cases. Much research has been and is being where Γ is the mass flow rate of condensate per unit width
done to develop durable heat exchangers with dropwise of plate. The rate of increase of mass flow rate with film
condensers but has not yet achieved success. thickness is obtained by differentiating with the boundary
Most of this chapter is concerned with film condensa- condition 𝛿 = 0 at z = 0.
dΓ 𝜌f (𝜌f − 𝜌g )g𝛿
tion. Dropwise condensation is discussed at the end of the 2

chapter in Section 2.10. All material up to and including = (2.2.4)


d𝛿 𝜇f
Section 2.8 is applicable only to non-metallic fluids unless
Neglecting subcooling of condensate, the latent heat of
otherwise noted. Section 2.8 deals with condensation of
the condensate added over a length dz. must equal the heat
mixtures; sections 2.2–2.7 are for single component vapors.
conducted through the liquid film:
Section 2.9 is devoted to liquid metals.
𝜌f (𝜌f − 𝜌g )g𝛿 2 d𝛿 TSAT − T𝑤
ifg = −kf dz (2.2.5)
𝜇f 𝛿
2.2 Condensation on Plates Rearranging and integrating with 𝛿 = 0 at z = 0 gives
[ ]4
2.2.1 Nusselt Equations 4𝜇f kf z(TSAT − T𝑤 )
𝛿= (2.2.6)
gifg 𝜚f (𝜌f − 𝜌g )
Nusselt (1916) presented analytical solutions to several
problems of laminar film condensation. His analysis for a The local heat transfer coefficient hz at distance z from
vertical plate is presented in the following. the top is then
Consider a vertical flat plate as shown in Figure 2.2.1. [ ]1∕4
The wall surface is at a temperature T w while the pure kf 𝜌f (𝜌f − 𝜌g )gifg kf3
hz = = (2.2.7)
saturated vapor is at the temperature T SAT , which is 𝛿 4𝜇f z(TSAT − T𝑤 )

Two-Phase Heat Transfer, First Edition. Mirza Mohammed Shah.


© 2021 John Wiley & Sons Ltd. This Work is a co-publication between John Wiley & Sons Ltd and ASME Press.
26 2 Heat Transfer During Condensation

Nusselt equations have been extensively tested against


experimental data with generally good results.

2.2.2 Modifications to the Nusselt Equations


McAdams (1954) reports a derivation by T.B. Drew, and
according to which liquid viscosity should be calculated at
the temperature T given by

T = TSAT − 0.75(TSAT − T𝑤 ) (2.2.13)


z The above derivation was made assuming linear variation
δ
of viscosity with temperature.
Bromley (1952) noted that the condensing vapor is sub-
cooled from the saturation temperature to the film temper-
ature. He showed that to account for this effect, ifg in the
Nusselt analysis should be replaced by ifg * defined as
Δz [ ]

Cpf (TSAT − T𝑤 )
ifg = ifg 1 + 0.4 (2.2.14)
ifg
y
Rohsenow (1956) has given an empirical equation for the
Figure 2.2.1 Condensation of stagnant vapor on a flat plate. effective latent heat. Based on their theoretical analysis,
Source: Rohsenow (1973a). © 1973 McGraw-Hill. Minkowycz and Sparrow (1966) recommended that liquid
properties, except latent heat, should be calculated at a
Integrating from z = 0 to z = L, the mean heat transfer reference temperature T ref given by
coefficient hm for the length L is
Tref = T𝑤 + 0.31(TSAT − T𝑤 ) (2.2.15)
L [ ]1∕4
1 𝜌f (𝜌f − 𝜌g )gifg kf3 As the temperature drop across the liquid film is gener-
hm = hz dz = 0.943
L∫ 𝜇f L(TSAT − T𝑤 ) ally small, the above modifications usually have negligible
0
effect.
(2.2.8)
Based on comparison with experimental data, McAdams
Equations (2.2.6) and (2.2.7) can be expressed in terms of (1954) recommended increasing the heat transfer coeffi-
Reynolds number defined as cients from Nusselt equations by 20%. He attributed the
4Γz increase to thinning of liquid film due to vapor shear
Rez = (2.2.9) besides the mixing action of ripples that become notice-
𝜇f
able at quite low Reynolds numbers. Stephan (1992)
The mean heat transfer coefficient can also be defined as recommends increasing Nusselt predictions by 15%.
ifg Γ Comprehensive analyses have been done by Koh et al.
hm = (2.2.10)
L(TSAT − T𝑤 ) (1961) and Chen (1961a). These authors considered several
Using Eq. (2.2.10), the alternative form of Nusselt factors that had been neglected in Nusselt’s derivation.
equations for mean heat transfer is Among them are the effect of vapor shear due to the rela-
[ ]1∕3 tive motion of liquid with respect to stagnant vapor, inertia
𝜇f2 effects, and the effect of energy convection. The results of
hm 3 = 1.47Rez=L (2.2.11)
kf 𝜌f (𝜌f − 𝜌g )g Koh et al. and Chen are in fairly good agreement. Both
found that for Prandtl number of the order of 1 or larger,
Similarly, the local heat transfer coefficient equation is
Nusselt was justified in neglecting these factors. At low
[ ]1∕3
𝜇f2 Prandtl numbers, which are characteristic of liquid metals,
hz 3 = 1.1Rez (2.2.12) these factors become important, and then heat transfer
kf 𝜌f (𝜌f − 𝜌g )g
coefficients are much lower than Nusselt predictions. Heat
If the plate is inclined at angle 𝜃 to the vertical, g is transfer coefficients decrease with increasing subcooling
replaced by (g sin 𝜃) in the above expressions. and are also reduced by vapor drag. Chen (1961a) has
2.2 Condensation on Plates 27

given the following formula to represent the results of his 2.2.4 Condensation on Underside of a Plate
analytical solution:
This problem was theoretically and experimentally inves-
( )
h 1 + 0.68𝛽 + 0.02𝛽𝛾 tigated by Gerstmann and Griffith (1967). They condensed
= (2.2.16)
hNu 1 + 0.85𝛾 − 0.15𝛽𝛾 R-113 at atmospheric pressure on the underside of plates
whose inclination to horizontal varied from 0∘ to 90∘ . For
where
horizontal surface, the condensate leaves the plate in the
Cpf ΔTSAT form of drops. For R-113, there was approximately one
𝛽= (2.2.17)
if g drop location per square centimeter. For the plates inclined
kf ΔTSAT between 3∘ and 7∘ , liquid ridges were formed from which
𝛾= (2.2.18) the drops fell. As angle of inclination was increased, ridges
𝜇f ifg
became less pronounced. At inclinations greater than
Equation (2.2.16) is applicable to liquid Prandtl numbers about 20∘ , the liquid drained from the trailing edge of the
less than 0.05 or greater than 1, 𝛽 ≤ 2, 𝛾 ≤ 20. plate. Data for inclination of 21∘ and higher were in good
agreement with the Nusselt theory. For lower inclinations,
measurements are higher than the Nusselt theory.
2.2.3 Condensation with Turbulent Film For inclined surfaces, their analysis of a ridge model
Ripples begin to form on condensate film at fairly low yielded the following expression for mean Nusselt number:
Reynolds number, about 30 according to Ghiaasiaan Num = 0.9Ra1∕6 [1 + 1.1Ra−1∕6 ] (2.2.22)
(2008). The condensate becomes turbulent when Reynolds
number exceeds a critical value, for which various values where
have been given. According to McAdams (1954), it occurs [ ]1∕2
h 𝜎
at 1800, while Isachenko et al. (1977) state it to be at 1600. Num = (2.2.23)
kf g(𝜌f − 𝜌g ) cos 𝜃
The mean heat transfer coefficient for a vertical plate or
on the outer surface of a vertical tube with both laminar and and
turbulent film may be calculated by the following correla- [ ]3∕2
tion given by Kirkbride (1934): g cos 𝜃𝜌f (𝜌f − 𝜌g )ifg 𝜎
Ra =
( )1∕2 ( )0.4 kf 𝜇f ΔTSAT g(𝜌f − 𝜌g ) cos 𝜃
𝜇f2 4Γ
hm = 0.0077 (2.2.19) (2.2.24)
g𝜌2f kf3 𝜇f
where 𝜃 is the inclination of plate to the horizontal.
This correlation was based on condensation of diphenyl Equation (2.2.22) is applicable only for Ra > 106 .
oxide and Dowtherm A on the outside of vertical tubes. For horizontal surfaces, they developed the following
Chen et al. (1987) presented the following equation for expressions through analysis:
vertical surfaces when Rez > 30: For 1010 > Ra > 108 ,
( )1∕3 Num = 0.81Ra0.193 (2.2.25)
hm 𝜈f2
1∕3 1∕2
= (Re−0.44 + 5.8 × 10−6 Re0.8
z Prf )
kf g z
For 108 > Ra > 106 ,
(2.2.20) Num = 0.69Ra0.2 (2.2.26)
The range of applicability includes both laminar and tur- All data were within about 15% of these equations.
bulent films. Howarth et al. (1978) presented a theoretical solution of
Isachenko et al. (1977) recommend the following condensation on underside of a plate, but they did not pro-
equation for mean heat transfer when both laminar vide comparison with any data.
and turbulent regimes are involved: Stein et al. (1985) studied condensation of steam on
( 2)1∕3 the underside of a horizontal plate enclosed in a vessel.
hm 𝜈f 0.25Rez Pressure ranged from 0.31 to 1.24 MPa. Steam contained
= −0.5
kf g 2300 + 41 Prf [(0.25Rez )3∕4 − 89](Prf ∕Pr𝑤 )0.25 air from 9 to 500 ppm. Heat transfer coefficients were
(2.2.21) compared to the Gerstmann and Griffith correlation given
earlier. The data for very low air content were lower than
This equation was shown to be in agreement with data this correlation at 0.31 MPa and higher than the correla-
from several sources for Reynolds numbers up to 50 000. tion at 1.24 MPa. Hence there appeared to be an effect of
28 2 Heat Transfer During Condensation

pressure. Their data were in the range of Eq. (2.2.26). They 180 and 700. They found the data to be about 20% higher
modified it to the following: than the Nusselt equation.
For 108 > Ra > 106 , Hayakawa et al. (1975) condensed ethanol, trichloroethy-
lene, and water on the outer surface of a 10 mm outside
Num = 0.787p0.464 Ra0.2 (2.2.27)
diameter (OD) vertical tube. All local heat transfer coef-
where p is in MPa. ficients were within ±20% of the Nusselt equation. They
point out that wall temperature is often not constant. In
such cases Nusselt equations for mean heat transfer coef-
2.2.5 Recommendations
ficient are not valid as these were arrived at by assuming
For vertical plates and inclined plates with condensation on (T w − T SAT ) to be constant. They appear to suggest that the
the upper side, use Nusselt equations for Rez < 1700, and larger deviations reported by experimenters may be due to
then increase their predictions by 20%. the wall temperature not being constant.
For vertical plates for Rez > 1700, use Eq. (2.2.21), the The Nusselt equations presented till now are based on the
correlation given by Isachenko et al. assumption that vapor shear is negligible. This assumption
For plates with condensation on underside, use the is valid only for very low velocities. Nusselt (1916) carried
following: out another analysis in which vapor shear was considered
For inclinations to horizontal ≥21∘ , use Nusselt but liquid film was assumed to be laminar. This solution
equations. has been outlined by Jakob (1949). Calculations with this
For inclinations <21∘ , use the correlations of Gerstmann method are very tedious. Very few reports of comparison of
and Griffith and Stein et al. in the range of their data. this solution with data have come to this author’s notice.
Dobran and Thorsen (1980) carried out an analysis for
the case of laminar liquid film with vapor shear in a vertical
2.3 Condensation Inside Plain Channels tube. They found that the heat transfer coefficients become
increasingly smaller than Nusselt equations for stagnant
2.3.1 Laminar Condensation in Vertical Tubes vapor as Fr/Re of vapor at entrance to tube decreases. The
As the condensate film thickness is small compared to effect is very strong at low liquid Prandtl numbers of the
curvature of all but the smallest diameter tubes, Nusselt order of 0.005. Their solution shows reasonable agreement
equations presented in Section 2.2 are also applicable to with some low pressure data for steam with vapor velocities
vertical tubes. The Reynolds number for tubes is defined as between 20 and 48 m s−1 . Their solution is too cumbersome
for practical use.
4𝑤 G(1 − x)D
Re = = (2.3.1)
𝜋D𝜇L 𝜇f
2.3.2 The Onset of Turbulence
where w is the mass flow rate of condensate and G is the
total mass flux of vapor and liquid. If condensation occurs All solutions discussed till now assumed that the liquid
on outside surface of the tube, D is the outside diameter. If film is laminar. For single-phase flow, it is known that
it occurs on the inside of tube, D is the inside diameter. liquid generally remains laminar till Reynolds number
Nusselt equations have been tested against a wide range exceeds 2100–2300. Hence in the absence of vapor shear,
of data by many researchers. McAdams (1954) summarized one would except the same behavior for condensate film.
the results of many early studies. The measured heat trans- Actually, transition to turbulent flow without vapor shear
fer coefficients were found to be 0.78–1.7 times the Nusselt occurs at Reynolds number of 1600–1800 though ripples in
predictions. In most cases, the measurements were higher. condensate film are seen at quite low Reynolds numbers,
Based on the data for ReLS from 80 to 2000, McAdams and Kirkbride (1934) found that friction factor for liquid
recommended that the mean heat transfer coefficient films exceeded those predicted by laminar film theory at
given by Nusselt formulas be multiplied by 1.2. Further, he Reynolds numbers exceeding 8.
recommended that the Nusselt equations be used only for In the presence of significant vapor velocities, the liquid
Reynolds numbers less than 1800. film becomes turbulent at very low Reynolds numbers. Car-
Walt and Kroger (1972) studied condensation of R-12 penter and Colburn (1951) compared their data for ethanol
inside a vertical tube. They found that the ratio of measured condensing in a vertical tube with the Nusselt solution that
to Nusselt predicted heat transfer coefficients increased includes vapor shear. The measured heat transfer coeffi-
from 1.03 to 1.15 as ReLS increased from 5 to 28. cients were found to be three times higher than the theo-
Cavallini and Zecchin (1966) compared their data for retical at vapor velocity of 27 m s−1 and seven times higher
R-11 at very low velocities in a vertical tube at ReLS between at vapor velocity of 71 m s−1 . Such wide discrepancies
2.3 Condensation Inside Plain Channels 29

could be explained only by turbulence in liquid film. They application of his mixing length theory. Their derivation
concluded that in the presence of vapor shear, liquid film can be seen in many textbooks including that by Knudsen
becomes turbulent at Reynolds number of 240. and Katz (1958).
Rohsenow et al. (1956) analyzed condensation on a verti- All the analyses mentioned previously apply these
cal plate in the presence of vapor shear. They found that the equations to the annular liquid film, which is assumed to
Reynolds number for transition to turbulent flow decreases be uniform around the circumference of the tube. Liquid
with increasing interfacial shear and can be as low as 50. entrainment is assumed zero, and the analogy between
Kim and Mudawar (2012) analyzed condensation in a heat and momentum transfer is assumed to be applica-
rectangular channel. They found the transition to turbulent ble. Thus a fixed ratio between 𝜀M and 𝜀H , usually 1, is
film to occur at Reynolds number of 25. assumed. Indeed, this assumption was implicit in Prandtl’s
Shah (2018a, 2018b) analyzed data for heat transfer derivation of these equations as he assumed the mixing
to gas–liquid flow in vertical and horizontal channels. lengths for heat and momentum transfer to be equal.
He found that the transition to turbulence occurred at However, experiments have shown that they are generally
ReLS = 170–175. not equal (Knudsen and Katz 1958).
Thus it is clear that laminar to turbulent transition dur- In order to solve for dT/dy, either the velocity dis-
ing two-phase flow occurs at much lower Reynolds number tribution or the eddy diffusivity must be known. Some
than in single-phase flow, even as low as 25. authors assumed that the velocity distribution in the
liquid film is the same as in single-phase flow and used
Von Karman’s universal velocity profile for smooth tubes.
2.3.3 Prediction of Heat Transfer in Turbulent Flow Others used eddy diffusivity expressions for single-phase
Numerous predictive methods, theoretical and empir- flow. For calculating the shear stress, frictional pres-
ical, have been published. These are discussed in the sure drop is calculated using some empirical correlation.
following text. Void fraction is also calculated with some empirical
correlation.
2.3.3.1 Analytical Models With either of these two approaches, a closed form
In many condensers, annular flow prevails over much solution is not possible without further simplifications and
of the length of the condenser. This has prompted many assumptions. Most researchers proceeded to solve them
attempts at analytical solutions for condensation in the numerically with computers. Dukler (1960) presented his
annular regime. The first to do so were Carpenter and numerical results in the form of graphs, which could be
Colburn (1951). However, the agreement of their model used for manual calculations. Azer et al. (1972) fitted an
with their own data was marginal, and it was found by equation to their computer output, which fitted their own
other researchers to have large deviations from their data, test data. However, most researchers did not try to develop
for example, Altman et al. (1959). Soliman et al. (1968) easily useable equations.
presented an improved version of the Carpenter–Colburn Among those who have presented closed form solutions
model and showed it to agree with some data. No inde- with the help of further assumptions and simplifications
pendent report of its evaluation has come to this author’s are Kosky and Staub (1971) and Traviss et al. (1973). The
notice. Calculations with this model are very tedious. latter is easier to use.
Many analyses based on a different approach have been An obvious difficulty with annular flow model analyses
made by many authors including Dukler (1960), Altman as well as those for other flow pattern-based correlations is
et al. (1959), Kosky and Staub (1971), Cavallini and Zecchin that prediction of flow patterns can have significant inac-
(1974), Azer et al. (1972), Razavi and Damle (1978), Traviss curacies as was discussed in Chapter 1. Many assumptions
et al. (1973), Bae et al. (1968), etc. The starting point for all used in the development of these models are of limited
these analyses is the following two equations: accuracy. They assume uniform liquid layer, smooth
vapor–liquid interface, and no entrainment of liquid. None
q dT
= −(𝛼 + 𝜀H ) (2.3.2) of these is generally valid. Further they use empirical
Cpf 𝜌f dy
correlations for pressure drop and void fraction that can be
𝜏 du
= (𝜈 + 𝜀M ) (2.3.3) in considerable error. Hence these analyses are not based
𝜌f dy
on secure theoretical or physical grounds. Their reported
where 𝛼 is the laminar thermal diffusivity, 𝜈 is the kine- accuracy is generally not good. For example, Razavi and
matic viscosity, 𝜀M and 𝜀H are the turbulent diffusivities for Damle (1978) compared their own solution as well as
heat and momentum, 𝜏 is the shear stress, and q is the heat several other solutions with data for water and other fluids
flux. These equations were obtained by Prandtl through from several sources. Their own solution came out best
30 2 Heat Transfer During Condensation

with mean absolute deviation (MAD) of 48%, while the The transition between the regimes occurs when J g = J g,t
Dukler solution had a MAD of 57%. given by
A similar though more sophisticated model was pre- −1∕3
⎧[ ]−3 ⎫
sented by Kim and Mudawar (2012). They analyzed ⎪ 7.5 −3 ⎪
annular flow in a rectangular minichannel. The model was Jg,t =⎨ 1.1
+ CT ⎬ (2.3.5)
⎪ 4.3Xtt + 1 ⎪
solved numerically. The results were in good agreement ⎩ ⎭
with their data for F-72 condensing in the same channel. It If J g > J g,t , the regime is ΔT independent, else it is
was not compared with any other data. ΔT-dependent regime.
In conclusion, the analytical solutions considered in this CT = 1.6 for hydrocarbons, =2.6 for other fluids. X tt is the
section are mainly of academic interest. These are not suit- Martinelli parameter defined as
able for practical design and hence none of them has been ( )0.1 ( )0.5
𝜇f 𝜌g ( )0.9
given here in detail. 1
Xtt = −1 (2.3.6)
𝜇g 𝜌f x

2.3.3.2 CFD Models In ΔT-independent regime, heat transfer coefficient is hA


There is presently much interest in computational fluid given by
( )0.3685 ( )0.2363
dynamics (CFD) modeling. An example of the application ⎛ 𝜌f 𝜇f
of this method is the work of Da Riva et al. (2012) who per- hA = hLT ⎜1 + 1.28x 0.817
⎜ 𝜚g 𝜇g
formed a numerical simulation of condensation in a 1 mm ⎝
( )2.144
diameter using the volume of fluid (VOF) method to track 𝜇f ⎞
the interface between liquid and vapor. There are many × 1− Pr−0.1 ⎟ (2.3.7)
𝜇g f ⎟
variations of the VOF method as well as of other methods. ⎠
Further there are many schemes for modeling turbulence. In ΔT-dependent regime,
Del Col et al. (2015) and Kharangate and Mudawar (2017) ( )0.8 ( )
⎡ Jg,t ⎤ J
hTP = ⎢hA − hstrat ⎥
have provided a detailed review of various CFD methods g
+ hstrat (2.3.8)
for condensation. ⎢ Jg ⎥ Jg,t
⎣ ⎦
As was noted in Section 1.9, the CFD methodology is as { [( ) ]}−1
1 − x 0.3321
yet not suitable for practical design. Therefore, a detailed hstrat = 1 + 0.741 hNu (2.3.9)
x
discussion of this topic has not been done. Those interested
hNu is the heat transfer coefficient for condensation on a
can find more information in the papers referenced herein.
single horizontal tube calculated by the following equation
given by Nusselt (1916):
2.3.3.3 Empirical Correlations [ 3 ]0.25
kf 𝜌f (𝜌f − 𝜌g )gifg
Numerous empirical correlations have been published. hNu = 0.725 (2.3.10)
Most of them are based only on the author’s own data. 𝜇f DΔT
Such correlations fail when compared with other data and hLT is the heat transfer coefficient for all mass flowing as
hence cannot be relied upon. However, there are a number liquid. It may be calculated by the following equation given
of correlations that have been shown to agree with data by McAdams (1954), generally known as the Dittus–Boelter
from many sources. Those are the ones that are discussed equation:
herein. ( )0.8
GD kf
hLT = 0.023 Pr0.4 (2.3.11)
𝜇f f D
Correlations for Conventional (Macro)Channels
This correlation was validated with data that included
Cavallini et al. Correlation Cavallini et al. (2006) have given
tube diameters from 3.1 to 17 mm, several fluids, and mass
a correlation that was verified with data for refrigerants and
velocity from 18 to 1002 kg m−2 s−1 . As ΔT appears on the
hydrocarbons in horizontal tubes from many sources. In
right-hand side of Eq. (2.3.10), iterative calculations are
this correlation, there are two regimes, a ΔT-independent
required with assumed ΔT. Kondo and Hrnjak (2012a)
(or heat flux independent) regime and a ΔT-dependent (or
found that this correlation overpredicted their data for
heat flux dependent) regime. The distinction between the
R-410A and carbon dioxide at pr > 0.7, the deviations
two regimes is made based on a dimensionless vapor veloc-
increasing as the critical pressure was approached. They
ity J g defined as
proposed a modification of Cavallini et al. correlation that
xG showed good agreement with their own data up to the
Jg = (2.3.4)
[gD𝜌g (𝜌f − 𝜌g )]0.5 critical pressure.
2.3 Condensation Inside Plain Channels 31

Shah Correlations A widely used correlation is Shah Equations (2.3.16) and (2.3.17) are used to calculate heat
(1979), which is expressed by the following equations: transfer coefficients in the three regimes as follows:
hTP In Regime I,
3.8
𝜓= = 1 + 0.95 (2.3.12)
hLS Z hTP = hI (2.3.18)
where In Regime II,
( )0.8
1
Z= −1 p0.4
r (2.3.13) hTP = hI + hNu (2.3.19)
x
where pr is the reduced pressure. hLO is the heat transfer In Regime III,
coefficient of the liquid phase flowing alone. It is calculated hTP = hNu (2.3.20)
by the following equation:
The boundaries of these three heat transfer regimes are
hLS = hLT (1 − x)0.8 (2.3.14) defined as follows:
hLT is calculated with Eq. (2.3.11). Equations (2.3.11)–(2.3.14) Horizontal tubes:
may be combined to give the correlation in the following Regime I occurs when
compact form: Jg ≥ 0.98(Z + 0.263)−0.62 (2.3.21)
hTP 3.8x0.76 (1 − x)0.04 Regime III occurs when
= (1 − x)0.8 + (2.3.15)
hLT p0.38
r
Jg ≤ 0.95(1.254 + 2.27Z 1.249 )−1 (2.3.22)
This correlation was verified with data for water, refriger-
ants, and many chemicals in horizontal and vertical tubes If neither of the above conditions is satisfied, it is
with downflow. The range of data included tube diameters Regime II.
7–40 mm, reduced pressures up to 0.44, and mass flux J g is the dimensionless vapor velocity defined by
11–1600 kg m−2 s−1 . It was realized that this correlation is Eq. (2.3.4).
likely to fail at low flow rates and higher reduced pressures. Equation (2.3.22) for the boundary of Regime III was
Shah (1981) recommended its use only if all the following given in Shah (2013), and the rest of the correlation was
three conditions are met: U GT > 3 m s−1 , ReLT > 350, and given in Shah (2009).
ReGT > 35 000. Vertical downflow:
This correlation has been widely used and continues to Regime I occurs when
be used even though the author has published improved 1
Jg ≥ (2.3.23)
versions to extend its range of applicability. Caution should 2.4Z + 0.73
be exercised to not use it beyond its recommended range, Regime III prevails when
especially not for pr > 0.44.
Jg ≤ 0.89 − 0.93 exp(−0.087Z −1.17 ) (2.3.24)
Shah (2009, 2013) presented a modified version of this
correlation to extend it to higher reduced pressures and If the Regime is not determined to be I or III by
lower flow rates. In this correlation, there are three heat Eqs. (2.3.23) and (2.3.24), it is Regime II.
transfer regimes called Regime I, Regime II, and Regime Figure 2.3.1 shows the boundaries of the regimes given by
III. The following two equations for heat transfer are used: Eqs. (2.3.21) to (2.3.24). It is seen that for most values of Z
( ) ( 𝜇 )(0.0058+0.557pr ) and J g , the area between the curves for Regimes I and III is
3.8 L much smaller for vertical tubes than for horizontal tubes.
hI = hLS 1 + 0.95 (2.3.16)
Z 14𝜇G This is the area in which Regime II occurs. Heat transfer
[ ]
3 1∕3 coefficients in Regime II are always higher than those in
−1∕3 𝜌L (𝜌L − 𝜌G )gkL
hNu = 1.32ReLO (2.3.17) Regime I or III as it is obtained by adding the equations
𝜇L 2
for those two regimes. Thus, the Shah correlation indicates
Equation (2.3.16) is a modification of the Shah (1979) that heat transfer in vertical tubes will be usually lower than
formula, done to extend it to high reduced pressures. At in horizontal tubes.
pr ≤ 0.4, it essentially equals the 1979 formula. It rep- This correlation was verified with a very wide range of
resents heat transfer exclusively by forced convection. data. The 1736 data points for 24 fluids from 51 studies were
Equation (2.3.17) is Nusselt’s equation for condensation predicted with MAD of 16.1%. The data included both hori-
in a vertical tube, Eq. (2.2.11), multiplied by 1.2 as rec- zontal and vertical tubes, while the correlations mentioned
ommended by McAdams (1954). This equation represents earlier were validated only with horizontal tube data. The
heat transfer controlled by heat flux alone. reduced pressure range is 0.0008–0.945. This is the most
32 2 Heat Transfer During Condensation

2.50 by Shah (2009, 2013) for his database with the flow pattern
Horizontal regimes I–II boundary
map given by El Hajal et al. (2003). The only data for water
2.00 was that of Varma (1977) for water in a 49 mm diameter
tube. The El Hajal et al. map is recommended only for diam-
Vertical regimes I–II boundary
1.50
eters ≤21.4 mm, and it was not validated with water data.
The map of Baker, discussed in Chapter 1, was validated
Jg

Horizontal II–III boundary


for diameters from 25 to 100 mm. It was therefore used for
1.00
Varma data. The comparison showed the following:
Vertical II–III bound.

0.50 • Heat transfer Regime I corresponded to the intermittent,


annular, and mist flow patterns.
0.00 • Heat transfer Regime II corresponded to stratified-wavy
0.01 0.1 1 10 100 flow pattern.
z
• Heat transfer Regime III corresponded to stratified flow
Figure 2.3.1 Boundaries of heat transfer regimes for vertical pattern.
and horizontal tubes in the Shah correlation (2009, 2013).
Source: Reprinted from Shah (2016a). © 2016, with permission In view of these findings, Shah (2014) gave the following
from Elsevier. flow pattern-based correlation:
If flow pattern is intermittent, annular, or mist,
verified correlation for macro channels and the only one hTP = hI (2.3.25)
that distinguishes between horizontal and vertical tubes.
It shows that heat transfer coefficients for the two orien- If flow pattern is stratified-wavy,
tations can be different for the same pressure and flow rate.
hTP = hI + hNu (2.3.26)
This is clear from Figure 2.3.1 as the regimes for horizontal
and vertical orientations are seen to differ considerably. If flow pattern is stratified,
For calculations in the heat flux-dependent range, this
correlation uses the Nusselt equation for flow inside ver- hTP = hNu (2.3.27)
tical tubes in terms of liquid Reynolds number. This has
the advantage that heat flux (or ΔT) are taken into account To determine flow patterns, it is recommended to use
implicitly and hence iterations with assumed values of heat the Baker map for low pressure water and the El Hajal
flux or ΔT are not required as is the case with the Cav- et al. map for other fluids. Applicability of other flow
allini et al. correlation and others using the Nusselt formula pattern maps is unknown. The MAD of the database with
for condensation outside tube in terms of ΔT. This makes this correlation was only slightly higher than of the Shah
design calculations easier and also enables analysis of test (2013) correlation. Figure 2.3.2 shows the comparison of
data for which ΔT is not given, as is usually the case. this and some other correlations with test data.
It is of interest to know whether the heat transfer regimes
in the Shah correlation have any physical interpretation. Some Other Correlations A few other well-known correla-
Shah (2014) compared the heat transfer regimes predicted tions are discussed herein.

Figure 2.3.2 Comparison of the data of Jung et al.


Heat transfer coefficient (W m−2 K−1)

6000
(2003) for R-142b with various correlations.
5000 D = 8 mm, T SAT = 40 ∘ C, and G = 300 kg m−2 s—1 .
“Present” is Shah (2014) flow pattern-based
correlation and “Shah” is Shah (2013) correlation.
4000
Measured Source: Reprinted from Shah (2014). ©2014, with
Shah permission from Begell House, Inc.
3000
Present
2000 Moser
Akers
1000

0
0 0.2 0.4 0.6 0.8 1
Vapor quality
2.3 Condensation Inside Plain Channels 33

A well-known correlation is the following by Ananiev For slug and bubbly flow where We* < 7X tt 2 ,
et al. (1961): [( )
hTP D 𝜙 2
0.34 g
( )0.5 = 0.048Re0.69 Pr
𝜌f kf LO f Xtt
hTP = hLT (2.3.28) ]
𝜌m 1.39 2 0.5
+ (3.2 × 10−7 Re−0.38
LO Sug ) (2.3.31)
where 𝜌m is the density of vapor–liquid mixture calculated
by the homogeneous model as 𝜙2g = 1 + CX + +X 2 (2.3.32)

𝜌f 𝜌 g (dp∕dz)f
𝜌m = (2.3.29) X2 = (2.3.33)
𝜌g + x(𝜌f − 𝜌g ) (dp∕dz)g
( )
dp 2ff G2 (1 − x)2
This correlation was originally based on water data but − = (2.3.34)
has been found by other researchers to work well with dz f 𝜌f D
( ) 2 x2
other fluids also. As it does not include heat flux, it fails in dp 2f g G
− = (2.3.35)
heat flux-dependent regime. Hence it should be used only dz g 𝜌fg D
at higher flow rates.
For Rek < 2000,
Thome et al. (2003) and Dobson and Chato (1998) have
given flow pattern-based correlations using data for hori- fk = 16Re−1
k
(2.3.36)
zontal tubes from many sources. These also use the Nusselt
For Rek = 2000–20 000,
formula Eq. (2.3.10) for condensation on horizontal tube for
predicting heat transfer in heat flux affected regimes and fk = 0.079Re−0.25 (2.3.37)
k
hence iterative calculations are required. As they require
prediction of flow patterns, their accuracy is limited by the For Rek > 20 000,
accuracy of flow pattern prediction that is far from perfect fk = 0.046Re−0.2 (2.3.38)
k
as discussed in Chapter 1.
Moser et al. (1998) developed an equivalent Reynolds For laminar flow in rectangular channel,
number-based semi-theoretical correlation that was shown fk Rek = 24(1 − 1.3553𝛽 + 1.9467𝛽 2 − 1.7012𝛽 3
to agree with data from several sources. Shah (2016c) found
+ 0.9564𝛽 4 − 0.2537𝛽 5 ) (2.3.39)
it to be generally inaccurate.
A correlation that is widely quoted in literature is that 𝛽 is the aspect ratio of channel; use its reciprocal if 𝛽 > 1.
of Akers et al. (1959). Most authors, such as Shah (2016c), The subscript k denotes for liquid or vapor.
have found it to be very inaccurate. For ReLO ≥ 2000, ReGO ≥ 2000,
( )0.35
0.03 0.1
𝜌f
C = 0.39ReLT Sug (2.3.40)
Correlations for Minichannels There has been a tremendous 𝜌g
amount of research on condensation in minichannels in
recent years. Many empirical or semiempirical correla- For ReLO ≥ 2000, ReGO < 2000,
tions have been published. Correlations for mini/micro ( )0.14
0.5
𝜌f
0.17
channels were reviewed by Awad et al. (2014) and Del C = 0.000 87ReLT Sug (2.3.41)
𝜌g
Col et al. (2015). Comparatively recent correlations for
minichannels include those by Bohdal et al. (2011), Huang For ReLO < 2000, ReGO ≥ 2000,
et al. (2010), Park et al. (2011), Jige et al. (2016), Rahman ( )0.36
et al. (2018), and Keinath and Garimella (2018). These 0.59 0.19
𝜌f
C = 0.0015ReLT Sug (2.3.42)
are based on limited databases and therefore cannot be 𝜌g
expected to have general applicability. The one that has For ReLO < 2000, ReGO < 2000,
had extensive verification is that of Kim and Mudawar ( )0.48
(2013). It is presented as follows: 0.5
𝜌f
For annular flow (smooth annular, wavy-annular, transi- C= 0.000 035Re0.44
LT Sug (2.3.43)
𝜌g
tion) where We* > 7X tt 2 ,
In Eqs. (2.3.30)–(2.3.43), DHYD is used in place of D for
hTP D 𝜙
0.34 g noncircular channels. The modified Weber number We* is
= 0.048Re0.69
LO Prf (2.3.30)
kf Xtt according to Soliman (1986) given as follows:
34 2 Heat Transfer During Condensation

For ReLO < 1250, PrTP = Prf (1 − x) + Prg x (2.3.53)


Re0.64
GO Equation (2.3.51) is dimensional; D used in it should
We∗ = 2.45 (2.3.44)
Su0.3
g (1 + 1.09Xtt0.039 )0.4 be in meter. For noncircular channels, D is to be replaced
by hydraulic equivalent diameter. They compared this
For ReLO > 1250,
correlation with a very wide range of data from many
( ) ( ) 0.084 sources that included many fluids, hydraulic diameters
Re0.79 ⎡ 𝜇 2 𝜌 ⎤
X 0.157

We = 0.85 0.3 GO tt ⎢ g f
⎥ 0.067–20 mm, and round and noncircular shapes. Good
0.039 0.4 ⎢ 𝜇 𝜌g ⎥
Sug (1 + 1.09Xtt ) ⎣ f
⎦ agreement with data was reported. The simplicity of this
(2.3.45) correlation is remarkable.

Sug is the Sugomel number defined as Shah Correlation Shah (2016b) compared his correla-
𝜌g 𝜎D tion for macro channels, Shah (2009, 2013), with data
Sug = (2.3.46) for minichannels (D ≤ 3 mm). It was found that while
𝜇g2
most data were correctly predicted, some of the data at
For channels heated only on three sides, heat transfer low mass flux were being considerably underpredicted.
coefficient calculated by the foregoing is corrected as Examination of results showed that in such cases, the
follows: correlation predicted Regime I while the data showed that
( )
Nu3 good agreement will be obtained if Regime II were used.
hTP = hTP,circular (2.3.47)
Nu4 Visual studies show that the flow patterns are affected as
where hTP,circular is the heat transfer coefficient for a circular flow passage becomes smaller. For example, Garimella
tube calculated by the foregoing equations: et al. (2001) found that the boundary between intermittent
and non-intermittent flow changes as hydraulic diameter
Nu3 = (8.235(1 − 1.833𝛽 + 3.767𝛽 2 − 5.814𝛽 3
varies from 4 mm down to 0.5 mm. This suggested that the
+ 5.361𝛽 4 − 2.0𝛽 5 )) (2.3.48) boundary between the two regimes at low flow rates was
Nu4 = (8.235(1 − 2.042𝛽 + 33.085𝛽 − 2.477𝛽 3 2 being modified by surface tension effects. At high mass
+ 1.058𝛽 4 − 0.186𝛽 5 )) (2.3.49) flow rates, inertia force dominates the surface tension
force, while at low flow rates, the surface tension force
This correlation was compared to a database that dominates. Weber number is the ratio between surface
consisted of single-channel and multi-channel data, 17 tension force and inertia force. It was therefore felt that
different working fluids (refrigerants, chemicals, CO2 ), this behavior may be quantified by its use. For all mass
hydraulic diameters from 0.424 to 6.22 mm, mass velocities flowing as vapor, it is defined as
from 53 to 1403 kg m−2 s−1 , ReLO from 276 to 89 798, qual-
G2 D
ities from 0 to 1, and reduced pressures from 0.04 to 0.91, WeGT = (2.3.54)
𝜌G 𝜎
and the MAD was 16.1%. However, Garimella et al. (2014)
found it to be in poor agreement with their database for It was found that the change from Regime I to II occurs
minichannels. when WeGT < 100. With this modification, good agreement
Garimella et al. (2014) developed a flow pattern-based with data was obtained. Further, it was found that devia-
predictive model that gave good agreement with their own tions become smaller if Eq. (2.3.16) in the Shah (2009) cor-
data collected over many years, which included many flu- relation is replaced by Eq. (2.3.7), which is the formula of
ids over a wide range of pressures and flow rate in channels Cavallini et al. (2006) for heat flux-independent regime.
of hydraulic diameters from 0.1 to 15 mm. In Shah (2016c), the correlation was compared to an
extensive database that included both mini and macro
channels. It was found that the use of Eq. (2.3.7) improved
2.3.3.4 Correlations Applicable to Both Macro
the agreement for horizontal minichannels, but for vertical
and Minichannels
minichannels and for macro channels in both orienta-
Dorao and Fernandino (2018) have given the following cor-
tions, Eq. (2.3.16) gave better agreement. Further data
relation:
analysis and modifications were done in Shah (2019a).
h1 = 0.023(ReLO + ReGO )0.8 PrTP
0.3
kf ∕D (2.3.50) The modifications included the effects of Froude number
and behavior of hydrocarbons. It was found that heat
h2 = 41.5D0.6 (ReLO + ReGO )0.4 PrTP
0.3
kf ∕D (2.3.51) transfer regimes for hydrocarbons are different from those
for other fluids. As noted earlier, this is also the case in
hTP = (h91 + h92 )1∕9 (2.3.52) the Cavallini et al. (2006) correlation. The reason for this
2.3 Condensation Inside Plain Channels 35

behavior is not known. Data analysis showed that when Table 2.3.1 Range of data with which Shah (2019a) was verified.
Fr LT < 0.012, Regimes I and III become Regime II, resulting
in higher heat transfer coefficients. This appears to be due Parameter Data range
to stratification in which condensate flows at the bottom
Fluids Water, R-11, R-12, R-22, R-32, R-41,
tube, and the liquid film in the upper circumference is R-113, R-123, R-125, R-134a, R-141b,
very thin, resulting in higher heat transfer. In developing R-142b, R-152a, R-161, R-236ea,
his correlation for boiling heat transfer, Shah (1976) found R-245fa, R-404A, R-410A, R-502, R-507,
stratification to occur at Fr LT < 0.04. Shah (2019a) reports R-1234fa, R-1234yf, R-1234ze(E), DME,
butane, propane, carbon dioxide,
that several data sets for condensation also indicated 0.04 methane, FC-72, isobutane, propylene,
as the transition value but deviations were minimized with benzene, ethanol, methanol, toluene,
transition at Fr LT = 0.012. Dowtherm 209, HFE-7100, ethane,
The final version of the correlation is as follows. pentane, HFE-7000, Novec 649, and
ammonia (43 fluids)
Heat transfer regimes are the same as in Shah (2013) cor-
Geometry Round, square, rectangle, semi-circle,
relation for both horizontal and vertical downflows if any triangle, and barrel shaped, single, and
of the following conditions are applicable: multi channels. All sides cooled or one
side insulated
• Fluid is a hydrocarbon; Regime is I and pr < 0.4.
Orientation Horizontal, vertical down
• Fluid is hydrocarbon and Regime is III.
Aspect ratio, 0.14–13.9
• ReLT < 100. width/height
If any of the mentioned conditions is fulfilled, the Shah DHYD (mm) 0.08–49.0
(2013) correlation is to be used. Reduced pressure 0.0008–0.946
If none of the mentioned conditions is applicable, heat G (kg m−2 s−1 ) 1.1–1400
transfer regimes are determined as follows: x 0.01–0.99
For horizontal flow: WeGT 0.15–79 060
Regime I occurs when WeGT > 100 and Fr LT > 0.012 and Fr LT for horizontal 0.00 035–133
channels
Jg ≥ 0.98(Z + 0.263)−0.62 (2.3.55)
Bond number 0.033–2392
Regime III occurs if Fr LT > 0.012 and Number of data 88 (75 Horizontal, 13 vertical down)
sources
Jg ≤ 0.95(1.254 + 2.27Z 1.249 )−1 (2.3.56) Number of data sets 182 (161 horizontal, 21 vertical down)
If it is not Regime I or III, it is Regime II. Source: From Shah (2019a). © 2019 Elsevier.
For vertical downflow:
Regime I occurs when WeGT > 100 and
1 This correlation was compared to an extensive database,
Jg ≥ (2.3.57)
2.4Z + 0.73 whose range of parameters is given in Table 2.3.1. It
Regime III occurs when includes 43 fluids, diameters 0.0.08–49.0 mm, reduced pres-
sures 0.0008–0.946, mass flux from 1.1 to 1400 kg m−2 s−1 ,
Jg ≤ 0.89 − 0.93 exp(−0.087Z −1.17 ) (2.3.58) channels of various shapes (round, rectangular, triangular,
If it is not Regime I or III according to Eqs. (2.3.57) and etc.) partially or fully cooled, and horizontal and vertical
(2.3.58), it is Regime II. downflow. There are 182 data sets from 88 sources. The
There are two alternative correlations. Correlation 1 uses same data were also compared to the following correla-
Eq. (2.3.7) for Regime I. Correlation 2 uses Eq. (2.3.16) tions: Kim and Mudawar (2013), Ananiev et al. (1961),
for Regime I. The two are the same in all other respects. Moser et al. (1998), Akers et al. (1959), and Dorao and
Equations (2.3.17)–(2.3.20) remain applicable. Correlation Fernandino (2018).
1 is recommended for horizontal flow when D ≤ 3 mm. Considering all data, the Shah correlations 1 and 2
Correlation 2 is recommended for horizontal channels have MAD of 19.0% and 18.6%, respectively. Using the
with D > 3 mm and for vertical channels of all diameters. two in their recommended ranges, the MAD of the Shah
For noncircular channels, DHYD is to be used as equivalent correlation is 18.2%. The other correlations have MAD
diameter in Weber and Froude numbers and DHP in all from 26.4% to 68.0%.
other places. DHP is defined as The deviations of various correlations in different ranges
4 × flow area are listed in Table 2.3.2. It is seen that the deviations of
DHP = (2.3.59) all correlations except Shah (2019a) are much higher for
perimeter with heat transfer
36 2 Heat Transfer During Condensation

30 000 Figure 2.3.3 Comparison of


Heat transfer coefficient (W m−2 K−1)

Measured correlations with the data of Wang


25 000 and Du (2000) for water in a
Shah (2019a)
horizontal tube. D = 4.98 mm,
Kim and Mudawar T SAT = 105 ∘ C, G = 14.1 kg m−2 s−1 ,
20 000 and WeGT = 24. Source: From Shah
Dorao and Fernandina
(2019a). © 2019 Elsevier.
15 000

10 000

5 000

0
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Vapor quality
Heat transfer coefficient (W m−2 K−1)

25 000 Figure 2.3.4 Data of Fronk and Garimella (2016)


Measured for ammonia in a 1.44 mm diameter horizontal
Present tube compared to various correlations.
20 000
Ananiev et al. (1961) T SAT = 40 ∘ C, G = 75 kg m−2 s−1 , and WeGT = 39.
Dorao and fernandino “Present” is Shah (2019a) correlation. Source:
15 000 From Shah (2019a). © 2019 Elsevier.

10 000

5 000

0
0.2 0.3 0.4 0.5 0.6 0.7
Vapor quality

WeGT < 100 than for WeGT > 100. This is true for D ≤ 3 mm 9000
Heat transfer coefficient (W m−2 C−1)

as well as D > 3 mm. For example, the MAD of Dorao and 8000
Measured
Fernandino correlation for D ≤ 3 mm is 39.3% at WeGT < 100 7000
Kim–Mudawar
Present 1
and 19.9% for WeGT > 100. For D > 3 mm, the deviations
6000 Present 2
of Dorao and Fernadino correlation are 56.5% and 24.1%
5000
for WeGT < 100 and WeGT > 100, respectively. The data at
the lower Weber number are underpredicted. Results with 4000
other correlations are similar. This indicates that the effects 3000
of surface tension become important at WeGT < 100 for all 2000
tube diameters, not only for D ≤ 3 mm that are usually con- 1000
sidered to be the one affected by surface tension.
0
Figures 2.3.3–2.3.5 show the comparison of data for vari- 0 0.2 0.4 0.6 0.8 1
ous fluids and parameters with various correlations. These Vapor quality
show the values and trends of heat transfer coefficients for
many conditions as well as the accuracy of various correla- Figure 2.3.5 Comparison of the data of Son and Lee (2009) for
R-134a with correlations. Present 1 and 2 are correlations of
tions. Shah (2016c). D = 3.36 mm, T SAT = 40 ∘ C, G = 400 kg m−2 s−1 , and
While the Kim and Mudawar (2013) correlation was WeGT = 1757. Source: From Shah (2016c). © 2016 Elsevier.
stated to be valid up to 6.2 mm diameter, it was found to be
inaccurate when diameter exceeded 3 mm.
In view of the results presented earlier, it is clear that the The Akers et al. correlation shows poor agreement with
Shah (2019a) is the only correlation that correctly predicts most data. The Ananiev correlation is suitable only for
data at all Weber numbers. The Dorao and Fernandino macro channels in the heat flux independent regime.
and Kim and Mudawar correlations can only be used at As seen in Table 2.3.1, the Shah correlation is in good
WeGT > 100. The latter is further limited to minichannels. agreement with data for noncircular channels of many
Table 2.3.2 Deviations of various correlations.

Deviation, %
Mean Absolute
Average
Shah Kim and Ananiev Moser Akers Dorao
Diameter (2009, Shah Mudawar et al. et al. et al. and
Orientation (mm) WeGT ReLT Fr LT Fluid N 2013) (2019a) (2013) (1961) (1998) (1959) Fernandez

Horizontal ≤3 <100 All All All 479 33.8 22.3 29.6 38.0 39.4 191.3 39.3
−23.6 3.0 −13.2 −33.2 5.6 190.6 −35.5
≤3 >100 1450 22.6 20.3 19.5 19.7 39.4 111.9 19.9
9.6 5.3 −6.3 −5.4 31.9 110.1 −4.4
≤3 All 1929 25.3 20.5 22.0 24.2 39.4 131.6 24.7
1.4 4.7 −8.1 −12.1 25.4 130.4 −12.1
>3 <100 184 30.1 26.4 72.6 54.2 38.1 34.2 56.5
−7.8 −6.0 25.6 −54.2 −35.2 12.9 −56.5
>3 >100 2517 16.7 16.1 28.3 24.1 32.7 27.8 24.1
2.3 3.2 −22.6 −13.8 17.3 −5.6 −16.0
>3 All 2701 17.6 16.4 31.3 26.2 33.1 28.3 26.3
1.6 4.1 −19.23 −16.6 13.7 −4.3 −18.8
All All ≤0.012 39 50.1 18.2 83.1 65.6 56.2 39.5 69.1
−50.1 −13.5 −14.4 −65.6 −56.2 39.3 −69.1
All All All 4630 20.8 18.5 27.4 25.4 35,7 71.3 25.6
1.5 3.1 −14.5 −14.7 18.6 51.8 −15.9
Vertical ≤3 <100 All All 67 31.0 19.6 32.1 39.3 21.6 115.7 36.1
downflow −24.9 1.5 −24.7 −36.8 −9.5 115.7 −33.9
≤3 >100 51 16.4 19.0 24.3 28.6 14.2 76.4 24.9
−14.6 −18.5 −24.7 −28.6 −2.7 76.4 −24.9
<3 All 118 24.7 19.5 28.7 34.6 18.4 98.7 31.3
−20.4 −6.0 −24.2 −33.2 −6.6 98.7 −30.0
>3 <100 56 20.7 20.5 52.9 51.1 57.9 42.2 76.8
0.1 2.5 −12.7 −50.0 −54.3 −34.2 −76.8
>3 >100 296 14.3 16.0 30.4 34.2 25.8 24.5 39.0
−0.5 1.0 6.7 24.7 15.6 −16.1 −36.9
>3 All 352 15.3 16.4 34.0 36.9 30.9 27.3 45.0
−0.4 2.2 3.6 10.3 4.5 −19.0 −43.2
All All 470 17.7 17.5 32.7 36.3 27.8 45.2 41.6
−5.5 −0.9 −3.4 −0.6 1.7 10.5 −39.9
(Continued)
Table 2.3.2 (Continued)

Deviation, %
Mean Absolute
Average
Shah Kim and Ananiev Moser Akers Dorao
Diameter (2009, Shah Mudawar et al. et al. et al. and
Orientation (mm) WeGT ReLT Fr LT Fluid N 2013) (2019a) (2013) (1961) (1998) (1959) Fernandez

Horizontal and All <100 All 786 31.8 23.1 41.5 42.8 38.9 137.4 45.7
Vertical −18.3 0.7 −5.1 −39.0 −9.5 126.6 −43.2
<100 All All HCN 56 24.8 18.5 22.5 20.9 41.5 229.8 24.0
14.9 8.3 9.8 −13.0 25.8 227.2 −9.4
>100 All 4314 18.5 18.0 25.4 23.4 34.3 56.4 23.7
4.4 2.3 −15.1 −8.2 21.9 33.7 −13.7
All All ≤100 All All 52 23.9 20.1 39.2 29.8 92.2 568.2 24.7
6.2 7.5 23.4 9.1 85.8 562.8 −13.2
All All All All All 5100 20.5 18.2 27.9 26.4 35.0 68.9 27.1
0.9 1.7 −13.6 −13.4 17.0 4.8 −18.2

Source: From Shah (2019a). © 2019 Elsevier.


2.3 Condensation Inside Plain Channels 39

Figure 2.3.6 Effect of inclination on heat transfer in a Vapor quality x = 0.25 R134a
square channel with DHYD = 1.25 mm. R-134a with 4000
T SAT = 40 ∘ C. Source: From Del Col et al. (2014). © 2014 G150
Elsevier. G135
3500 G120
G100

3000

HTC (W m−2 K−1)


2500

2000

1500

1000
0 15 30 45 60 75 90
Inclination angle (°)

shapes. The theoretical analyses by Wang and Rose (2005, Table 2.3.3 lists the salient features and range of param-
2006, 2011) showed heat transfer in noncircular channels eters in various studies on inclined channels. These are
to be higher than in round tubes. This was attributed to discussed in the following.
condensate accumulating in corners, thus thinning the All studies showed that effect of inclination decreased
liquid film on the channel surface, which reduced its with increasing mass flow rates. Figure 2.3.6 shows a typical
thermal resistance. These analyses assumed laminar film. case. It is seen that at G of 150 kg m−2 s−1 , there is negli-
As discussed in Section 2.3.2, the liquid film becomes tur- gible effect of inclination while it has considerable effect
bulent due to the effect of vapor shear. The liquid Reynolds at lower mass flux. The values of mass flux at which the
number at which this occurs can be as low as 25. Hence effect of inclination ends found by different researchers dif-
the assumption of laminar film is applicable to only a very fer. In the studies in which measurements in vertical down-
small length of the channel. This explains why correlations flow were included, heat transfer coefficient in this orienta-
for round tubes also give good agreement with data for tion were found to be lowest. The only exception is Wurfel
noncircular channels. This has been further discussed in et al. (2003) who found it to increase continuously from 0∘
Shah (2019b). to −90∘ . The results for other orientations vary greatly as
seen in Table 2.3.3. While some studies show increase in
heat transfer coefficient with increasing downward incli-
Inclined Channels Many applications involve condensa- nation, others show it to decrease. Similar is the situation
tion inside inclined tubes. By inclined tubes is meant for upward inclinations. Most remarkable is the difference
tubes at angles other than 0∘ (horizontal) and − 90∘ (ver- between the results of Mohseni et al. (2013) and Meyer et al.
tical downwards). In the foregoing, the flow orientations (2014). The two studies had the same tube diameter, the
considered have been horizontal and vertical downwards. same fluid, and overlapping range of saturation tempera-
The possible range of angles of inclination is from −90∘ ture and mass flux. While Lips and Meyer report the incli-
(vertical downwards) to +90∘ (vertical upwards). Examples nation for highest heat transfer as −30∘ , Mohseni et al. had
are air-cooled condensers for air conditioning and refriger- highest heat transfer at +30∘ inclination. This discrepancy
ation, air-cooled condensers for power plants, condensers cannot be explained by any mechanism.
for accident mitigation in advanced nuclear reactors, and Many prediction methods, theoretical and empirical have
thermosiphons. Thermosiphons have liquid and vapor been published for calculating the effect of inclination on
flowing in opposite directions; these are not considered heat transfer coefficients. These are discussed in the follow-
here but in Section 2.7.10. ing text.
Research on condensation in inclined channels was Chato (1960) developed an analytical model that requires
reviewed by Lips and Meyer (2011) and Shah (2016a). numerical solution of three differential equations. It agreed
Table 2.3.3 Salient features of studies on inclined tubes.

Maximum G for
Diameter Inclination T SAT G Inclination for effect of inclination, Data
Source (mm) (∘ ) Fluid (∘ C) pr (kg m−2 s−1 ) ReLT ReGT maximum h kg m−2 s−1 analyzable?

Chato (1960) 14.5 −3 R-113 −10 Not known No


0
Meyer et al. (2014) 8.34 −90 R-134a 30 0.189 100 5156 6.7E4 −30 to 300 Yes
+90 50 0.325 400 11 742 1.3E5 −15
Mohseni et al. (2013) 8.34 −90 R-134a 35 0.218 53 2577 3.7E4 +30 Not known Yes
+90 170 8267 1.2E5
Del Col 1.23 −90 R-134a 40 0.2494 100 760 9947 +60 150 Yes
et al. +90 390 2966 3.9E4
(2014)
−90 R-32 40 0.427 100 1292 8902 0 200 Yes
0 390 5041 34 717
Wang and 1.94, −45 Water 105 0.0065 28.8 208 4500 −45 Not known Yes
Du (2000) 0 65.1 680 15 000
2.8 19.9 208 4500 −45 Not known Yes
680 15 000
3.95 14.1 208 4500 0 Not known Yes
680 15 000
4.98 11.2 208 4500 0 Not known Yes
680 15 000
Lyulin et al. (2011) 4.8 −90 Ethanol 58 0.068 0.28 2.2 137 −30 Not known No
0 2.44 19.4 1195 −40
Tepe and 18.5 −15 Methyl 64.4 0.0122 16.3 920 27 000 Not Not known Yes
Mueller alcohol 30.5 1726 52 000 known
(1947)
Benzene 80.8 0.021 24.9 1527 52 000 Not Not known Yes
65.9 3444 120 000 known

Xing et al. (2015) 14.81 −90 R-245fa 55.4 0.113 199 11 000 260 000 +90 199 Yes
+90 699 38 000 900 000
Wurfel et al. (2003) 20.0 −90 n-Heptane −90 Not known No
0
Cho et al. (2013) 44.8 −3 Water 290 0.335 Not Not known No
+3 known

Yunxiao and Li (2015) 8.0 +90 R-410A 31 0.0065 103 7559 52 274 NA NA No
48 0.427 390 36 809 278 786

Source: From Shah (2016a). © 2016 with permission from Elsevier.


2.4 Condensation Outside Tubes 41

with his own data for inclinations of −3∘ to 0∘ . Rufer and • For horizontal or vertical downflow, use the Shah (2019a)
Kezios (1966) pointed out that the assumption made by correlation.
Chato that depth of liquid decreases toward the outlet • For other orientations, use the Shah (2016a) correlation
holds only for open channel flow. Inside tubes, liquid for inclined tubes. Use the Shah (2019a) correlation for
depth increases from inlet to outlet as more condensate is calculating hTP,0 and hTP,−90∘ required by it.
formed. They gave their own correlation. • Dorao and Fernadino (2018), Cavallini et al. (2006), and
Wang and Du (2000) noted that gravity forces tend to Kim and Mudawar (2013) correlations may be used as
stratify the flow while surface tension and inertia forces alternatives to Shah correlation for horizontal and verti-
tend to spread the liquid uniformly around the circum- cal downflow when WeGT > 100. The last mentioned is
ference. Considering these forces, they developed an further restricted to D ≤ 3 mm. The Cavallini et al. corre-
analytical model that agreed with their own data. lation should be restricted to pr < 0.7.
Lips and Meyer (2012) developed a theoretical model of
stratified flow. It agreed well with their own data for R-134a
at 200 kg m−2 s−1 . 2.4 Condensation Outside Tubes
Saffari and Naziri (2010) developed a model of strati-
fied downward flow. It consists of a system of differential 2.4.1 Single Tube
equations that are to be solved numerically. Their model
2.4.1.1 Stagnant Vapor
predicts maximum heat transfer to occur at −30∘ and
For condensation outside vertical tubes, the analyses
lowest heat transfer occurs in the horizontal position, with
and correlations for vertical plates are applicable (see
that in vertical downflow in between the two.
Section 2.2).
Xing et al. (2015) and Mohseni et al. (2013) have also
For condensation on the outer surface of a single hori-
given correlations of their own data.
zontal tube, Nusselt performed an analysis similar to that
Wurfel et al. (2003) fitted the following formula to their
for vertical plates. For the local heat transfer coefficient h𝜂
own data for inclinations −90∘ to 0∘ :
at angle 𝜂 to the vertical, he derived the following equation:
hTP,𝜃 [ ]1∕3
= (1 + sin 𝜃)0.214 (2.3.60) 𝜌f (𝜌f − 𝜌g )g sin 𝜂kf3
hTP,0
h𝜂 = 0.693 (2.4.1)
According to this formula, heat transfer is highest in ver- Γ′𝜂
tical down orientation. Γ𝜂 ′ is the mass flow rate per unit length at that angular
Shah (2016a) developed the following correlation: position. The mean heat transfer coefficient for the entire
For inclinations of −30∘ to +90∘ , perimeter is
hTP,𝜃 = hTP,0 (2.3.61) [ ]1∕4
𝜌f (𝜌f − 𝜌g )gifg kf3
hm = 0.725 (2.4.2)
For inclinations −90∘ to −30∘ , D𝜇f ΔTSAT
hTP,𝜃 = hTP,0 + (hTP,0 − hTP,−90 )(𝜃 + 30)∕60 (2.3.62) These can be expressed in terms of Reynolds number as
where hTP,𝜃 is the heat transfer coefficient at angle 𝜃 to [ ]1∕3
𝜇f2
−1∕3
the horizontal. In Shah (2016c), this correlation was used hm = 1.51Ref (2.4.3)
together with Shah (2016b) version of the correlation. The 𝜌f (𝜌f − 𝜌g )gk3f
550 data points from eight sources were predicted with The Reynolds number is defined as
MAD of 17%.
4Γ′
While this correlation does not reconcile all the dif- Ref = (2.4.4)
𝜇f
ferences in the trends reported by various researchers, it
allows prediction of heat transfer in all orientations with Γ′ is the mass flow rate of condensate per unit length
reasonable accuracy. It could be that some of the data are of tube.
incorrect. More experimental work is needed to resolve the In his analysis, Nusselt showed that 59.4% of condensate
differences. is formed on the upper half of the tube. The reason is that
the liquid film is thicker at the lower half as the condensate
from the upper half flows to the lower part from where it
2.3.4 Recommendation
drains down.
For condensation of stagnant vapor inside vertical chan- According to Collier and Thome (1994), Eq. (2.4.3) is
nels, calculate as for vertical plates. See Section 2.2.5. valid up to Ref = 3200. According to Ghiaasiaan (2008),
For condensation of vapor with forced flow in channels: transition to turbulent flow occurs at Ref = 3600.
42 2 Heat Transfer During Condensation

Comparing Eq. (2.2.8) for vertical plate/tube with due to forced convection given by
Eq. (2.4.2) for horizontal tube at the same ΔT, it is seen
hFC = 0.9kf (𝜌f ug ∕𝜇f D)1∕2 (2.4.6)
that the horizontal tube will have the same heat transfer
coefficient as the vertical tube if length of tube is 2.87 times At high velocities, contribution of hNu becomes very
the diameter. For longer lengths, heat transfer coefficient small, and at very low velocities, Eq. (2.4.5) approximates
of the horizontal tube will be higher. The reason is that the to the case of stagnant vapor.
film thickness on vertical tube keeps on increasing with Shekriladze and Gomelauri (1966) also analyzed the case
length while the condensate continuously drains down of vapor boundary layer separation. They argued that with
from the horizontal tube. boundary layer separation, there will be very little conden-
McAdams (1954) examined the data of early researchers sation on the area of tube beyond 82∘ from top and therefore
and found that the average deviation of data from Nusselt recommended multiplying hm from Eq. (2.4.5) by 0.66. But-
equation was 1.23 for steam and 0.94 for various organic terworth (1977) recommended multiplying only hFC by that
vapors. Berman (1969) examined data for steam from factor. With Butterworth modification, Eq. (2.4.6) becomes
nine studies. The data for ΔT > 10 ∘ C were within +30% hFC = 0.59kf (𝜌f ug ∕𝜇f D)1∕2 (2.4.7)
and −20%; deviations increased as ΔT decreased. Collier
and Thome (1994) state that most data are within 15% of Butterworth (1977) analyzed data for steam flowing
Nusselt equation. It may be concluded that the Nusselt down on a tube bank. He found that generally heat transfer
equation is reliable. coefficients were in between predictions with and without
Many analyses have been performed to determine the boundary layer separation. Hence use of Eq. (2.4.7) gives
effect of factors neglected in Nusselt’s analysis. A compre- conservative predictions.
hensive analysis was performed by Chen (1961b). He took In the foregoing discussions, vapor was considered to be
into consideration the effects of subcooling of condensate, flowing vertically downwards. Honda and Fujii (1974) ana-
inertial forces, and shear on moving liquid film from stag- lyzed the case of vapor approaching the tube at different
nant vapor. He found that for ordinary fluids with Prandl angles. Their results for horizontal flow are essentially the
numbers about 1 or greater, there is little departure from same as for downwards flow.
the Nusselt equations. For low Prandl numbers typical of
2.4.2 Bundles of Horizontal Tubes
liquid metals, heat transfer coefficients are much lower
than given by Nusselt equation. The results of this analysis 2.4.2.1 Vapor Entry from Top
are closely represented by Eq. (2.2.16) for flat plates, with Nusselt (1916) assumed that condensate from tubes leaves
hNu calculated by Nusselt equation for condensation on in the form of sheet and falls on the tube below without
tubes. The results of the analysis by Sparrow and Gregg splashing. Accordingly, he developed the following for-
(1959a) are close to those of Chen. Berman (1969) has mula for a vertical column of horizontal tubes. The mean
reviewed several other analyses besides these two. heat transfer coefficient hn for the nth tube in the column
is given by
2.4.1.2 Moving Vapor hn
= n0.75 − (n − 1)0.75 (2.4.8)
Moving vapor disturbs the liquid film and causes signif- h1
icant increase in heat transfer coefficient compared with where h1 is the heat transfer coefficient on the top tube. The
that of stagnant vapor. Berman and Tumanov (1962) per- mean heat transfer coefficient hN for a column of N tubes
formed tests with steam flowing downwards on tubes. Heat is given by
transfer coefficients were up to 1.6 times the predictions of hN
Nusselt equation as vapor Reynolds number reached 800. = N −1∕4 (2.4.9)
h1
Shekriladze and Gomelauri (1966) developed an analyt-
ical predictive formula by analyzing the case of downward Kern (1958) gave the following correlation:
flow of vapor on a horizontal tube. They assumed that the hN
= N −1∕6 (2.4.10)
condensate is laminar, and there is no separation of vapor h1
boundary layer. Their formula has been put in the following hn
more convenient form by Butterworth (1977): = n5∕6 − (n − 1)−5∕6 (2.4.11)
h1
hm = [0.5h2FC + (0.25h2FC + h4Nu )1∕2 ]1∕2 (2.4.5) Eisenberg (1972) has given the following correlation
based on his own data for steam condensation:
where hNu is obtained from Nusselt equation for stagnant hN
vapor, Eq. (2.4.2), and hFC is the heat transfer coefficient = 0.6 + 0.42N −1∕4 (2.4.12)
h1
2.4 Condensation Outside Tubes 43

Belghazi et al. (2001) condensed R-134a at 40 ∘ C in a


13 × 3 staggered bundle of 16.8 mm OD tubes arranged
at 60∘ . Their data are compared to three correlation in
Figure 2.4.1. The Nusselt theory underpredicts all data
that are seen to be in between the predictions of Kern and
Eisenberg formulas.
Berman (1969) studied a wide range of data. He found
that a lot of data exceeded predictions of Nusselt equation.
He points out two major reasons for the failure of the Nus-
selt theory. Firstly, condensate does not flow down from
tubes as a continuous sheet but as widely separated drops
and columnar streams. Figure 2.4.2a shows the condensate
flow assumed by Nusselt, while Figure 2.4.2b shows the (a) (b)
actual flow as described by Berman. The distance between
the liquid columns may be as large as 80–100 mm. Thus, Figure 2.4.2 Downward flow of condensate: (a) in Nusselt model
the thickening of liquid film at the bottom of tube assumed and (b) actual. Source: From Butterworth (1977). © 1977
American Society of Mechanical Engineers.
by Nusselt does not occur. The liquid where the drops and
streams fall get agitated and hence its thermal resistance is 1.0
much reduced. Berman also points out that in a large tube
0.9
bank, the vapor velocity cannot be considered negligible
except at the lowest tubes. The vapor velocity is highest at Data
0.8
the top tubes as the entire vapor condensed in the bundle
has to pass over them. Then it progressively decreases as hLN 0.7 Kern

it passes through the bundle. Hence effect of vapor veloc- hL1


0.6
ity has to be taken into consideration. He recommends Nusselt
step-by-step calculations from top to bottom. 0.5
Figure 2.4.3 shows the data examined by Berman and the
0.4
predictions of Nusselt and Kern models. Neither is seen to
be satisfactory. 0.3
Honda et al. (1988) have given a correlation that was 1 2 3 4 5 6 7
verified with data for steam and refrigerants from several Tube number, N

sources.
Figure 2.4.3 Comparison of heat transfer coefficients for lower
Eckels (2005) condensing R134a downwards at 35 tubes with those for the top tube in the bundle. Source: From
and 40 ∘ C saturation temperatures on a bundle of 25 Butterworth (1977). © 1977 American Society of Mechanical
water-cooled smooth tubes and 10 dummy half tubes. The Engineers.

Figure 2.4.1 Ratio of heat transfer


coefficients of first and nth row of a tube 1
bundle compared to three correlations.
Source: From Belghazi et al. (2001). © 2001
Elsevier. 0.8

0.6
αN /α1

Nusselt
Kern
0.4 Eissenberg
ΔT = 21 K
ΔT = 17 K
0.2 ΔT = 14 K
ΔT = 9 K
ΔT = 4 K
0
1 3 5 7 9 11 13
Row
44 2 Heat Transfer During Condensation

tubes have 19.1 mm outer diameter and are arranged in Fujii et al. (1972) carried out experiments on condensa-
a staggered pattern with a horizontal and a vertical pitch tion of saturated steam at 30 ∘ C on a bundle with 14 mm
of 22.9 mm. He compared the data with the formulas of diameter tubes. Vapor entered from the side. Their results
Nusselt, Kern, and Honda et al. (1988). Best agreement show little difference from those with downflow.
was found with Honda et al. method. However, even better Butterworth (1977) has quoted some unpublished data
agreement was found with the Nusselt formula if the for side flow in which the measured heat transfer coeffi-
exponent of N was changed to −0.08 form −0.25. cients with approach velocity of 30 m s−1 were about half of
Zeinelabdeen et al. (2018) compared a number of empir- those from Nusselt theory. Butterworth noted that in those
ical models for calculating the heat transfer coefficients for tests, the velocity was higher and temperature difference
condensation on banks of tubes with the experimental data was lower than in any published data for downflow or side
set obtained by previous investigators consisting of more flow. Hence the reason for low heat transfer coefficients
than 4000 data points for 6 different condensing fluids and may not necessarily be side flow.
13 different tube bank configurations. The configurations The conclusion is that at low to moderate velocities. Heat
included inline and staggered bundles. The fluids were transfer in side flow may be calculated by the same methods
steam, R-11, R-21, R-113, methanol, and iso-propanol. as for downflow. The situation at high velocities is unclear.
Vapor velocity approaching the bundle varied from 0 to It will be advisable to avoid high velocities.
21 m s−1 . They found the Fujii and Oda model (1986) to
be the most accurate model with MAD of about 21.5%. It 2.4.3 Recommendations
accounts for the effects of the shear stress on the surface of
the film condensate and the inundation within the bank. Stagnant vapor on vertical tube: Use the recommendations
It is given by the following equations: for vertical plate in Section 2.2.5
Stagnant vapor on horizontal tube: Use the Nusselt formu-
hm D
Nu = = (Nu4gr + Nu4sh )1∕4 (2.4.13) las, Eq. (2.4.2) or Eq. (2.4.3).
kf
Moving vapor on single tube: Eq. (2.4.5) with Eq. (2.4.7) for
Nugr = NuNu N −s (2.4.14) hFC .
1∕2 Tube bundles: Use the Fujii and Oda model, Eq. (2.4.13).
Nush = 0.9(1 + 𝜁 −1 )1∕3 ReTP,mv N −0.14 (2.4.15)
[ ]1∕2
kf ΔT 𝜇f 𝜌f 2.5 Condensation with Enhanced Tubes
𝜁= (2.4.16)
𝜇f ifg 𝜇g 𝜌g
2.5.1 Condensation on Outside Surface
umv 𝜌g D
ReTP,mv = (2.4.17) A wide variety of enhanced tubes are in use. A commonly
𝜇g
used type is the integral low finned tube. It has fins of
The exponent s is 0.16 and 0.08, respectively, for the
rectangular or trapezoidal shape. With higher surface
inline and staggered bundles. The subscripts “gr” and “sh”
tension fluids, these tend to trap condensate between
indicate gravity dominated and shear dominated. umv is the
fins in the lower part of the tube that reduces heat trans-
mean vapor velocity through the area not occupied by the
fer. A wide variety of tubes with two-dimensional and
tubes. The other models that were compared to the same
three-dimensional fins have been developed to minimize
data included those of McNaught (1982) and Honda et al.
liquid trapping.
(1986).
2.5.1.1 Single Tubes
2.4.2.2 Vapor Entry from Side
Beatty and Katz (1948) were the first to present a model
In the foregoing, vapor was considered as flowing vertically
for condensation on horizontal low finned tubes. They
downwards on the bundle. Vapor entry horizontally from
assumed that in the space between fins, there is no effect
the side is herein discussed.
of fins, and condensation can be calculated by Nus-
Fuks and Zernova (1970) performed experiments in
selt equation for single tube. Heat transfer on the fin is
which measurements were done with both downwards
calculated by the Nusselt equation for vertical plate. An
and side entry horizontal flow of vapor. The bundle had
effective length of fin Lfin was defined as
eight staggered rows of 19 mm diameter tubes. Steam was
saturated at 36–80 ∘ C. While there were some differences 𝜋(D2o − D2r )
Lfin = (2.5.1)
between the results for top and side entry of vapor, there 4Do
was little difference between the bundle average heat where Do is the diameter at the tip of fin and Dr is diameter
transfer coefficient. at the root of fin (see Figure 2.5.1).
2.5 Condensation with Enhanced Tubes 45

Figure 2.5.1 Definition of geometric P


variables for tube with trapezoidal fins.
Source: From Marto (1988). © 1988 b t
American Society of Mechanical Engineers.

h
θ
Do/2

bb tb Dr/2

The heat transfer coefficient of fin is calculated with Nus- fins. See Figure 2.5.1 for definition of geometrical variables.
selt equation for vertical plate using Lfin from Eq. (2.5.1) as Figure 2.5.2 shows predictions of Eq. (2.5.4) for rectangular
the characteristic length. Heat transfer coefficient for the fins of thickness 0.5 mm on a tube with outside diameter of
tube surface is calculated using the Nusselt formula for con- 21.05 mm. It is seen that there is a critical fin spacing below
densation on horizontal tubes. The final relation is which flooding angle decreases sharply, which causes
[ ]1∕4 deterioration in heat transfer. The dependence of critical
𝜌f (𝜌f − 𝜌g )gifg kf3
h = 0.689 spacing on (𝜎/𝜌) is also shown. The larger this ratio, the
𝜇f (TSAT − T𝑤 )
[ ] greater the critical fin spacing.
Atube −0.25 𝜂finAfin See Ali (2017) for various other methods for calculating
× Dtube + 1.3 L−0.25 (2.5.2)
Atotal Atotal fin the flooding angle.
The constant 0.689 is empirical. Atube is the area of tube Many predictive methods have been proposed for con-
surface, Afin is the area of fin surface, and Atotal is the sum densation on finned tubes taking condensate retention
of the two. The fin efficiency 𝜂 fin is calculated considering into consideration. The most conservative approach is to
conduction from the tube surface into the fin; formulas for assume no condensation in the flooded part of surface.
it may be found in most books. This equation gave satis- This will give the following formula:
factory agreement with data for six low surface tension flu-
𝜙f
ids (methyl chloride, SO2 , R-22, n-pentane, propane, and h = hbk (2.5.5)
n-butane). The tubes had 433–633 fins m−1 . 𝜋
Note that the heat transfer coefficient given by Eq. (2.5.2) where hbk is the heat transfer coefficient assuming no flood-
is based on an effective area Aeff given by ing calculated by the Beatty and Katz model. This model
Aeff = Atube + 𝜂fin Afin (2.5.3) underpredicts heat transfer.
Kumar et al. (2002) measured heat transfer coefficients
Beatty and Katz assumed that condensate flows down during condensation of steam (p = 101–110 kPa) and
freely as for bare tubes. This assumption is valid only for R134a at 39 ∘ C on five different single horizontal circular
low finned tubes and low surface tension liquids. In other integral-fin tubes and on two spine integral-fin tubes. Fin
cases, surface tension causes condensate to be retained tip diameter of the copper integral fin tubes ranged from
between the fins, making the lower part of tube ineffective.
23.82 to 24.98 mm, fin height ranged from 0.6 to 1.11 mm.
This phenomenon is known as condensate retention or
Longitudinal fin pitch varied from 0.53 to 1.07 mm for
flooding. Many expressions have been proposed for pre-
R134a and was 2.57 mm for steam. They presented a new
dicting the flooding angle 𝜙f measured from the top; see
model that is stated to be able to predict 90% of their
Figure 2.5.2. Honda et al. (1983) performed a mechanistic
own experimental data within ±15% and data from 13
analysis to develop the following expression for flooding in
investigators within 30%. Their model is described by the
trapezoidal and rectangular fins with h > b/2:
[( ) ] following equations:
4𝜎 cos 𝜃 ( )
𝜙f = cos −1
−1 (2.5.4)
𝜌f gbDo q 𝜌2f kf3 g
−1∕3
h= 0.024Ref We0.3 Y 1.4
where b is the space between fins at the tip and 𝜃 is one-half 𝜋Do LΔTSAT 𝜇f2
of the apex angle at the tip of fins; it is zero for rectangular (2.5.6)
46 2 Heat Transfer During Condensation

1000 600 400 300 Fins per meter Figure 2.5.2 Finned tube with condensate
180 retention. Source: From Marto (1988). ©
σ/p at 1 atm 1988 American Society of Mechanical
160 (10–6 m3/s2) Engineers.
R-113
11
Ethylene
140 glycol 34
Water 61
120

100
ϕf Fin
80
Tube ϕf
60 wall
ψ

40
Condensate
20

0
0 1 2 3 4 5 6
b (mm)

4𝑤
Ref = (2.5.7) (1985) models overpredicted their data, while Beatty and
𝜇f pfin
Katz (1948) and Owen et al. (1983) models underpredicted.
[ ]
4𝜎 t1 + 1 It is concluded that none of the predictive methods can be
(pfin −tr )
We =
o
(2.5.8) considered generally applicable. Their use has to be limited
Hfin 𝜌f g to the database for which they have been validated.
[ ]
(D2o −D2r ) 2.5.1.2 Tube Bundles
4𝜋 + Do to + Dr (pfin − tr )
4A 2 Eckels (2005) tested two bundles with R-134a at 35 and
Y= =
Dr − pfin Dr pfin 40 ∘ C. One of the bundles had circular integral fins, while
(2.5.9) the other had three-dimensional integral fins. Nominal
diameter of both was 19.1 mm. There were 1575 fins m−1 ,
A is the true area of the tube surface for one pitch length, and there were 10 rows of tubes arranged in staggered
tr and to are the fin thickness at root and top, Do and Dr pattern. By pumping liquid to the top, simulation was
are diameters of the fin at top and root, H fin is the height of made for up to 30 rows. The measurements at the top
fin, pfin is the fin pitch, and w is mass flow rate of conden- row were in good agreement with the Beatty and Katz
sate. Kumar et al. compared the same data with the model model. The bundle average and tube-by-tube heat transfer
of Honda and Nozu (1987) and found it to have deviations coefficients were in satisfactory agreement if the Beatty
up to 50%. and Katz prediction was adjusted as follows:
Cavallini et al. (2003) compared this model with data For bundle average,
from several sources for steam and refrigerants. Large hN
deviations were found with many data sets. They con- = N −m (2.5.10)
h1
cluded that the Kumar et al. equation is applicable only to
condensation on single copper tubes with fin height less For the nth row,
than 1.11 mm and fin pitches between 0.5 and 1 mm for hn
= n−m − (n − 1)(1−m) (2.5.11)
halogenated refrigerants and larger than 2.0 mm for steam. h1
Sajjan et al. (2015) measured condensation heat trans- where h1 is the heat transfer coefficient of the first tube
fer coefficients of refrigerant R-600a (iso-butane) over a calculated by Beatty and Katz model. They used the expo-
horizontal smooth tube of outer diameter 19 mm and five nent m = 0.08. This is half that in the Kern formula and
integral fin tubes of different fin-densities (945, 1024, 1102, 1/3rd that in the Nusselt formula. This shows that the
1181, and 1260 fpm). They compared their fin tube data effect of inundation was small. This is in agreement with
with several models. Rudy and Webb (1985) and Webb et al. the results of most other researchers as stated by Browne
2.5 Condensation with Enhanced Tubes 47

and Bansal (1999) after review of data for integral low heat transfer on the lower tubes gets lowered, resulting in
finned tubes from several sources. the row effect.
Eckels also compared his data with the models of Sarde- For other studies and prediction methods, see the review
sai et al. (1983), Briggs and Rose (1994), Sreepathi et al. papers of Browne and Bansal (1999) and Cavallini et al.
(1996), and Murata and Hashizume (1992). These models (2003).
are more complex and take into considerations the effects
of surface tension. Deviations of Sardesai et al. model were 2.5.2 Condensation Inside Enhanced Tubes
comparable to those with Beatty and Katz model, while
those of the others were higher. Many types of internally enhanced tubes are in use. More
Their tests on the bundle with three-dimensional tubes common are those with helical or herringbone fins shown
showed progressively lower heat transfer coefficients at the in Figure 2.5.4.
lower rows. Figure 2.5.3 is an example. They have given Condensation in enhanced tubes has been reviewed by
equations to fit their row-by-row data. After reviewing data Cavallini et al. (2003) and Dalkilic and Wongwises (2009).
for bundles of three-dimensional tubes, Browne and Bansal Cavallini et al. (1995, 1999) presented a correlation
(1999) noted that most studies show that there is significant for heat transfer during condensation in microfin and
row effect in such bundles. Thus, the results of Eckels are cross-grooved tubes by modifying the correlation of Cav-
in line with those of other researchers. allini and Zecchin (1974) for smooth tubes. It was based
The difference in row effect between integral finned tubes on data for refrigerants from 15 sources. Wang and Honda
and three-dimensional enhanced tubes can be explained (2003) compared data for one-and two-dimensional inte-
by different condensate flow in the two. Honda et al. (1987) gral finned tubes with several predictive techniques. The
studied condensate flow in a bundle of integral fin tubes. Cavallini et al. (1999) correlation was found to under-
They found four flow patterns: drop, column, column and predict data for R-11 and R-123, which was at pr < 0.1.
sheet, and sheet. Columns occur at constant pitch. At high They found best results with their own theoretical model
condensate rates, sheet mode occurs. Webb and Murawski that is difficult to use as it involves numerical solution
(1990) studied condensation on three-dimensional tubes. of a set of equations. The next best results were with the
They found that the drainage pattern was very unstable correlation of Yu and Koyama (1998), which predicted
the low pressure data well. This correlation is applicable
and that there was a lateral flow of condensate not seen
only to microfin tubes. Cavallini et al. (2003) found it to
with integral fin tubes. Condensate falling on a tube drains
give unsatisfactory agreement with data. They also report
down not directly below there but at some distance away
unsatisfactory results with the correlation of Kedzierski
along the tube. There was no lateral flow of condensate
and Goncalves (1997).
in integral fins because condensate flows down through
Cavallini et al. (2009) developed a correlation for integral
the fins at column locations, and thus the length between
fin tubes by modifying their correlation for plain tubes, Cav-
the columns remains largely condensate free. Therefore,
allini et al. (2006). It is given by the following equations:
heat transfer in lower tubes is not reduced much and
the row effect is small if any. Due to lateral movement of hTP = [h3D + h3A ]1∕3 (2.5.12)
condensate in three-dimensional enhanced tubes, no part
of the tube is completely free of condensate and therefore hA = A ChAS (2.5.13)

Figure 2.5.3 Variation of heat transfer with tube Row number (n)
rows in a bundle of three-dimensional tubes.
Source: From Eckels (2005). © 2005 ASHRAE. 0 10 20 30 40 50
3D tube, pure R-134a 8000
45 000
35 °C, 15 620 W m−2
40 000 7000
ho,tube (W m−2 K−1)

ho,tube Btu/(h ft2 R)

35 000 6000
30 000 5000
25 000
4000
20 000
3000
15 000
10 000 2000
5 000 1000
0 0
0 10 20 30 40 50
Row number (n)
48 2 Heat Transfer During Condensation

Figure 2.5.4 Internally enhanced tubes: (a) helical


microfin tube and (b) herringbone tube. Source: From
Cavallini et al. (2003). © 2003 Elsevier.

(a) (b)

( )0.3685 ( )0.2363
⎛ 𝜌f 𝜇f (2.5.21)
hAS = hLT ⎜1 + 1.28x0.817
⎜ 𝜚g 𝜇g Figure 2.5.5 shows the geometrical parameters in these

( )2.144 equations. D in all expressions above is that at the fin
𝜇f ⎞ tip. hfin is the fin height, ng is the number of grooves,
× 1− Pr−0.1 ⎟ (2.5.14)
𝜇g f ⎟ and nopt is the optimum number of grooves. The param-
⎠ eter Rx accounts for the effect of the heat transfer area
hD = C[2.4x0.1206 (Rx − 1)1.466 C10.6875 + 1]hD,S increase. In Eq. (2.5.11), the heat transfer coefficient is
defined with reference to the cylindrical envelope surface
+ C(1 − x0.087 )Rx hD,S (2.5.15)
area of the finned surface at the fin tip. This correlation
{ [( ) ]}−1 was verified with 3500 data points from many sources
1 − x 0.3321
hD,S = 1 + 0.741 hNu (2.5.16) covering the range: 6 mm < D < 14 mm, 2 l < ng < 82,
x
0.12 mm < hfin < 0.35 mm, hfin /D < 0.04, 25∘ < 𝛾 < 64∘ ,
where hNu is calculated with Eq. (2.4.2), the Nusselt
0∘ < 𝛽 < 30∘ , 0.1 < pr < 0.67, and 90 < G < 900 kg m−2 s−1 .
formula for condensation on horizontal tubes. hLT is
The experimental data used for the validation of the model
calculated by the McAdams/Dittus–Boelter equation for
includes CO2 and halogenated refrigerants. It is not to
single-phase flow, Eq. (2.3.11).
be used for hydrocarbons, ammonia, and water. While
A = 1 + 1.119Fr −0.3821 (Rx − 1)−0.3821 (2.5.17)

G2 A
Fr = (2.5.18)
gD(𝜌f − 𝜌g )2
[ ( ( ))] β
⎧ 𝛾 ⎫ D
⎪ 2h n
fin g 1 − sin 2 ⎪
Rx = ⎨ [ ( )] + 1⎬ ∕ cos 𝛽
𝛾
⎪ 𝜋D cos 2 ⎪
⎩ ⎭ A
(2.5.19) 360°/n
γ
nopt = 4064.4D + 23.257D (2.5.20)
with D in meter.
C = 1 if (nopt /ng ) ≥ 0.8 h
C = (nopt /ng )1.904 if (nopt /ng ) < 0.8
A–A
C1 = 1 if J g ≥ J g,* D
C1 = (J g /J g,* ) if J g < J g,*
−0.3333
⎧[ ]−3 ⎫
⎪ 7.5 −3 ⎪ Figure 2.5.5 Geometrical parameters in the correlation of
Jg,∗ = 0.6⎨ 1.111
+ 2.5 ⎬ Cavallini et al. (2009). Source: From Cavallini et al. (2009).
⎪ 4.3Xtt + 1 ⎪ © 2009 Elsevier.
⎩ ⎭
2.6 Condensation of Superheated Vapors 49

it showed good agreement with its database, a number where hsh is the heat transfer coefficient of superheated
of other correlations were also compared to it and found vapor and hSAT is the heat transfer coefficient of saturated
less accurate. Correlations for Kedzierski and Goncalves vapor. As the sensible heat is usually very small compared
(1997), Han and Lee (2005), Chamra and Mago (2006), and with the latent heat, the increase predicted by Eq. (2.6.1)
Chamra et al. (2005) were found to be fairly good. does not exceed a few percent.
Miyara et al. (2000), Wellsandt and Vamling (2005), and Minkowycz and Sparrow (1966) performed a detailed
Olivier et al. (2007) have given correlations for herringbone analysis of condensation of steam on a vertical plate in
tubes. Aroonrat and Wongwises (2018) have given a corre- which they considered the effect of superheat. Saturation
lation for dimpled tubes to fit their own test data. Data and temperature varied from 27 to 100 ∘ C. Their result was that
limited correlations for many other types of enhanced tubes increase in heat transfer due to superheat does not exceed
have been published. 5%. They also analyzed the case of steam containing air and
found that superheat causes considerably more increase in
heat transfer compared with that for pure vapor.
2.5.3 Recommendations Many experimental studies have confirmed that the
Outside single tubes: Use the Beatty and Katz model for effect of superheat is small. Spencer and Ibele (1966) found
low surface tension fluids for low finned tubes. Use the increases of 4–12%. Walt and Kroger (1972) condensed
Kumar et al. correlation for integral circular and spine R-12 with superheats up to 128 ∘ C on a vertical surface and
fins in the range of their database. This includes fin found that increase in heat transfer coefficients compared
height less than 1.11 mm and fin pitches between 0.5 with saturated vapor by Nusselt equation did not exceed
and 1 mm for halogenated refrigerants and larger than 8%.
2.0 mm for steam. It is concluded that effect of superheat on condensation
Tube bundles: For integral low finned tubes, use Eqs. (2.5.10) heat transfer coefficient can be predicted with sufficient
and (2.5.11) with heat transfer coefficient of the first row accuracy by Eq. (2.6.1).
calculated with a method suited for the tube geometry The total heat flux is that which passes from liquid film
and fluid. For other enhanced tube, use test data and to the wall and is calculated by
correlations suited for the design conditions. q = hsh (T𝑤 − TSAT ) (2.6.2)
Condensation inside tubes: For condensation in microfin
tubes, use the Cavallini et al. correlation (2009) in the Heat transfer from the superheated vapor to the inter-
range recommended by these authors. Outside this face is
range and for various other types of tubes and fluids, use qsensible = hg (Tg − TSAT ) (2.6.3)
the correlations proposed for those geometries in the
range of their database. where hg is heat transfer coefficient between vapor and
the interface. It may be calculated by correlations for
single-phase convection. The condensate rate w from a
plate of height L and unit width is
2.6 Condensation of Superheated L(q − qsensible )
Vapors 𝑤= (2.6.4)
ifg
2.6.1 Stagnant Vapor on External Surfaces
2.6.2 Forced Flow on External Surfaces
As long as the surface temperature is above the satura-
tion temperature of vapor, heat transfer occurs by natural Minkowycz and Sparrow (1969) analyzed laminar con-
convection, and no condensation occurs. Once the wall densation of superheated steam with forced flow on a
temperature drops below the saturation temperature, con- horizontal plate. Saturation temperatures of 27–100 ∘ C and
densation starts. A thermal boundary layer is created in superheats up to 204 ∘ C were considered. For saturation
which the vapor temperature drops to the interface tem- temperature of 27 ∘ C, there was virtually no effect of
perature. If the effects of natural convection in the vapor superheat for ΔT SAT > 2 ∘ C and rose to about 5% above that
boundary layer are neglected, the Nusselt analysis may be for saturated vapor when superheat was 204 ∘ C. At 100 ∘ C
extended by including the de-superheating of vapor. Thus saturation temperature, the increase at 204 ∘ C reached
ifg is replaced by [ifg + Cpg (T g − T SAT )]. This gives about 10%. It is interesting to note that Eq. (2.6.1) predicts
[ ] a 17% increase for this condition. The increase was neg-
hsh ifg + Cpg (Tg − TSAT ) ligible at superheat of 60 ∘ C. They also analyzed the case
= (2.6.1)
hSAT ifg of steam containing air and found that superheat causes
50 2 Heat Transfer During Condensation

considerable increase in heat transfer. This is discussed included using the actual vapor quality instead of the ther-
further in Section 2.7.8 modynamic vapor quality and defining the heat transfer
Michael et al. (1989) condensed steam at atmospheric coefficient based on (T bulk − T w ). Agarwal and Hrnjak
pressure flowing down on a tube of 14 mm diameter. Vapor (2015) condensed R134a, R1234ze(E), and R32 at super-
velocity was 5–81 m s−1 and superheat was up to 40 ∘ C. No heats up to 50 ∘ C in a 6.1 mm tube. They found satisfactory
effect of superheat on heat transfer was found. agreement with the modified Cavallini et al. correlation
mentioned earlier. Xio and Hrnjak (2018) have given a flow
pattern map for superheated vapor.
2.6.3 Flow inside Tubes
Webb (1998) have given a theoretically based meth-
When superheated vapor enters a tube, heat transfer occurs ods for calculation of heat transfer during condensation
by single-phase convection as long as the wall temperature of superheated vapor that is expressed by the following
is above the saturation temperature of vapor; heat transfer equation:
in this region may be calculated by single-phase equations. ( )
Cpg qlat
Once the wall temperature falls below saturation tempera- hsup = hSAT + F hs−ph + (2.6.8)
ture, condensation can start even while the core of vapor is hfg
superheated. McAdams (1954) studied a number of exper- where hsup is the heat transfer coefficient of superheated
imental studies on whose basis he recommends that heat vapor, hSAT is the heat transfer coefficient of saturated
transfer coefficient h in this region be calculated as for sat- vapor, hs-ph is the local single-phase heat transfer coef-
urated vapor and the total heat flux is then calculated as ficient between super-heated vapor and the condensate
q = h(T𝑤 − TSAT ) (2.6.5) interface, qlat is the local heat flux due only to phase
change, and F is a factor defined as
Based on their tests on condensation of steam at
Tg − TSAT
8–180 bar, Miropolskiy et al. (1974) state that even if F= (2.6.9)
TSAT − T𝑤
the wall temperature is below the saturation temperature,
condensation will not start until the vapor temperature According to Webb, this model may be applied to
is below a border temperature T br and quality is below a different types of condensers by using the appropriate
quality xbr defined by correlations to compute the saturated vapor condensation
q heat transfer coefficient hSAT and the single-phase heat
Tbr = TSAT + (2.6.6) transfer coefficient hs-ph .
hGT
qCpg Kondo and Hrnjak (2012c) measured heat transfer of
xbr = 1 + (2.6.7) superheated CO2 and R410A flow in horizontal microfin
hg ifg
tubes of 6 mm inner diameter at reduced pressures of
Specific heat Cpg is calculated at the mean of bulk and 0.55–0.95. They found that condensation of superheated
saturation temperatures. hGT is the heat transfer coeffi- vapor starts when wall temperature reaches saturation
cient of vapor calculated with a suitable forced convection temperature. They proposed a correlation that combines
equation such as Eq. (2.3.11) but its constant doubled. the correlations of Carnavos (1980) for single-phase flow
This was attributed to the variation of properties across and Cavallini et al. (2009) for condensation, both for
the vapor core. They have given a graphical correlation for microfin tubes. It evaluates the liquid properties at film
calculation of heat transfer for qualities between xbr and 1. temperature and was validated with experimental results
Mizoshina et al. (1977) condensed superheated benzene at the reduced pressure ranging from 0.55 to 0.95.
in vertical tubes at vapor Reynolds number between 3000
and 35 000. They concluded that heat transfer coefficient
2.6.4 Plate-Type Heat Exchangers
can be calculated by Eq. (2.6.1).
Kondo and Hrnjak (2011, 2012a, 2012b) condensed Sarraf et al. (2016) studied condensation of pentane in
R-410A and carbon dioxide in a 6.1 mm diameter plain a brazed plate heat exchanger (BPHE) with chevron
tube. Reduced pressure varied from 0.55 to 0.95 and angle of 55∘ and a hydraulic diameter equal to 4.4 mm.
superheats from 0 to 40 ∘ C. They found that condensation Mass flux varied from 9 to 30 kg m−2 s−1 , and the vapor
started when the wall temperature reached the saturation superheat 5–25 K, at a mean constant saturation temper-
temperature. This is in agreement with generally accepted ature of 36.5 ∘ C (1.029 bar). Flow inside the channels was
view and in contrast to that of Miropoloskiy et al. men- simultaneously studied by an infrared camera. Condensa-
tioned earlier. They correlated their data by modifying tion was found to be occurring while the vapor was still
the Cavallini et al. (2006) correlation. The modification superheated. It was found that heat transfer coefficient
2.7 Miscellaneous Condensation Problems 51

1,7 verified range as given in this paper and in Agarwal and


× 10 000 Gr = 9
Gr = 12.3 Hrnjak (2015). For other conditions, use the following
1,5 Gr = 15.5 simple method. Calculate heat transfer coefficient at
Gr = 18.3 x = 0.995 with a correlation for saturated vapor such as
Gr = 21.2
htp (W m−2 K−1)

Shah (2019a). Quality slightly less than 1 is suggested


1,3 Gr = 24
Gr = 27.1 as many correlations crash at x = 1. Then calculate
Gr = 29.8 total heat flux by Eq. (2.6.2). Calculate sensible heat
1,1 flux by Eq. (2.6.3) between vapor and interface with hg
calculated by a single-phase correlation such as that of
0,9
Dittus–Boelter, Eq. (2.3.11), using vapor properties in it.
Condensation inside microfin tubes: Use the correlation of
Kondo and Hrnjak (2012c) in its verified range. Outside
0,7 its range, use the same method as for plain tubes.
0 5 10 15 20 25 30
Condensation in plate heat exchangers: Use the method of
ΔT (°C)
Webb (1998), Eq. (2.6.8).
Figure 2.6.1 Effect of superheat on mean heat transfer
coefficient in a brazed plate heat exchanger. Gr is the mass flux,
kg m−2 s−1 , of condensing fluid pentane. Source: From Sarraf 2.7 Miscellaneous Condensation
et al. (2016). © 2016 Elsevier.
Problems
increased with superheat, up to 70% at 9 kg m−2 s−1 flow 2.7.1 Condensation on Stationary Cone
and 25 K superheat. The effect of superheat decreased
with increasing flow rate. Figure 2.6.1 shows their mea- For a vertical cone, Rohsenow (1973a) analyzed condensa-
surements. They fitted the following equation to their tion of stagnant vapor on a vertical cone in a way similar
data: to the Nusselt analysis for an inclined plate. His expression
for the local heat transfer coefficient for a cone with apex
hTP = [(300891G−3.1 )2 + 73.32 ]0.5 (Tg − TSAT ) + 8491G−0.03 angle (2𝜃) is
(2.6.10) [ ]1∕4
g cos 𝜃𝜌f (𝜌f − 𝜌g )kf3 ifg
Based on study of infrared images, it was concluded that hz = 0.875 (2.7.1)
z𝜇f ΔTSAT
the increase in heat transfer with superheat is caused by its
effect on fluid distribution near the inlet. They found the where z is the distance from top along the slope. The expres-
method of Webb (1998) for calculating the effect of super- sion for mean heat transfer coefficient for a cone of length
heat unsatisfactory for their data. z = L is
[ ]1∕4
Other researchers have also reported increase of heat g cos 𝜃𝜌f (𝜌f − 𝜌g )kf3 ifg
transfer due to superheat. These include Longo (2009, hm = (2.7.2)
L𝜇f ΔTSAT
2011) and Mancin et al. (2012, 2013) with a variety of
fluids. Increase due to superheat in mean heat transfer
coefficients up to 20% were reported. 2.7.2 Condensation on a Rotating Disk
Longo et al. (2015) have given a computational model for
Rohsenow (1973a) analyzed condensation of stagnant
calculating heat transfer during condensation in plate heat
vapor on the upper side of a rotating horizontal disk. By
exchangers. It includes the effect of superheat by using the
equating shear force on the liquid film to the centrifugal
method of Webb (1998). It was verified with data from sev-
force on it, he arrived with the following equation:
eral sources. It is discussed in Section 2.7.5.
[ 2 2 3 ]1∕4
2𝜔 𝜌f kf ifg
hm = (2.7.3)
2.6.5 Recommendations 3𝜇f ΔTSAT
Stagnant vapor on external surfaces: Use Eq. (2.6.1). where 𝜔 is angular velocity in radians per second. He
Forced flow on external surfaces: Neglect effect of superheat. reports that the data of Beatty and Nandkapur (1959)
Condensation inside plain tubes: Condensation to be for condensation of methanol, ethanol, and R-113, at
considered to start when wall temperature reaches sat- atmospheric pressure were 25% below this equation.
uration temperature. For condensation of superheated Sparrow and Gregg (1959a) performed a comprehensive
vapor, use the Kondo and Hrnjak (2011) correlation in its analysis of condensation on a rotating disk taking into
52 2 Heat Transfer During Condensation

consideration shear and centrifugal forces and also taking 2.7.3 Condensation on Rotating Vertical Cone
into consideration momentum and convection terms.
Sparrow and Hartnett (1961) analyzed condensation of a
The model was solved numerically and the results pre-
vertical cone rotating about its axis in stagnant vapor. They
sented graphically. For Prandtl number ≥ 1, these results arrived at the following result for a cone with apex angle
are represented at (Cpf ΔT SAT /ifg ) < 0.1 by the following (2𝜃):
equation: hcone
= (sin 𝜃)0.5 (2.7.5)
hdisk
( )1∕4
h(𝜈f ∕𝜔)1∕2 Prf where hdisk is the heat transfer coefficient of a rotating hori-
= 0.094 (2.7.4) zontal disk given graphically in Sparrow and Gregg (1959b)
kf Cpf ΔTSAT ∕ifg
as well as by Eq. (2.7.4). This analysis took into considera-
tion momentum and convection effects that were neglected
Yanniotis and Kolokotsa (1996) performed tests on in Rohsenow’s analysis in arriving at Eq. (2.7.3).
condensation of low-pressure steam on the underside of Howe (1972) performed theoretical and experimental
a rotating horizontal disk. Saturation temperature was studies on condensation of steam on rotating cones.
40–50 ∘ C. Speeds were up to 1000 rpm. Heat transfer coef-
ficients were independent of the radial distance from the 2.7.4 Condensation on Rotating Tubes
center. The experimental heat transfer coefficients were
Mathews (1962) measured condensation of steam on
only 50–66% of those predicted by Sparrow and Gregg’s
horizontal cylinders of diameter 102, 204, and 254 mm,
formula, Eq. (2.7.4). They report that data from two other
rotating around their own axis. For centrifugal accelera-
studies were 25–55% lower than Eq. (2.7.4), besides the tions up to 2 g, the heat transfer coefficients were similar
data of Beatty and Nandkapur (1959) that were 25% below to the stationary-cylinder values but increased above 2 g
this equation. Hence despite its rigorous derivation, the accelerations.
Sparrow and Gregg formula does not seem reliable. Gacesa (1967) and Nicol and Gacesa (1970) condensed
It would be interesting to compare the data of Yanniotis steam on a 25 mm diameter vertical tube, rotating on its
and Kolokotsa for stationary disk with the formulas of axis. The condensing heat transfer coefficients increased
Gerstmann and Griffith (1967) presented in Section 2.2.4. with speed of rotation, and for the maximum rotational
Visual observation by Yanniotis and Kolokotsa (1996) speed of 2700 rpm investigated, heat transfer coefficients
during operation showed the following. Condensation on were found to be four or five times the stationary value.
the underside of the disk at 0 rpm forms a stationary film. However, for speeds up to about 600 rpm, heat transfer
As the thickness of the film increases, drops are formed coefficients remained unchanged from the stationary
and fall down. When the disk is put in rotation, centrifugal value.
forces act on the film and throw the liquid out toward the Chandran and Watson (1976) condensed methanol, iso-
periphery. The higher the rotational speed, the thinner propanol, ethyl acetate, n-butanol, and water, on horizontal
the film becomes. At 1000 rpm observation using a stro- 19 mm tubes. Some of the tubes were plain, some had pins
boscopic light showed that the condensate forms rivulets on their surface, and some were coated with polytetrafluo-
on the outer area of the disc. Under such conditions it is roethylene (PTFE). Tests were made with tubes stationary
and rotating. Heat transfer coefficients were found to ini-
doubtful that the disk surface will be fully wetted. The low
tially decrease with increasing speed and then rose with
visibility of the Plexiglas wall did not allow more definite
further increase in speed. At a speed of 200 rad s−1 , a three-
observation.
fold increase in heat transfer was obtained. Pins and PTFE
Alasadi et al. (2013) studied condensation of steam at sat-
coatings increased heat transfer, and it was attributed to bet-
uration temperature of 40–50 ∘ C on a rotating disk. They
ter drainage of condensate. A dimensionless correlation of
found heat transfer coefficients to increase with rotational their own data was developed but there was large scatter.
speed. They developed an analytical formula that agrees They did not compare their correlation with data other than
fairly well with their own data. They did not compare their their own. They compared the predictions of their correla-
data with the Sparrow–Gregg formula, nor did they com- tion with those proposed by others. The predictions of the
pare their formula with any other data. various correlations differed greatly.
Rifert et al. (2018) have reviewed the literature on boiling In conclusion, there is no prediction method that has
and condensation on rotating disks. However, they have not been verified with data other than the authors’ own. The
offered any resolution of the differences between test data data and correlations in these references may be used for
and theory. guidance in design.
2.7 Miscellaneous Condensation Problems 53

2.7.5 Plate-Type Condensers multiplied by the enlargement factor 𝜙 to get the heat trans-
fer coefficient based on projected area S in the gravity con-
The most widely used plate heat exchangers have her-
trolled regime, hgrav,m . Thus,
ringbone pattern plates. Their geometry is shown in
Figure 2.7.1. To ensure leak-tightness, condensers have hgrav,m = 𝜙hNu,m (2.7.8)
brazed plates, and these are therefore known as BPHE.
For Reeq > 1600, regime is of forced convection and the
Many correlations have been proposed for condensation
mean heat transfer coefficient hFC is
heat transfer in BPHE based on limited data. Among these
1∕3
are Han et al. (2003), Yan et al. (1999), and Kuo et al. (2005). hFC = 1.875𝜙(kf ∕DHYD )Re0.445
eq Prf (2.7.9)
A correlation based on data from many sources has been As was discussed in Section 2.6.5, superheated vapor
presented by Longo et al. (2015). In this correlation, there begins to condense if wall temperature falls below the
are two heat transfer regimes, a gravity dominated regime saturation temperature and that Webb (1998) gave a model
and a forced convection regime. The boundary between the for heat transfer in this region. That model is expressed
two regimes is at Reeq = 1600, where Reeq is an equivalent by Eq. (2.6.8) given in Section 2.6.3. Longo et al. have
Reynolds number defined as incorporated the Webb model into the above correlation.
Local single-phase heat transfer coefficient of vapor hs-ph
Reeq = G[(1 − x) + x(𝜌f ∕𝜌g )1∕2 ]DHYD ∕𝜇f (2.7.6) is calculated by the following equation given by Thonon
(1995):
The mean of the inlet and outlet quality is used for x. 1∕3
hs−ph = 0.2267(kg ∕DHYD )Re0.631
GS PrG (2.7.10)
DHYD = 2b, where b is the height of corrugation.
The heat transfer coefficient calculated in this correlation They compared this correlation with their own data as
well as those from six other sources. The data included sev-
are based on the projected area S of active plates given by
eral fluids (halogenated refrigerants, propane, and carbon
dioxide), chevron angles 30∘ –65∘ , (measured from the ver-
S = NWLp (2.7.7) tical axis) and mass flux 2.6–42 kg m−2 s−1 . The MAD for all
data was 16%.
where N is the number of effective plates. At Reeq < 1600, Zhang et al. (2019) condensed R1234ze(E) and R1233zd(E)
heat transfer coefficient hNu,m is calculated with the Nus- in a plate heat exchanger with chevron angle of 65∘ . Satura-
selt formula for condensation on vertical plates, Eq. (2.2.8), tion temperatures ranged from 30 to 70 ∘ C, mass flux from
with Lp replacing L. This heat transfer coefficient hNu,m is 16 to 90 kg m−2 s−1 . They found satisfactory agreement

Figure 2.7.1 Corrugated plate heat exchanger. + Cross section


Note that in many publications, chevron angle normal to
𝛽 is measured from the vertical axis. Source: + + direction of
From ASHRAE (2017). © 2017 ASHRAE. troughs
Developed
length

Protracted
length
β
Lp
t
W
b p
+ +

+ +
λ
+

β = Chevron angle
ϕ = Enlargement factor = Developed length/protracted length
λ = Corrugation pitch
54 2 Heat Transfer During Condensation

with the Longo et al. correlation, while several others The most commonly used refrigerants in the past and at
had large deviations. They also gave a correlation of their present are halogenated type. These are mostly miscible
own data. with and soluble in oil. Shen and Groll (2005) reviewed
Shon et al. (2018) condensed R-1233zd(E) in a BPHE the research on condensation of such refrigerants. They
with chevron angle 60∘ . They compared their data with concluded that most studies show that condensation heat
correlations of Kuo et al. (2005), Yan et al. (1999), and Han transfer decreases with increasing oil content though its
et al. (2003). These were found unsatisfactory so they gave extent varies. Tichy et al. (1984) condensed R-12 with oil in
a correlation of their own data. a horizontal tube. Oil caused deterioration of heat transfer,
Eldeeb et al. (2016) have reviewed experimental studies which they expressed by the following formula:
and predictive techniques for plate condensers. hwith oil
= exp(−5xoil ) (2.7.11)
hno oil
2.7.5.1 Recommendation where xoil is the mass fraction of in the mixture.
Longo et al. correlation is recommended in its verified Sur and Azer (1991) condensed refrigerant–oil mixtures
range. Note that it has not been verified for water or in smooth and finned tubes. They found reductions in heat
ammonia. Their properties differ considerably from the transfer coefficients of 7%, 12%, and 16%, with oil concen-
fluids for which Longo et al. correlation was verified. trations of 1.2%, 2.8%, and 4.0% in R-113, respectively.
Some studies also showed no effect of oil. Williams and
2.7.6 Effect of Oil in Refrigerants Sauer (1981) condensed R-11 on a horizontal tube. Two
types of oil were used: SUS 150 and SUS 500. Oil concen-
In most compressors used in refrigeration and air condi- trations up to 7% had no effect on heat transfer but higher
tioning, some lubricating oil gets mixed with refrigerant concentrations caused deterioration.
during compression and is carried into the system. While Eckels (2005) studied the effect on condensation of
the larger systems use oil separators, they are not 100% R-134a on horizontal banks of smooth and enhanced
effective. Small systems do not have an oil separator at all. tubes. Two types of oils were used, ISO-68 and ISO-120. Oil
Oil is thus carried into the condenser where it can affect its concentrations up to 12.4% were used. It was found that
performance. ISO-120 oil had essentially no effect on heat transfer, while
Depending on the type of oil, refrigerants may be mis- ISO-68 oil reduced heat transfer by up to 25%. The only
cible or immiscible with them. Immiscible oils were notable difference between the two oils is that the viscosity
exclusively used in ammonia systems in the past and are of ISO-120 oil is about twice that of ISO-68 oil. Eckels had
also presently used in ammonia systems with flooded no explanation for the different behavior with the two oils.
evaporators. Mirmov and Yemelyanov (1976) state that Some prediction methods for miscible refrigerant–oil
in the USSR, it was generally believed that oil films of mixtures are now discussed. Many authors have given
0.05–0.08 mm were formed on condensing surface and this empirical correlations based on their own data. Equation
resistance was added in the design of condensers. However, (2.7.11) is an example of such correlations. Huang et al.
Abdukmanov and Mirmov (1971) state that they found in (2010) compared many of these correlations with their data
their experiments on horizontal tubes that oil increased for R-410A in 4.18 and 1.6 mm diameter tubes. All of them
the heat transfer coefficient by 30%. Mazukewitch (1952) gave large errors and so they gave their own correlation of
found that oil reduced heat transfer in a vertical annulus their data.
by 30%. Short (2017) states that normal naphthenic or A number of authors have taken the approach that the
paraffinic lubricants and synthetic hydrocarbon oils have effect of oil on heat transfer is due to change in fluid
low solubility and miscibility in ammonia. These oils are properties. Thome (1998) recommended the use of pure
heavier than ammonia and tend to form an oil film on the fluid correlations but using the viscosity of oil–refrigerant
heat transfer surfaces. Shah (1975) studied evaporation mixture in place of the pure refrigerant viscosity. Several
of ammonia in a horizontal tube. Oil films were visually authors have recommended use of pure refrigerant con-
observed in both single-phase liquid flow and during evap- densing correlations but using mixture properties. Among
oration. Single-phase heat transfer coefficients were lower them are Sur and Azer (1991), Schlager et al. (1990), and
than predicted by the Dittus–Boelter equation, Eq. (2.3.11). Shao and Granryd (1995). These authors were able to cor-
The difference in heat transfer was equivalent to the relate their own data by this method. No comprehensive
resistance oil films 0.04–0.1 mm. So, most of the evidence comparison of any method with data from many sources
indicates that insulting oil films can form with immiscible could be found.
oils in ammonia systems and it will be advisable to take it It should be noted that the composition of mixture as
into account when sizing the condensers. well as the saturation temperature change as condensation
2.7 Miscellaneous Condensation Problems 55

progresses. Many methods for prediction of mixture prop- A similar analysis can be done for condensation of
erties have been proposed. Shen and Groll (2005) have stagnant vapor on a flat plate as follows. Let y be the
discussed them and made recommendations for their thickness of condensate layer at time t. Taking heat
choice. Calculations with this method are quite complex. balance,
Considering that the actual amount of oil is unknown and kf dy
that no comprehensive verification of this method has = 𝜌f ifg (2.7.14)
𝛿 dt
been done, such calculations may be worthwhile only in Integrating with y = 0 at t = 0 and y = 𝛿 at time t,
exceptional situations, perhaps to get an idea of trends. ( )0.5
As the oil content in normally operating systems is small, 2kf t
𝛿= (2.7.15)
assumption of a reduction of 5–10% should be adequate for 𝜌f ifg
design purposes. The heat transfer coefficient h is
( )
kf kf 𝜌f ifg
2.7.6.1 Recommendation h= = (2.7.16)
For immiscible oils, assume an oil film 0.04–0.08 mm thick. 𝛿 2t
Add its resistance to that of pure fluid to calculate the effec-
2.7.7.2 Experimental Studies
tive heat transfer.
Azzolin et al. (2018) measured heat transfer during con-
When refrigerants and oils are miscible, assume a 5–10%
densation of HFE-7000 in a horizontal 3.4 mm diameter
decrease in heat transfer due to the effect of oil.
tube. Tests were done in normal gravity as well as in
microgravity achieved in parabolic flights. The result was
2.7.7 Effect of Gravity that at 170 kg m−2 s−1 flow, heat transfer coefficients were
the same in both gravity conditions. At lower flow rates,
In recent years there has been much interest in space travel heat transfer coefficients in microgravity were lower, the
and space colonization. Low to zero gravity occurs under difference increasing with decreasing mass flow rates. At
those conditions, and it is important to know their impact 70 kg m−2 s−1 , heat transfer coefficients were 10–24% lower
on heat transfer. at various qualities under microgravity compared with nor-
It is easy to see that the absence of gravity will have a huge mal gravity. Flow patterns were observed simultaneously.
impact and many of the predictive techniques used under It was seen that at 170 kg m−2 s−1 , there was uniform liquid
terrestrial gravity will be inapplicable. For example, Nusselt layer around the tube circumference at both gravity levels.
analysis for condensation on horizontal and vertical sur- At lower flow rates, liquid film was thicker at the bottom
faces is based on drainage of condensate by gravity. In the and thinner at top at normal gravity. At microgravity, liquid
absence of gravity, condensate will not flow and hence Nus- film remained uniform around the circumference at all
selt formulas will be inapplicable. During condensation in flow rates. The thinner liquid film at top under normal
horizontal tubes, stratification occurs due to gravity. In the gravity results in less thermal resistance and hence higher
absence of gravity, there will be no stratification, and heat heat transfer. It is gravity that causes stratification and
transfer will be the same in all orientations. there is no stratification in its absence.
Lee et al. (2013) studied condensation of FC-72 at micro-
2.7.7.1 Some Formulas for Zero Gravity gravity as well as gravity at Mars and the moon. Heat
Rohsenow (1973a) analyzed condensation from stagnant transfer measurements were in a horizontal tube while
vapor on a tube at zero gravity. Condensate does not drain flow observations were made in a horizontal annulus.
away. Instead, it forms a uniform layer around the tube Liquid film thickness was uniform at microgravity while at
whose thickness grows with time. Heat transfer in liquid Lunar and Martian gravities, liquid film was thicker at the
occurs by conduction. The thickness of condensate 𝛿 = 0 at bottom for lower flow rates. They did not compare their
time t = 0. By taking an energy balance and integrating, the heat transfer measurements with any correlation. They
following expression was obtained for value of 𝛿 at time t: developed a control volume-based model that had MAD of
( ) ( ) 2kf ΔT 25.7% with their data.
R+𝛿 2 R+𝛿 1 1
ln − + = t (2.7.12) Mudawar (2017) has reviewed the literature on micro-
R R 2 2 𝜚f ifg R2
gravity condensation.
R is the radius of tube. Heat transfer coefficient h is
given by 2.7.7.3 Conclusion
kf Amount of available experimental data for condensation
h= ( ) (2.7.13) in microgravity conditions is very limited and there is no
R+𝛿
R ln R well-verified method for calculating heat transfer.
56 2 Heat Transfer During Condensation

2.7.8 Effect of Non-condensable Gases non-condensables greatly reduces heat transfer. The effect
is much greater in stagnant vapor than in forced flow.
Presence of non-condensable gases reduces heat trans-
For steam at 100 ∘ C, 2% mass content of air reduces heat
fer. In the pioneering study by Othmer (1929), presence
transfer by over 70% for stagnant vapor and by 10% for
of 0.5% air in steam was found to reduce heat transfer
forced convection. Minkowycz and Sparrow (1966) also
by 50%. Similar results have been found in numerous
studied the effect of superheat. For pure steam, superheat
subsequent experimental studies as well as theoretical
had essentially negligible effect. For steam–air mixtures,
analyses. The reason for this is explained by considering
there was considerable effect of superheat. For stagnant
condensation of steam–air mixture on a vertical plate as
steam at 100 ∘ C saturation and 2% air, heat transfer was
shown in Figure 2.7.2. As steam condenses on the plate
about 40% higher at 222 ∘ C superheat.
surface, air is left behind. A layer of air–vapor mixture of
thickness 𝛿 is formed on the condensate layer. Steam vapor 2.7.8.1 Prediction Methods
has now to diffuse through this air-rich layer to reach the Colburn and Hougen (1934) presented a general method to
interface. The partial pressure of steam pgo drops through calculate condensation in the presence of non-condensable
this layer, while that of air pao rises through this layer. that was further developed by later researchers. It may be
The total pressure at the condensate surface remains the written as
same as in the region outside this layer. The temperature of
the liquid–vapor interface is at the saturation temperature hc (Ti − T𝑤 ) = 𝜍hg (Tg − Ti ) + Gc ifg (2.7.17)
T gi corresponding to the partial pressure of vapor at the T i is the interface temperature and T g is the bulk vapor–gas
interface. Heat transfer is reduced due to the mass diffusion mixture temperature. 𝜍 is the Ackerman (1937) constant
resistance and also because air has to be cooled from the that corrects for mass transfer effects. hc is the heat trans-
bulk temperature to the interface temperature. fer coefficient for condensation of pure vapor, and hg is the
Comprehensive laminar boundary layer analyses were single-phase convective heat transfer coefficient between
done by Minkowycz and Sparrow (1966) for condensa- mixture and interface. 𝜍 is given by
tion from stagnant vapor and Sparrow et al. (1967) for
forced convection on plates. They found that presence of 𝜍 = 𝜆∕(1 − e−𝜆 ) (2.7.18)

𝜆 = Gc Cpg ∕hg (2.7.19)


p The mass flux of condensate Gc is given by
[ ]
pgo
p − pg,i
pai Gc = KM 𝜚g ln (2.7.20)
pg p − pg,b
where p is the total pressure, pg,b is the partial pressure of
pgi vapor in the bulk mixture, and pg,I is the partial pressure
pa
of vapor at the liquid–gas interface. The mass transfer coef-
pao ficient K M may be calculated by the analogy between heat
and mass transfer, empirical correlations, or by other meth-
ods. Iterative solution of the above listed equations together
Tgo
with an equation for heat transfer between wall and coolant
will yield the interface temperature T i . This method can
δ
be used for external as well as internal condensation with
appropriate choice of for hc and hg .
Tgi
Many empirical correlations have been proposed. An
early one is by Meisenburg et al. (1935) for condensation of
steam–air outside a vertical tube. Ge et al. (2016) obtained
data on a vertical plate, with the average vapor velocity
of 1.2 m s−1 , CO2 mass fraction in steam of 20–94%, and
a pressure of 1 atm. They found the Meisenberg et al.
correlation and several others to greatly underpredict their
data. They fitted the following equation to their own data:
Figure 2.7.2 Effect of non-condensable gas on condensation of
vapor on a vertical surface. Source: From Collier and Thome hm 1.9864 − 0.498 ln(C + 16.2724)
(1994). © 1994 Oxford University Press. = (2.7.21)
hNu (Tg − T𝑤 )0.18
2.7 Miscellaneous Condensation Problems 57

where C is the mass percentage of carbon dioxide in the 2.7.9 Flooding in Upflow
mixture.
The correlations for heat transfer presented in Section 2.3
A number of authors have given correlations in terms of
are applicable only past the flooding point. Hence predic-
a degradation factor defined as the ratio of heat transfer of
tion of flooding point is important.
mixture to that of pure vapor. Vierow and Shrock (1991),
To understand the flooding phenomenon, consider a
Kuhn et al. (1997), and Siddique et al. (1993) gave correla-
vapor entering at very low velocity into the bottom of a
tions based on their data for steam–gas mixtures in vertical
vertical tube at wall temperature below the saturation
tubes. Lee and Kim (2008) found them to be unsatisfac-
temperature of vapor. Condensate forms on the wall and
tory on comparison of their own data for steam–nitrogen
flows down while the uncondensed vapor flows upwards.
mixtures in a 13 mm diameter vertical tube. They gave the
As velocity of vapor is increased, pressure drop rises,
following correlation that fitted their own data as well as
liquid continues flowing down though some droplets are
those of Siddiqui et al. and Kuhn et al.:
carried away by the vapor. As vapor velocity continues to
hmixture increase, a point is reached where pressure drop reaches
= 𝜏mix
0.3124 0.402
(1 − 0.964Cnc ) (2.7.22)
hNu a maximum, downwards flow of condensate ends, and
For 0.06 < 𝜏 mix < 46.65, 0.038 < Cnc < 0.814. Cnc is the condensate begins to move upwards. This is the flooding
mass fraction of non-condensable gas and 𝜏 mix is the dimen- point and the vapor velocity at this point is called flooding
sionless shear stress calculated as follows: velocity. With further increase in vapor velocity, pressure
0.5𝜌mix u2mix f drop falls though liquid keeps moving upwards.
𝜏mix = (2.7.23) There are several possible mechanisms that may be the
g𝜌f L
cause of flooding. These are interfacial vapor shear, entrain-
Remix 𝜇mix
umix = (2.7.24) ment of liquid in vapor stream, formation of large waves in
𝜌mix D
liquid near the bottom of tube that may prevent downflow
( 2 )1∕3
𝜈f of liquid, and Bernoulli effect at tube entrance caused by
L= (2.7.25) high vapor velocity that causes liquid drops to be sucked
g
upwards.
The friction factor f is given by Many prediction methods, theoretical and empirical,
For Remix > 2300, have been proposed. Wallis (1969) considered flooding to
0.079 occur due to vapor shear and gave the following correlation:
f = (2.7.26)
Re0.25
mix ∗1∕2 ∗1∕2
uLS + uGS = C (2.7.28)
For Remix < 2300,
16 Wallis gave C = 0.775. Other values of C (0.88–1) have
f = (2.7.27)
Remix been proposed by other researchers as noted by McQuil-
lan and Whalley (1985). In this equation uLS * and uGS * are
The range of data covered by this correlation includes
dimensionless velocities for liquid and gas flowing alone,
tube diameters 13–49.5 mm, inlet non-condensable gas
defined as
mass fraction up to 40%, pressure 0.03–0.5 MPa, and steam
flow 6–62 kg m−2 s−1 . ( )1∕2
𝜌 f
Caruso et al. (2013) performed tests with steam–air mix- u∗LS = uLS (2.7.29)
gD(𝜌f − 𝜌g )
tures in a tube inclined at 7∘ to the horizontal and gave
an empirical correlation of their data. Xu et al. (2016) per- ( )1∕2

𝜌g
formed experiments with condensation of steam–air mix- uGS = uGS (2.7.30)
gD(𝜌f − 𝜌g )
ture in horizontal 25 mm diameter tube. The correlation of
Caruso was found to overpredict their data. They have given McQuillan and Whalley (1985) compared data for flood-
a correlation to fit their own data. ing in gas–liquid and steam–water systems from many
Various correlations and theoretical methods have been source to a number of correlations. Best agreement was
reviewed by Huang et al. (2015). found with the correlation of Alekseev et al. (1972). They
modified it to the following equation, which gives even
2.7.8.2 Recommendation better agreement:
The Lee and Kim correlation may be used for vertical tubes
( )−0.18
in its verified range. Colburn and Haugen method is recom- 𝜇f
mended for general use. Kg∗ = 0.286Bd0.26 Fr ∗−0.22 (2.7.31)
𝜇water
58 2 Heat Transfer During Condensation

𝜇 water is the viscosity of water, taken to be 0.001 N s m−2 K g * It may be concluded that Nusselt equations can be used
is known as Kutateladze number defined as with confidence for ReLS > 10. For lower Reynolds num-
𝜌0.5
g
bers, Eq. (2.7.35) may be used but with caution as it has not
Kg∗ = Jg (2.7.32) had validation with other data.
[𝜎g(𝜌f − 𝜌g )]0.25
Bd is the Bond number defined by
D2 g(𝜌f − 𝜌g ) 2.7.11 Condensation in Helical Coils
Bd = (2.7.33)
𝜎 Wongwises and Polsongkram (2006) condensed R-134a in
Fr * is a dimensionless Froude number defined as a coil with vertical axis. Tube inside diameter was 8.3 mm
( ) 0.25
g (𝜌f − 𝜌g )0.25 and coil diameter 305 mm and coil pitch was 35 mm. Sat-
∗ 𝑤
Fr = (2.7.34) uration condensing temperatures were from 40 to 50 ∘ C
𝜋D𝜌f 𝜎 0.75
and mass flux was 400–800 kg m−2 s−1 . They found that
where w is the mass flow rate of liquid. coil average heat transfer coefficients were 33–53% higher
Palen and Yang (2001) reviewed a number of flooding than for horizontal tubes. They correlated their data by the
correlations. They suggested a different form of correlation following equation:
but did not give its details, nor its validation.
Berrichon et al. (2015) compared their data for air–water hTP
= 0.1325De0.7654
eq Pr0.8144
f
Xtt0.0432 p−0.3356
r (Bo × 104 )0.112
in a 54 mm diameter tube with several correlations. The Dtube kf
Wallis correlation with C = 0.88 gave the best agreement. (2.7.36)
The McQuillan and Whalley correlation also gave satis-
Deeq is equivalent Dean number defined as
factory agreement. They also did tests with low pressure
steam. ( )( )0.5 ( )0.5 ⎤
⎡ 𝜇g 𝜌f Dtube
The McQuillan and Whalley correlation was vali- Deeq = ⎢ReLS + ReGO ⎥
dated with a very wide-ranging database that included ⎢ 𝜇f 𝜌g Dcoil ⎥
⎣ ⎦
condensing steam. It is therefore recommended. (2.7.37)

2.7.10 Condensation in Thermosiphons Bo is the boiling number defined as [q/(Gifg )], where q is
the heat flux.
Closed vertical thermosiphons are considered in this Gupta et al. (2014) measured heat transfer coefficients
section. In such thermosiphons, evaporation occurs at the during condensation of R-134a in a horizontal coil. Tube
bottom. The vapor rises up to the section with cold wall diameter was 8.33 mm and coil diameter was 90.48 mm.
where it condenses. The condensate flows down back into Mass flux was 100–350 kg m−2 s−1 and evaporation tem-
the evaporation section and thus the cycle continues. perature was 35–40 ∘ C. They found satisfactory agreement
Shiraishi et al. (1981) tested a closed vertical ther- with the correlation of Wongwises and Polsongkram (2006)
mosiphon formed by a 37 mm diameter tube with water, given earlier. They also compared their data with the Shah
ethanol, and R-113. Condensation heat transfer agreed (1979) correlation for straight tubes. Measurements were
fairly well with Nusselt equation. mostly 1.5–2.0 times the predictions. Deviations increased
Jouhara and Robinson (2010) studied a 6 mm diame- with decreasing mass flux. As was discussed in Section
ter vertical thermosiphon with water, FC-84, FC-77, and 2.3.3.2, the Shah (1979) correlation is inaccurate at low
FC-3283. In the condensation region, agreement was good mass flux and has therefore been replaced by later versions,
with Nusselt equation at ReLS > 10, but it over predicted
the latest being Shah (2019a).
at lower ReLS . Hashimoto and Kaminaga (2002) had also
Salimpour et al. (2017) condensed R-404A in horizontal
found similar behavior in their tests on a thermosiphon.
coils. Tube diameter was 7.52 mm, while coil diameter was
They attributed it to entrainment of liquid that thinned
from 87 to 153 mm. Coil pitch varied from 15 to 35 mm.
out liquid film and gave a formula to fit their own data.
Mass flux was from 125 to 188 kg m−2 s−1 and satura-
Jouhara and Robinson found it to somewhat underpredict
tion temperature was 29–39 ∘ C. They fitted the following
their data and modified it to the following formula:
equation to their data:
( )
𝜌f ( ) ( )
0.1
hm = 0.85ReLS exp −0.000 067 − 0.6 hNu Pcoil −0.0355 kf
𝜌g 0.8934 0.9744 0.0832
hTP = 0.086Deeq Prf Xtt
𝜋Dcoil Dtube
(2.7.35)
(2.7.38)
where hNu is the prediction of Nusselt formula, Eq. (2.2.8).
Pcoil is the pitch of the coil.
2.8 Condensation of Vapor Mixtures 59

Mozafari et al. (2015) condensed R-600a in horizontal cannot be calculated by methods for pure vapor. Their heat
and inclined coils. Tube diameter was 8.3 mm. The coil transfer coefficients are lower than that of the components.
diameter, pitch, height, and the number of coil turns were Figure 2.8.1 shows an example. Condensation of zeotropic
305 mm, 35 mm, 210 mm, and 6, respectively. Experiments mixtures of vapors of miscible fluids is discussed herein.
were also done in a straight tube. Tests were performed The boiling point of a mixture of a particular composition
at average saturation temperatures ranging between 38.5 at a particular pressure is different from the condensation
and 47 ∘ C. Refrigerant mass fluxes varied in the range of point of a mixture of the same composition. The former is
155–265 kg m−2 s−1 . Inclinations of coil axis to horizontal known as the bubble point, while the latter is known as the
were 0∘ , 30∘ , 60∘ , and 90∘ . The highest and lowest values of dew point. The dew point is higher than the bubble point.
heat transfer coefficient occurred at 30∘ and 90∘ inclination The difference between the two is known as temperature
angles, respectively. Heat transfer coefficients in the coils glide or simply glide.
were 1.24–2.65 times those in the straight horizontal tube. As condensation proceeds, the less volatile (that is with
Al-Hajeri et al. (2007) have reported data for condensa- higher boiling point) component condenses at a more rapid
tion of R-134a in a coil but have not compared it to any rate. This results in the vapor mixture getting richer in the
correlation. more volatile (that is with lower boiling point) component.
Yu et al. (2018) condensed propane in a tube bent into a As a result, the dew point temperature falls. With more and
horizontal half coil with helix angle of 10∘ . Tube diameter more condensation, dew point temperature keeps falling.
was 32 mm, while the coil diameter was 2 m. Shah (2019a) When condensation is completed, the composition of the
found their data to be in good agreement with his correla- liquid is the same as of the original vapor. Sensible heat
tion. The larger the coil diameter compared to tube diam- has to be removed from both vapor and liquid as their
eter, the more the coil resembles a straight tube. The coil temperature falls from the initial dew point temperature to
diameter to tube diameter ratio in these tests was 62. It indi- the bubble point temperature. As the composition of vapor
cates that for this ratio and higher ratios, straight tube cor-
relations can be used for coils. 30
The preceding discussions show that heat transfer coef- CO2%
ficients can be considerably higher than in straight tubes. 20 0% 21% 39%
Ratio of pitch to coil diameter and ratio of tube diameter
10 200 (kg m−2 s−1)
to coil diameter affect heat transfer. While several correla-
tions for heat transfer have been proposed, they have not
been verified beyond the data on which they were based. 30

Therefore, these correlations should be used only for the


20
data range on which they were based. For Dcoil /Dtube ≥ 62,
300 (kg m−2 s−1)
straight tube correlation may be used.
10
αR (kW m−2 K−1)

30
2.8 Condensation of Vapor Mixtures
20
Condensation of vapor mixtures is required in many indus- 450 (kg m−2 s−1)

tries that include chemical, petroleum, and refrigeration. 10


In recent years, mixtures are being increasingly used as
refrigerants because the traditional refrigerants such as 30
R-12 were found to harm the ozone layer and cause global
warming. 20 500 (kg m−2 s−1)

10
2.8.1 Physical Phenomena
Mixtures of vapors may be azeotropic or zeotropic (also 0
0.0 0.2 0.4 0.6 0.8 1.0
called non-azeotropic). Azeotropic mixtures behave like
1–x
single-component fluids and their heat transfer may be
calculated by the same methods as for pure vapors using Figure 2.8.1 Effect of composition of mixture on heat transfer.
the mixture properties. In zeotropic mixtures, the compo- CO2 /DME mixture condensing in a 4.35 mm diameter horizontal
nents condense at different rates and their heat transfer tube. Source: From Afroz et al. (2008). © 2008 Elsevier.
60 2 Heat Transfer During Condensation

and liquid is varying, mass transfer resistance occurs in tubes, Del Col et al. (2005) have presented two methods
both vapor and liquid phases. The resistance in the vapor applicable to different flow pattern groups. For the inter-
phase has much more effect. Heat transfer of mixtures is mittent, annular, and mist flow patterns, they reasoned that
lowered compared to pure vapor because of these sensible the heat transfer coefficient of the vapor phase is increased
heat transfer and mass transfer effects. because of interfacial waves and interfacial shear. To take
these factors into account, they modified Eq. (2.8.3) to the
2.8.2 Prediction Methods following:
1 1 q ∕q
The classical method used for correcting the performance = + sens tot (2.8.5)
heff hc fi hga
of binary mixtures is the film theory method proposed by
Colburn and Drew (1937), which subsequently has been where hga is the heat transfer coefficient of vapor phase in
extended to multicomponent mixtures. Such methods are the tube cross section occupied by vapor and is calculated
described in Taylor and Krishna (1993). These methods using the vapor velocity through the area occupied by the
are quite tedious, especially when there are more than two vapor. The interfacial friction factor, f i , is calculated using a
components. Bell and Ghaly (1973) presented a simple number of equations that include those for the calculation
method for calculating heat transfer during condensation of flow patterns and void fraction. Their correction factor
of mixtures that has been widely used. for the stratified and stratified-wavy flow patterns requires
Bell and Ghaly assumed that the mass transfer resistance that heat flux be known.
is approximately equal to the sensible heat transfer resis- According to McNaught (1979), the mass transfer resis-
tance of the vapor phase. Sensible heat is transferred from tance can be significantly higher than that calculated by the
the vapor to the interface by single-phase convection. The Bell and Ghaly method (1973). McNaught replaced hg in
sum of sensible and latent heats is transferred through the Eq. (2.8.3) by hg,mod given by the following equation:
liquid film to the wall. Sensible heat flux from vapor to liq-
hg 𝜙
uid interface is hg,mod = (2.8.6)
(exp𝜙 − 1)
qsens = hg (Tg − Ti ) (2.8.1)
where 𝜙 is given by
where T i is the interface temperature and hg is the

n
mi Cpg,i
single-phase heat transfer coefficient of vapor calcu- 𝜙= (2.8.7)
lated by single-phase correlations suitable for the geometry i=1
hg
and conditions ignoring the presence of liquid. The total where m is the individual mass flux of a component. The
heat flux is subscript i indicates for component i. The summation is car-
qtot = hc (Ti − T𝑤 ) (2.8.2) ried out for all n components of the mixture. The predicted
heat transfer coefficient by the McNaught method (1979)
where hc is the condensing heat transfer coefficient to be is always equal or lower than that by the Bell and Ghaly
calculated with a pure vapor correlation. Eliminating T i method (1973).
between these two equations, the following expression for Shah et al. (2013) analyzed a database for condensation of
the effective heat transfer coefficient heff is obtained: inside tubes that included 529 data points from 22 studies,
1 Tg − T𝑤 1 q ∕q for 36 fluid mixtures, and temperature glides up to 35.5 ∘ C
= = + sens tot (2.8.3)
heff qtot hc hg for a wide range of reduced pressures and mass flow rates.
These data were compared to the Shah (2009) correlation
Note that hg and hc are calculated using properties of mix-
for pure vapors together with the corrections factors by
ture. hg is calculated assuming liquid is not present. Thus
Bell–Ghaly, Del Col et al., and McNaught. The correction
for condensation in tubes, hg is the superficial heat transfer
factors of Bell–Ghaly and McNaught gave MAD of 17.7%
coefficient assuming that the gas phase occupies the entire
and 17.6%, respectively. The correction factor by Del Col
cross section. The ratio of sensible to total heat transfer may
et al. was evaluated in its heat flux-independent regime
be written as
and had MAD of 20.1%. The deviations of Bell–Ghaly and
qsens xCpg ΔTDP
= (2.8.4) McNaught methods are about the same. Also, the Del
qtot ΔH Col et al. method is applicable only to horizontal tubes.
where ΔH is the change in enthalpy of mixture and ΔT DP It should also be noted that the Bell–Ghaly method has
is the change in dew point temperature. been evaluated by numerous researchers with mostly
Several authors have proposed modifications of the good results and is applicable to all types of condensers.
Bell–Ghaly method. For condensation inside horizontal There has not been much independent verification of the
2.9 Liquid Metals 61

McNaught method other than that by Shah et al. (2013) for Bonilla (1956) with mercury and sodium, Cohn (1960) with
condensation in tubes mentioned earlier. cadmium and mercury, and Roth (1962) with rubidium.
The heat transfer coefficients measured in those studies
2.8.3 Recommendation were found to be much lower than Nusselt theory as well
as modifications of Nusselt theory given in the compre-
Use the Bell–Ghaly method for miscible mixtures for all hensive theoretical analyses by Chen (1961a), Sparrow and
types of condensers. Gregg (1959a), and Koh et al. (1961). Figure 2.9.1 shows
the findings of Sukhatme and Rohsenow. It is seen that all
the data are much lower than the theoretical values.
2.9 Liquid Metals Other studies on condensation of liquid metal vapors
include Sukhatme and Rohsenow (1966) with mercury,
2.9.1 Stagnant Vapors Bakulin et al. (1967) with sodium, Ivanovskii et al. (1968)
Results of several studies on condensation from stag- with mercury, and Subbotin et al. (1964) with potassium.
nant liquid metal vapors were studied by Sukhatme and All of them showed heat transfer coefficients much lower
Rohsenow (1964, 1966). These included those by Misra and than theory.

Figure 2.9.1 Data from some No. investigator Liquid metal condensed Coolant used
early studies on condensation of
1 Misra; Bonilla Mercury Water
stagnant liquid metal vapors 2 Misra; Bonilla Mercury Air
compared to Nusselt theory and 3 Cohn Mercury Air
its modifications. 𝜆 is the latent 4 Cohn Cadmium Air
heat. Source: From Sukhatme and 5 Roth Rubidium Air
Rohsenow (1966). © 1966 6 Misra; Bonilla Sodium Air
American Society of Mechanical Lines drawn through some of the data are correlations
Engineers. suggested by the respective authors

10

Pr = 0.03
Nusselt’s
theory
1
1/4

2
0.08
gL3·Pr

1 Modification to
ν2

0.003 nusselt
6 theory
) hLk ) /0.943

4
0.1

0.01

ΔT = (Tv – Tw)
For experimental data 5
h = hc
ΔT = ΔTf
For theory
h = hNu

0.001
0.0001 0.001 0.01 0.1 1
CP· ΔT/λ
62 2 Heat Transfer During Condensation

2.9.2 Interfacial Resistance all cases in which low condensation coefficient had been
found, it could be attributed to inadequate evacuation
The low heat transfer coefficients discussed earlier were
and/or experimental uncertainties. He pointed out that
attributed to interfacial resistance. For vapor to condense,
at pressure above 0.1 bar, (T SAT − T i ) is small compared
pressure of vapor has to be higher than that of liquid at the
with (T SAT − T w ) and therefore is difficult to measure with
interface. Thus, there is an interfacial resistance. The ques-
sufficient accuracy. Further, gases may be released from the
tion is about its magnitude. One theory is that all molecules
vessel walls at the higher temperatures. In the later study
striking the liquid interface are not condensed. The ratio
of Sakhuja (1970), traces of secondary gases were collected
of molecules condensed to those striking the interface is
in a secondary condenser and bled to the atmosphere.
called the condensing coefficient or the accommodation
When bleeding was stopped, condensation coefficients
coefficient. The theory of interfacial resistance was mainly
fell to the low values in other studies. When bleeding was
developed by Schrage (1953). Rohsenow (1973a) has given
resumed, the coefficients rose back to 1. He concluded that
the following simplified theoretical formula for interfacial
the condensing coefficient is really always = 1. Berman
heat transfer coefficient hi :
(1969) also gave the opinion that condensing coefficient
( ) 2
2𝜎 M 0.5 ifg is always 1, the lower values reported by some are due to
hi = (2.9.1) presence of non-condensable gases.
2 − 𝜎 2𝜋R 𝑣fg Tg
3∕2

This is applicable for


𝑤∕A 2.9.3 Moving Vapors
< 0.1 (2.9.2)
𝜌f (2TR∕M)1∕2 Experiments on condensation of potassium containing
where 𝜎 is the condensation coefficient, A is the area of sur- 1.5% sodium in a 4 mm diameter tube were done by Ala-
face, R is the universal gas constant, M is the molecular dyev et al. (1966). Vapor velocities were 40–290 m s−1 and
mass, and vfg is the specific volume change during conden- pressures ranged 0.05–1.3 bar. They estimate that the resis-
sation. This interfacial resistance is in series to the resis- tance of the condensate film was less than 9% of the total.
tance of liquid film. The actual heat transfer coefficient h is They correlated their data by the following formula:
thus hTP = 7.94p0.22 (2.9.4)
1 1 1
= + (2.9.3)
h hc hi where hTP is in W m−2 K−1 and p is in N m−2 . They com-
pared this equation with data for condensation of stagnant
where hc is the heat transfer coefficient in the absence of
vapor from several sources including Misra and Bonilla
interfacial resistance. Many methods to calculate 𝜎 have
(1956). All of them agreed with this equation. It would
been proposed. Dannon (1962) has given a method to
therefore appear that vapor velocity and geometry have
calculate the condensation coefficient. According to it,
no effect. This behavior is quite different from that of
the condensation coefficient varies from 0.014 for ethyl
non-metallic fluids as was discussed in Section 2.3. It is
alcohol to 1.01 for carbon tetrachloride and that for water
unlikely that the mechanism of heat transfer of liquid
is 0.051. These were reported to be in good agreement with
metals be quite different from that of non-metallic fluids.
test data.
Most probably, the apparent discrepancy is because the
From their data on condensation of mercury, Sukhatme
resistance of metallic condensate is very small compared
and Rohsenow (1966) determined the condensing coeffi-
with that of non-condensables, which prevents accurate
cient to be 0.45. In various studies on condensation, 𝜎 was
measurement of heat transfer coefficient.
found to be around 1 till a pressure of about 0.05 bar and
Kroger and Rohsenow (1967) gave a correlation for con-
dropped rapidly to as low as 0.02 (Rohsenow 1973a). How-
densation coefficients that was in good agreement with data
ever, Wilcox and Rohsenow (1970) and Sakhuja (1970) in
for stagnant metal vapors as well as the data of Aladyev
their studies obtained condensation coefficients close to 1
et al. for high velocity vapor discussed earlier. This indi-
at pressures up to 0.2 bar. Huang et al. (1972) performed an
cated that the data of Aladyev et al. were also affected by
analysis that assumes traces of non-condensables that pre-
non-condensables.
dicts condensing coefficient to be 0.88–0.92 at all pressures
up to the atmospheric.
Rohsenow (1973a, 1973b) carried out a careful review
2.9.4 Recommendation
of the earlier experimental studies done at M.I.T. as to the
extent to which the apparatus had been evacuated and For stagnant vapor, use Eq. (2.9.3) with hi by Eq. (2.9.1)
the uncertainty in measurements. He concluded that in using 𝜎 = 1, and hc by the Chen (1961a,1961b) analytical
2.10 Dropwise Condensation 63

solution represented by Eq. (2.2.16). When using 2.10.2 Theories of Dropwise Condensation
Eq. (2.2.16), calculate hNu by Nusselt equation appro-
Theories of dropwise condensation have been reviewed
priate for the condensing surface geometry. Correct for
among others by Rose (2002) and Huang et al. (2015).
effect of non-condensables as described in Section 2.7.8.
There are two basic theories of dropwise condensation.
For moving vapors, no verified general method is avail-
According to the first theory, a thin condensate layer is
able. Conservatively, heat transfer coefficient may be cal-
initially formed by the film condensation process. When
culated assuming stagnant vapor.
this film reaches a certain critical thickness, surface ten-
sion forces fracture it and rolls it up in the form of drops.
This theory was first put forward by Jakob (1936) and has
2.10 Dropwise Condensation
been further developed by others such as Baer and Mck-
All discussions in this chapter till now have been on elvey (1958), Welch and Westwater (1961), and Sugawara
film-type condensation in which condensate forms a con- and Katsuta (1966). Welch and Westwater estimated the
tinuous layer on the surface. In dropwise condensation, critical thickness of film to be 0.5–1 μm. Majumdar and
condensation occurs in the form of drops with space Mezic (1999) actually observed with a microscope that
between them. Heat transfer coefficients in dropwise there was a liquid film with a few nanometers thickness
condensation are an order of magnitude higher than in on the condenser surface. According to other theories,
film condensation. Hence there is, and has been for a condensation occurs on random nucleation sites such as
long time, a keen interest in developing heat exchangers pits and scratches. In their experiments, Umur and Griffith
with dropwise condensation. As yet durable and econom- (1965) observed that on the surface between drops, the
ical surfaces with dropwise condensation have not been thickness of liquid film was no more than a monomolecu-
developed. Ahlers et al. (2019) have reviewed current lar layer, and essentially no condensation occurs on it. This
technologies and discussed future possibilities. contradicts the critical film thickness theory. Other studies
supporting the random nucleation theory include those of
Liu et al. (2007) and Mu et al. (2008). As yet there is no
2.10.1 Prediction of Mode of Condensation
consensus on the mechanism of dropwise condensation.
It has been observed that dropwise condensation occurs
only when the liquid does not wet the surface. Thus pure
mercury does not wet metals and always gives dropwise
condensation. Steam produces drop condensation on 2.10.3 Methods to Get Dropwise Condensations
noble metals but on ordinary metals condenses as a film.
To achieve dropwise condensation of steam on ordinary Most metals have high surface energy and hence are
metals, promoters have to be applied on the surfaces that hydrophilic. For dropwise condensation, the surface
render them non-wetting. It is of theoretical and practical should be hydrophobic, that is, can repel water. Ordinary
interest to be able to predict the mode of condensation. It metals can be made hydrophobic by treating the surface
has been found that dropwise condensation occurs if the with substances known as promoters. The method that has
liquid–vapor surface tension is greater than the surface free been in use from the earliest is to continuously or intermit-
energy or surface tension of the solid surface, Merte (1973). tently inject a promoter into the condenser. The promoters
Davies and Ponter (1968) studied data for condensation that have been used in steam condensers include stearic
on Teflon surfaces and found that dropwise condensation acid, montan wax, benzyl mercaptan, oleic acid, dioctade-
occurred when liquid–vapor surface tension was between cyle disulfide, and cupric oleate. Another method is to
0.033 and 0.061 N m−1 and film condensation occurred coat the surface with a durable layer of a promoter such
when its value was between 0.019 and 0.02. The surface as Teflon. However, it has not been possible to achieve a
tension of Teflon is 0.018 N m−1 according to Shaffrin and durable layer.
Zisman (1960). In recent years, a variety of surfaces have been developed
A related topic is transition from dropwise to film con- in an effort to achieve long-lasting dropwise condensation.
densation. Rose (2002) reviewed the experimental data and This is done by changing the surface characteristics. The
theories. He concluded that this phenomenon is still not techniques used include lithography, plasma treatment,
clearly understood. The important mechanisms appear to chemical deposition, solution casting, electro-spraying, etc.
be the proximity of active site with increasing temperature These have been described and reviewed by Ahlers et al.
difference between vapor and surface and with the speed at (2019). However, none of these surfaces has achieved the
which drops coalesce as compared with their growth rate. goal of durability together with economic viability.
64 2 Heat Transfer During Condensation

2.10.4 Some Experimental Studies a method to calculate it. Their method combines an
expression for the heat transfer through a drop of given
Figure 2.10.1 shows data on dropwise condensation of
size with an expression for the average distribution of drop
steam at near atmospheric pressure from 1930 onwards put
sizes. These are integrated overall drop sizes to obtain, for
together by Rose (2002). These were obtained on surfaces
a given vapor–surface temperature difference, the heat flux
with promoters. It is seen that there are very large differ-
for the surface. Rose (2002) reports that this method was
ences between the data from different sources. According
found to be in good agreement with wide-ranging steam
to Rose, the data in the hatched portion are the correct ones.
data as well as data for several organics form independent
The data with large ΔT are erroneous due to experimental
sources. Calculations with this method are quite complex.
errors and/or presence of non-condensing gases. As was
Griffith (1973) has given the following correlation for
discussed in Section 2.7.8, non-condensable gases greatly
steam on vertical surfaces:
reduce heat transfer. Rose points out that careful experi-
25 < T SAT < 100 ∘ C,
mentation shows that heat transfer coefficient increases
with increasing temperature difference, the reason being hdc = 47 628 + 2041TSAT (2.10.2)
that more nucleation points get activated. The data in the
hatched portion of Figure 2.10.1 show this trend, while the T SAT > 100 ∘ C,
data lines with large slope show decreasing heat transfer hdc = 255 150 (2.10.3)
coefficients with increasing temperature difference.
Koch et al. (1998) studied the effect of orientation of The units are T SAT in C and h in W m−2 K−1 .
a flat condensing surface. Their results are shown in Rose (2002) gives the following equation that fits the
Figure 2.10.2. It is seen that heat transfer coefficients steam data considered by him to be reliable:
decrease as inclination changes from vertical to horizontal. 0.8
q = TSAT [5ΔTSAT + 0.3(ΔTSAT )2 ] (2.10.4)
Heat transfer coefficient in the horizontal position is about
40% of that in the vertical position. Graham (1969) studied This is applicable to pressures near atmospheric and
data from three sources and also found similar effect of lower. T is in C and q in kW m−2 .
inclination. This effect is because drainage of drops is hp is the promoter heat transfer coefficient. Heat is con-
retarded as inclination becomes less. Further, a drop sitting ducted through the promoter before it reaches the condens-
on a horizontal surface occupies more area than a drop ing surface. Thus,
hanging from a vertical surface and hence has higher k
thermal resistance. hp = (2.10.5)
𝛿
where 𝛿 is the thickness of promoter layer and k is its ther-
2.10.5 Prediction of Heat Transfer mal conductivity.
Usually this resistance is negligible as the promoter is
There are several resistances involved in the overall heat
only a monolayer. For some promoters such as Teflon,
transfer during dropwise condensation. Griffith (1973)
the thermal conductivity is low and the layer thickness is
have represented the overall heat transfer coefficient U as
significant. The minimum required thickness of Teflon is
follows:
around 1.5 μm (Griffith 1973), and its thermal conductivity
1 1 1 1 1 1 is 0.17 W m K−1 . This gives hp = 113 300 W m−2 K−1 .
= + + + + (2.10.1)
U hnc hi hdc hp hc
hc is the constriction heat transfer coefficient attributed
here hnc is the heat transfer coefficient due to presence of to thermal conductivity of condensing surface. Some
non-condensables. Effect of non-condensables on film con- researchers reported that thermal conductivity has great
densation was discussed in Section 2.7.8. This resistance effect on heat transfer. As the heat of condensation has
can be treated in the same way for dropwise condensation. to pass through the metal surface, heat transfer is “con-
The impact of non-condensable is usually more severe in stricted” if the thermal conductivity is low. Rose (2002)
dropwise condensation as the heat transfer coefficients of examined the data on the effect of thermal conductivity
pure vapor are much higher than in film condensation. of condensing surface. He concluded that the balance of
hi is the heat transfer coefficient due to interfacial resis- evidence suggests that this is significant only at very low
tance. It can be calculated in the way described in Section heat fluxes and for very small condensing surfaces. Most
2.9.1. The condensation coefficient is to be taken as 1. of the data showing large constriction effect are affected
hdc is the heat transfer coefficient for conduction through by non-condensables and experimental errors. Hence the
liquid drops. Its magnitude depends on the size and pop- constriction effect is not pertinent to most practical heat
ulation of drops. Le Fevre and Rose (1966) developed exchangers.
2.10 Dropwise Condensation 65

Figure 2.10.1 Heat transfer 25


measurements on dropwise
condensation of steam at near 6 1
atmospheric pressure. Source: From 10
Rose (2002). © 2002 SAGE Publications. 20
23

15
4 24

ΔT/(K)
5

10 8
2
16 12

9, 11, 13, 14, 15, 17, 18, 19, 20, 21, 22


5
3

0.5 1.0 1.5 2.0


q (MW m−2)

Figure 2.10.2 Effect of surface inclination on 1,2


heat transfer coefficient α during dropwise
condensation. Source: From Koch et al. (1998). α ( β)
© 1998 Elsevier. α ( β = 90°)

1
0.976
0.960

0.898

0.830
0.8
0.761
0.3 MW m–2
–2
0.4 MW m
0.5 MW m–2
Ratio

0.6 MW m–2
0.6
0.7 MW m–2
0.8 MW m–2
0.9 MW m–2
1.0 MW m–2
0.443
0.4

Face down
Face up

0.2

0
0 30 60 90 120 150 180
Surface inclination angle β(°)
66 2 Heat Transfer During Condensation

2.10.6 Recommendations k thermal conductivity (W m−1 K−1 )


L length (m)
For steam near atmospheric pressure and lower pressures,
MAD mean absolute deviation
use Eq. (2.10.4) to obtain overall heat transfer coefficient
N number of data points, or number of tube
in the absence of non-condensable gases. For other fluids
rows (−)
and conditions, use the correlation of Le Fevre and Rose for
Nu Nusselt number (−)
hdc . If it is a flat surface and it is not vertical, correct for the
pr reduced pressure (−)
effect of orientation using Figure 2.10.2. Correct for other
Pr Prandtl number, (−)
resistances as described in Section 2.10.5.
R universal gas constant, = 8.134 J m−1 K−1
Ra Raleigh number (−)
Nomenclature Re Reynolds number = GD μ−1 (−)
ReGT Reynolds number for all mass flowing as
A area (m2 ) vapor = GD μG −1 (−)
Bd Bond number = g(𝜌L − 𝜌G )D2 𝜎 −1 (−) ReGS Reynolds number assuming vapor phase
D diameter of tube (m) flowing alone = GxDμG −1 (−)
Deeq equivalent Dean number defined by ReLS Reynolds number assuming liquid phase
Eq. (2.7.37) (−) flowing alone, = G (1 − x)DμL −1 (−)
DHP equivalent diameter based on perimeter with ReLT Reynolds number for all mass flowing as
heat transfer (m) liquid = GDμL −1 (−)
DHYD hydraulic equivalent
( diameter
) (m) Su Sugomel number (−)
G2
Fr Froude number = 𝜌2 gD (−) T temperature (K)
( )
2 ΔT (T w − T SAT ) (K)
Fr LT Froude number = 𝜌G2 gD (−) ΔT SAT (T w − T SAT ) (K)
f

G total mass flux (liquid + vapor) (kg m−2 s−1 ) u velocity (m s−1 )
g acceleration due to gravity (m s−2 ) WeGT Weber number for all mass flowing as vapor,
h heat transfer coefficient (W m−2 K−1 ) given by Eq. (2.3.54) (−)
hI heat transfer coefficient in Shah correlation w mass flow rate of condensate (kg s−1 )
given by Eq. (2.3.16) (W m−2 K−1 ) x vapor quality (−)
hTP two-phase heat transfer coefficient X tt Martinelli parameter defined by Eq. (2.3.6)
(W m−2 K−1 ) (−)
hTP,𝜃 two-phase heat transfer coefficient at tube Z Shah’s correlating parameter defined by
inclination 𝜃 (W m−2 K−1 ) Eq. (2.3.13) (−)
Jg dimensionless vapor velocity defined by
Eq. (2.3.4) (−)

Greek Letters
𝜇 dynamic viscosity (Pa s) 𝜃 angle (∘ or rad)
𝜈 kinematic viscosity (m2 s−1 ) Γ condensate mass flow rate per unit width(
𝜌 density (kg m−3 ) kg m−1 s−1 )
𝜎 surface tension (N m−1 )

Subscripts
f liquid LO liquid phase flowing alone, same as LS
FC forced convection LS liquid phase flowing alone, superficial liquid
G vapor LT all mass flowing as liquid
GO vapor phase flowing alone, same as GS m mean
GS vapor phase flowing alone, superficial vapor mix mixture
GT all mass flowing as vapor Nu by Nusselt equation
g vapor SAT saturation
i interface w wall
L liquid z at location z
References 67

References
Abdukmanov, K.A. and Mirmov, N.I. (1971). Experimental flowing through dimpled tubes with different helical and
study of oil-contaminated ammonia vapor on horizontal dimpled pitches. Int. J. Heat Mass Transf. 121: 620–631.
tubes. Heat Transfer Sov. Res. 3 (6): 176–180. ASHRAE (2017). Handbook Fundamentals, 4.24. Atlanta, GA:
Ackerman, G. (1937). Simultaneous heat and molecular mass ASHRAE.
transfer by overall temperature and partial pressure Awad, M.M., Dalkilic, A.S., and Wongwises, S. (2014). A
differences. VDI Forschungsh. 382: 1–16. critical review on condensation heat transfer in
Afroz, H.M.M., Miyarah, A., and Tsubaki, K. (2008). Heat microchannels and minichannels. J. Nanotechnol. Sci. Med.
transfer coefficients and pressure drops during in-tube 5: 010801-1–010801-25.
condensation of CO2 /DME mixture refrigerant. Int. J. Azer, N.Z., Abis, L.V., and Soliman, H.M. (1972). Local heat
Refrig. 31: 1458–1466. transfer coefficients during annular flow condensation.
Agarwal, R. and Hrnjak, P. (2015). Condensation in two phase ASHRAE Trans. 78: 135–143.
and desuperheating zone for R1234ze(E), R134a and R32 Azzolin, M., Bortolina, S., Nguyen, L.P.L. et al. (2018).
in horizontal smooth tubes. Int. J. Refrig. 50: 172–183. Experimental investigation of in-tube condensation in
Ahlers, M., Buck-Emden, A., and Bart, H. (2019). Is dropwise microgravity. Int. Commun. Heat Mass Transfer 96: 69–79.
condensation feasible? A review on surface modifications Bae, S., Maulbetsch, J.L. and Rohsenow, W.M. (1968). Forced
for continuous dropwise condensation and a profitability Convection Condensation Inside Tubes. M.I.T. Report.
analysis. J. Adv. Res. https://doi.org/10.1016/j.jare.2018.11 DSR-79760-59.
.004. Baer, E. and Mckelvey, J.M. (1958). Heat transfer in dropwise
condensation. ACS Delaware Science Symposium. Quoted
Akers, W.W., Deans, H.A., and Crosser, O.K. (1959).
in Huang et al. (2015).
Condensing heat transfer within horizontal tubes. Chem.
Bakulin, N.V., Ivanovskii, M.N., Sorokin, V.P., and Subbotin,
Eng. Prog. Symp. Ser. 59 (29): 171–176.
V.I. (1967). Intensity of heat transfer in film condensation
Aladyeev, I.T., Kondratyev, N.S., Mukhin, V.A. et al. (1966).
of pure sodium vapour. Teplofiz. Vysok. Temper 5 (5): 930.
Thermal resistance of phase transition with condensation
Quoted in Necme and Rose (1976).
of potassium vapor. In: Proceedings of the Third
Beatty, K.O. and Katz, D.L. (1948). Condensation of vapors on
International Heat Transfer Conference, vol. 2, 312–317.
outside of finned tubes. Chem. Eng. Prog. 44: 55–70.
Chicago: Begell House.
Beatty, K.O. & Nandapurkar, S.S. (1959) Condensation on a
Alasadi, A.A.M., Ezzat, A.W., and Muneer, A. (2013).
horizontal rotating disc. Preprint 112, presented at the
Investigation of steam condensation process on rotating
Third National Heat Transfer Conference, ASME-AIChE,
disk condenser at different rotation speed. Int. J. Comput.
Storrs, Conn., August, 1959. Quoted in Rohsenow (1973a).
Appl. 84 (14): 10–18.
Belghazi, M., Bontemp, A., Signe, J.C., and Marvillet, C.
Alekseev, V.P., Poberezkin, A.E., and Gerasimov, P.V. (1972).
(2001). Condensation heat transfer of a pure fluid and
Determination of flooding regular packings. Heat Transfer binary mixture outside a bundle of smooth tubes,
Sov. Res. 4: 159–163. comparison of experimental results and a classical model.
Al-Hajeri, M., Koluib, A., Mosaad, M., and Al-Kulaib, S. Int. J. Refrig. 24: 841–855.
(2007). Heat transfer performance during condensation of Bell, K.J. and Ghaly, M.A. (1973). An approximate
R-134a inside helicoidal tubes. Energy Convers. Manag. 48: generalized method for multicomponent partial
2309–2315. condenser. AIChe. Symp. Ser. 69: 72–79.
Ali, H.M. (2017). Condensation heat transfer on Berman, L.D. (1969). Heat transfer during film condensation
geometrically enhanced horizontal tube: a review. In: Heat of vapor on horizontal tubes in transverse flow. In:
Exchangers - Advanced Features and Applications (ed. Convective Heat Transfer in Two-Phase and One Phase Flow
S.M.S. Murshed). InTech https://doi.org/10.5772/65896. (eds. V.M. Borishnaskii and I.I. Paleev). Jerusalem: I. P. S. T
Altman, M., Staub, F.W., and Norris, R.H. (1959). Local Heat Press.
Transfer and Pressure Drop for Refrigerant 22 Condensing in Berman, L.D. and Tumanov, Y.A. (1962). Condensation heat
Horizontal Tubes. Storrs, CT: ASME-AIChE. transfer in a vapour flow over a horizontal tube.
Ananiev, E.P., Boyko, I.D., and Kruzhilin, G.I. (1961). Heat Teploenergerika 10: 77–84. Quoted in Berman (1969).
transfer in the presence of steam condensation in Berrichon, J.D., Louahlia-Gualous, H., Bandelier, P. et al.
horizontal tubes. Int. Dev. Heat Transfer 2: 290–295. (2015). Experimental study of flooding phenomenon in a
Aroonrat, K. and Wongwises, S. (2018). Condensation heat power plant reflux air-cooled condenser. Appl. Therm. Eng.
transfer and pressure drop characteristics of R-134a 79: 214–224.
68 2 Heat Transfer During Condensation

Bohdal, T., Charun, H., and Sikora, M. (2011). Comparative microfin tubes: a new computational procedure. Int. J.
investigations of the condensation of R134a and R404A Refrig. 32: 162–174.
refrigerants in pipe minichannels. Int. J. Heat Mass Transf. Chamra, L.M. and Mago, P.J. (2006). Modeling of
54 (9–10): 1963–1974. condensation heat transfer of refrigerant mixture in
Briggs, A. and Rose, J.W. (1994). Effect of fin efficiency on a micro-fin tubes. Int. J. Heat Mass Transf. 29: 1915–1921.
model for condensation heat transfer on a horizontal, Chamra, L.M., Mago, P.J., Tan, M., and Kung, C. (2005).
integral-fin tube. Int. J. Heat Mass Transf. 37 (1): 457–463. Modeling of condensation heat transfer of pure refrigerant
Bromley, L.A. (1952). Effect of heat capacity of condensate. in micro-fin tubes. Int. J. Heat Mass Transf. 48: 1293–1302.
Ind. Eng. Chem. 44: 2966–2968. Chandran, R. and Watson, F.A. (1976). Condensation on
Browne, M.W. and Bansal, P.K. (1999). An overview of static and rotating pinned tubes. Trans. Inst. Chem. Eng. 54
condensation heat transfer on horizontal tube bundles. (2): 65–72.
Appl. Therm. Eng. 19: 565–594. Chato, J.C., (1960). Laminar condensation inside horizontal
Butterworth, D. (1977). Developments in the Design of Shell and inclined tubes. PhD thesis at the Massachusetts
and Tube Condensers. ASME Paper 77-WA/HT-24. Institute of Technology, Cambridge, MA.
Carnavos, T.C. (1980). Heat transfer performance of Chen, M.M. (1961a). An analytical study of laminar film
internally finned tubes in turbulent flow. Heat Transfer condensation part 1- flat plates. J. Heat Transf. 83 (1):
Eng. 1 (4): 32–37. 48–54.
Carpenter, F.G. and Colburn, A.P. (1951). The effect of vapor Chen, M.M. (1961b). An analytical study of laminar film
velocity in condensation inside tubes. In: General
condensation part 1-single and multiple horizontal tubes.
Discussion on Heat Transfer, 20–26. Institution of
J. Heat Transf. 83 (1): 55–60.
Mechanical Engineers and ASME.
Chen, S.L., Gerner, F.M., and Tien, C.L. (1987). General film
Caruso, G., Vitale Di Maio, D., and Naviglio, A. (2013).
condensation correlations. Exp. Heat Transfer 1 (2): 93–107.
Condensation heat transfer coefficient with
Cho, Y., Kim, S., Bae, B. et al. (2013). Assessment of
noncondensable gases inside near horizontal tubes.
condensation heat transfer model to evaluate performance
Desalination 309: 247–253.
of the passive auxiliary feedwater system. Nucl. Eng.
Cavallini, A. and Zecchin, R. (1966). High velocity
Technol. 45 (6): 759–766.
condensation of R-11 vapor inside vertical tubes.
Cohn, P. D. (1960). MS thesis. Oregon State College. Quoted
Proceedings of the International Institute of Refrigeration
in Sukhatme and Rohsenow (1964).
1966, Commission 2, Trondheim, Norway.
Colburn, A.P. and Drew, T.B. (1937). The condensation of
Cavallini, A. and Zecchin, R. (1974). A dimensionless
mixed vapors. Trans. Am. Inst. Chem. Eng. 33: 197–215.
correlation for heat transfer in forced convection
Colburn, A.P. and Hougen, O.A. (1934). Design of cooler
condensation. In: Proceedings of the Fifth International
condensers for mixture of vapors with noncondensing
Heat Transfer Conference, vol. 3, 309–313. Tokyo: Begell
gases. Ind. Eng. Chem. 26: 1178–1182.
House.
Cavallini, A., Doretti, L., Klammsteiner, N. et al. (1995). Collier, J.G. and Thome, J.R. (1994). Convective Boiling and
Condensation of new refrigerants inside smooth and Condensation, 3e. Oxford, UK: Oxford University Press.
enhanced tubes. In: Proceedings o0f the 19th International Da Riva, E., Del Col, D., Garimella, S.V., and Cavallini, A.
Congress of Refrigeration, vol. 4, 105–114. The Hague. (2012). The importance of turbulence during condensation
Cavallini, A., Del Col, D., Doretti, L. et al. (1999). A new in a horizontal circular minichannel. Int. J. Heat Mass
computational procedure for heat transfer and pressure Transf. 55 (13–14): 3470–3481.
drop during refrigerant condensation inside enhanced Dalkilic, A.S. and Wongwises, S. (2009). Intensive literature
tubes. Enhanced Heat Transfer 6: 441–456. review of condensation inside smooth and enhanced tubes.
Cavallini, A., Censia, G., Del Col, D. et al. (2003). Int. J. Heat Mass Transf. 52: 3409–3426.
Condensation inside and outside smooth and enhanced Dannon, F. (1962). Topics Statistical Mechanics. USAEC
tubes — a review of recent research. Int. J. Refrig. 26: Report UCRL 10029.
373–392. Davies, G.A. and Ponter, A.B. (1968). The prediction of the
Cavallini, A., Del Col, D., Doretti, L. et al. (2006). mechanism of on tubes coated with tetrafluoroethylene.
Condensation in horizontal smooth tubes: a new heat Int. J. Heat Mass Transf. 11: 375–377.
transfer model for heat exchanger design. Heat Transfer Del Col, D., Cavallini, A., and Thome, J.R. (2005).
Eng. 27 (8): 31–38. Condensation of zeotropic mixtures in horizontal tubes:
Cavallini, A., Del Col, D., Mancin, S., and Rossetto, L. (2009). new simplified heat transfer model based on flow regimes.
Condensation of pure and near-azeotropic refrigerants in J. Heat Transf. 127: 221–230.
References 69

Del Col, D., Bortolato, M., Azzolin, M., and Bortolin, S. Gacesa, M. (1967). Condensation of steam on a vertically
(2014). Effect of inclination during condensation inside a rotating cylinder. Electronic Thesis and Dissertations.
square cross section minichannel. Int. J. Heat Mass Transf. https://scholar.uwindsor.ca/etd/6476.
78: 760–777. Garimella, S., Killion, J.D., and Coleman, J.W. (2001). An
Del Col, D., Bortolin, S., and Das Riva, E. (2015). Predicting experimentally validated model for two-phase pressure
methods and numerical modeling of condensation in drop in the intermittent flow regime for circular
microchannels. In: Encyclopedia of Two-Phase Heat microchannels. In: Proceedings of NHTC’01,
Transfer and Fluid Flow II – Special Topics and NHTC2001-20103. ASME.
Applications, Special Topics in Condensation, vol. 3 (ed. J.R. Garimella, S., Fronk, B.M., Milkie, J.A., and Keinath, B.L.
Thome), 37–84. New York: World Scientific Publication. (2014). Versatile models for condensation of fluids with
Dobran, F. and Thorsen, R.S. (1980). Forced flow laminar widely varying properties from the micro to macroscale.
filmwise condensation of a pure saturated vapor in a Proceedings of the 15th International Heat Transfer
vertical tube. Int. J. Heat Mass Transf. 23: 161–171. Conference. IHTC-15, IHTC15-10516, Kyoto, Japan (10–15
Dobson, M.K. and Chato, J.C. (1998). Condensation in August 2014).
smooth horizontal tubes. J. Heat Transf. 120: 193–213. Ge, M., Wang, S., Zhao, J. et al. (2016). Condensation of steam
Dorao, C.A. and Fernandino, M. (2018). Simple and general with high CO2 concentration on a vertical plate. Exp.
correlation for heat transfer during flow condensation Thermal Fluid Sci. 75: 147–155.
inside plain pipes. Int. J. Heat Mass Transf. 122: 290–305. Gerstmann, J. and Griffith, P. (1967). Laminar film
Dukler, A.E. (1960). Fluid mechanics and heat transfer in condensation on the underside of horizontal and inclined
surfaces. Int. J. Heat Mass Transf. 10: 567–580.
vertical falling film systems. Chem. Eng. Symp. Ser. 56 (30):
Ghiaasiaan, S.M. (2008). Two Phase Flow, Boiling and
1–10.
Condensation in Conventional and Miniature Systems.
Eckels, S. J. (2005). Effects of Inundation and Miscible Oil
Cambridge: Cambridge University Press.
Upon Condensation Heat Transfer Performance of R-134a.
Graham, C. (1969). The limiting mechanisms of dropwise
ASHRAE report 984.
condensation. Doctoral dissertation. Masschusetts Institute
Eisenberg, D.M. (1972). An investigation of the variables
of Technology. Cambridge, MA.
affecting steam condensation on the outside of a horizontal
Griffith, P. (1973). Dropwise condensation. In: Handbook of
tube bundle. PhD thesis. University of Tennessee,
Heat Transfer (eds. W.M. Rohsenow and J.P. Hartnett),
Knoxville. Quoted in Belghazi et al. (2001).
12-34–12-47. New York: McGraw-Hill.
El Hajal, J., Thome, J.R., and Cavallini, A. (2003).
Gupta, A., Kumar, R., and Gupta, A. (2014). Condensation of
Condensation in horizontal tubes. Part 1: two-phase flow
R-134a inside a helically coiled tube-in-shell heat
pattern map. Int. J. Heat Mass Transf. 46: 3349–3363.
exchanger. Exp. Thermal Fluid Sci. 54: 279–289.
Eldeeb, R., Aute, V., and Radermacher, R. (2016). A survey of
Han, D. and Lee, K.J. (2005). Experimental study on
correlations for heat transfer and pressure drop for
condensation heat transfer enhancement and pressure
evaporation and condensation in plate heat exchangers. drop penalty factors in four microfin tubes. Int. J. Heat
Int. J. Refrig. 65: 12–26. Mass Transf. 48: 3804–3816.
Fronk, B.M. and Garimella, S. (2016). Condensation of Han, D., Lee, K., and Kim, Y. (2003). The characteristics of
ammonia and high-temperature-glide ammonia/water condensation in brazed plate heat exchangers with
zeotropic mixtures in minichannels – part I: different chevron angles. J. Korean Phys. Soc. 43 (1): 66–73.
measurements. Int. J. Heat Mass Transf. 101: 1343–1356. Hashimoto, H. and Kaminaga, F. (2002). Heat transfer
Fujii, T. and Oda, K. (1986). Correlation equations of heat characteristics in a condenser of closed two-phase
transfer for condensate inundation on horizontal tube thermosyphon: effect of entrainment on heat transfer
bundles. Trans. Jpn. Soc. Mech. Eng. B 52 (474): 822–826, in deterioration. Heat Transfer Asian Res. 31: 212–225.
Japanese. Quoted by Zeinelabedien et al. (2018). Hayakawa, T., Kawasaki, J., and Muraki, M. (1975).
Fujii, T., Uehera, H., Hirata, K., and Oda, K. (1972). Heat Condensation of single vapor on a vertical surface. Heat
transfer and flow resistance in condensation of low Transfer Jpn. Res. 4 (2): 55–62.
pressure steam flowing through tube banks. Int. J. Heat Honda, H. and Fujii, T. (1974). Effect of the direction of an
Mass Transf. 15: 247–260. oncoming vapor on laminar filmwise condensation on a
Fuks, S.N. and Zernova, E.P. (1970). Heat and mass transfer horizontal tube. In: Proceedings of the Fifth International
with condensation pure steam and steam containing air Heat Transfer Conference, vol. 3, 299–303. Tokyo: Begell
supplied from the side of a bank. Therm. Eng. 17 (3): 84–90. House.
70 2 Heat Transfer During Condensation

Honda, H. and Nozu, B. (1987). A prediction method for heat Jung, D., Song, K., Cho, Y., and Kim, S. (2003). Flow
transfer during condensation on horizontal low integral fin condensation of heat transfer coefficients of pure
tubes. J. Heat Transf. 109: 218–225. refrigerants. Int. J. Refrig. 26: 4–11.
Honda, H., Nozu, S., and Mitsumori, K. (1983). Augmentation Kedzierski, M.A. and Goncalves, J.M. (1997). Horizontal
of condensation on finned tubes by attaching a porous convective condensation of alternative refrigerants within
drainage plate. In: Proceedings of the ASME-JSME Thermal a micro-fin tube. NISTIR 6095, US Department of
Engineering Joint Conference, vol. 3, 289–295. ASME. Commerce.
Honda, H., Nozu, S., Bunken, U., and Fujii, T. (1986). Effect of Keinath, B.L. and Garimella, S. (2018). High-pressure
vapour velocity on film condensation of R-113 on condensing refrigerant flows through microchannels, part
horizontal tubes in a crossflow. Int. J. Heat Mass Transf. 29 II: heat transfer models. Heat Transfer Eng. https://doi.org/
(3): 429–438. 10.1080/01457632.2018.1443258.
Honda, H., Nozu, S., and Takeda, Y. (1987). Flow Kern, D.Q. (1958). Mathematical development of loading in
characteristics of condensate on a vertical column of horizontal condensers. AIChE J 4: 157–160.
horizontal low finned tubes. In: ASME-JSME Thermal Kharangate, C.R. and Mudawar, I. (2017). Review of
Engineering Joint Conference, 517–524. ASME. computational studies on boiling and condensation. Int. J.
Honda, H., Uchima, B., Nozu, S. et al. (1988). Condensation Heat Mass Transf. 108: 1164–1196.
of downward flowing R-113 vapor on bundles of horizontal Kim, S.M. and Mudawar, I. (2012). Theoretical model for
smooth tubes. Trans. Jpn. Soc. Mech. Eng.: 1453–1460. Also annular flow condensation in rectangular micro-channels.
in Heat Transfer Japanese Research, 18(6) 1989. Quoted in Int. J. Heat Mass Transf. 55: 958–970.
Collier and Thome (1984). Kim, S. and Mudawar, I. (2013). Universal approach to
Howarth, J.A., Poots, G., and Wynne, D. (1978). Laminar film predicting heat transfer coefficient for condensing
condensation on the underside of an inclined flat plate. mini/micro-channel flow. Int. J. Heat Mass Transf. 56
Mech. Res. Commun. 5 (6): 369–374. (1–2): 238–250.
Howe, M. (1972). Heat transfer by steam condensing onto Kirkbride, C.G. (1934). Heat transfer by condensing vapor on
rotating cones. Doctoral Thesis. Durham University. http:// vertical tubes. Int. J. Ind. Eng. Chem. 26 (4): 425–428.
etheses.dur.ac.uk/8824. Knudsen, J.G. and Katz, D.L. (1958). Fluid Dynamics and
Huang, Y.S., Lyman, F.A., and Lick, W.J. (1972). Heat transfer Heat Transfer. New York: McGraw-Hill.
by condensation of low pressure metal vapors. Int. J. Heat Koch, C., Kraft, K., and Leipertz, A. (1998). Parameter study
Mass Transf. 15: 741–754. on the performance of dropwise condensation. Rev. Gen.
Huang, X., Ding, G., Hu, H. et al. (2010). Influence of oil on Therm. 37: 539–548.
flow condensation heat transfer of R410A inside 4.18 and Koh, J.C.Y., Sparrow, E.M., and Hartnett, J.M. (1961). The
1.6 mm inner diameter horizontal smooth tubes. Int. J. two-phase boundary layer in laminar film condensation.
Refrig. 33: 158–169. Int. J. Heat Mass Transf. 2: 69–82.
Huang, J., Zhang, J., and Wang, L. (2015). Review of vapor Kondo, C. and Hrnjak, P. (2011). Heat rejection from R744
condensation heat and mass transfer in the presence of flow under uniform temperature cooling in a horizontal
non-condensable gas. Appl. Therm. Eng. 89: 469–484. smooth tube around the critical point. Int. J. Refrig. 34 (3):
Isachenko, V.P., Osipova, V.A., and Sukomel, A.S. (1977). 719–731.
Heat Transfer. Moscow: Mir Publishers. Kondo, C. and Hrnjak, P. (2012a). Heat rejection in
Ivanovskii, M.N., Milovanov, Y.V., and Subbotin, V.I. (1968). condensers close to critical point-Desuperheating,
Condensation coefficient of mercury vapour. Atomn. condensation in superheated region and condensation of
Energ. 24 (2): 146. Quoted in Necsa and Rose (1976). two phase fluid. International Refrigeration and Air
Jakob, M. (1936). Heat transfer in evaporation and Conditioning Conference, Purdue (16–19 July 2012).
condensation-II. Mech. Eng. 58: 729–739. Kondo, C. and Hrnjak, P. (2012b). Condensation from
Jakob, M. (1949). Heat Transfer, vol. 1. New York: Wiley. superheated vapor flow of R744 and R410A at sub-critical
Jige, D., Inoue, N., and Koyama, S. (2016). Condensation of pressures in a horizontal smooth tube. Int. J. Heat Mass
refrigerants in a multiport tube with rectangular Transf. 55: 2779–2791.
minichannels. Int. J. Refrig. 67: 202–213. Kondo, C. and Hrnjak, P. (2012c). Effect of microfins on heat
Jouhara, H. and Robinson, A.J. (2010). Experimental rejection in desuperheating, condensation in superheated
investigation of small diameter two-phase closed region and two phase zone. International Refrigeration and
thermosyphons charged with water, FC-84, FC-77 and Air Conditioning Conference, West Lafayette (Indiana),
FC-3283. Appl. Therm. Eng. 30: 201–211. 16–19 July, 2012. http://docs.lib.purdue.edu/iracc/1345.
References 71

Kosky, P.G. and Staub, F.W. (1971). Local condensation heat Majumdar, A. and Mezic, I. (1999). Instability of ultra-thin
transfer coefficients in the annular flow regime. AIChE J water films and the mechanism of droplet formation on
17 (5): 1037–1043. hydrophilic surfaces. J. Heat Transf. 121 (1999): 964–971.
Kroger, D.G. and Rohsenow, W.M. (1967). Film condensation Mancin, S., Del Col, D., and Rossetto, L. (2012). Condensation
of saturated potassium vapor. Int. J. Heat Mass Transf. 10: of superheated vapour of R410A and R407C inside plate
1891–1894. heat exchangers: experimental results and simulation
Kuhn, S.Z., Schrock, V.E., and Peterson, P.F. (1997). An procedure. Int. J. Refrig. 35 (7): 2003–2013.
investigation of condensation from steam-gas mixtures Mancin, S., Del Col, D., and Rossetto, L. (2013). R32 partial
flowing downward inside a vertical tube. Nucl. Eng. Des. condensation inside a brazed plate heat exchanger. Int. J.
177: 53–69. Refrig. 36 (2): 601–611.
Kumar, R., Varma, H.K., Mohanty, B., and Agrawal, K.N. Marto, P.J. (1988). An evaluation of film condensation on
(2002). Prediction of heat transfer coefficient during horizontal integral fin tubes. J. Heat Transf. 110:
condensation of water and R-134a on single horizontal 1287–1305.
integral-fin tubes. Int. J. Refrig. (25): 111–126. Mathews, D. (1962). The transfer of heat to cylinders of
Kuo, W., Lie, Y., Hsieh, Y., and Lin, T. (2005). Condensation various diameters rotating in a steam atmosphere, with
heat transfer and pressure drop of refrigerant R-410A flow varying conditions of temperature, pressure, and rotational
in a vertical plate heat exchanger. Int. J. Heat Mass Transf. speed. PhD Thesis, Imperial College, London, UK.
48: 5205–5220. Mazukewitch, L.V. (1952). Condensation of ammonia on
Le Fevre, E.J. and Rose, J.W. (1966). A theory of heat transfer vertical surface. Cholodilnaja Technika 29 (2): 50–51.
by dropwise condensation. In: Proceedings of 3rd Quoted here from the abstract in Kalteteknikk, 12, 334-335,
International Heat Transfer Conference, vol. 2, 362–375. 1952.
Chicago. McAdams, W.H. (1954). Heat Transmission, 3e. New York:
Lee, K.-Y. and Kim, M.H. (2008). Experimental and empirical McGraw-Hill.
study of steam condensation heat transfer with a McNaught, J.M. (1979). Mass-transfer correction term in
noncondensable gas in a small-diameter vertical tube. design methods for multicomponent/partial condensers.
Nucl. Eng. Des. 238: 207–216. In: Condensation Heat Transfer (eds. J.P. Marto and P.G.
Lee, H., Mudawar, I., and Hasan, M.M. (2013). Experimental Kroeger), 111–118. New York: ASME.
and theoretical investigation of annular flow condensation McNaught, J.M. (1982). Two-phase forced-convection heat
in microgravity. Int. J. Heat Mass Transf. 61: 293–309. transfer during condensation on horizontal tube banks. In:
Lips, S. and Meyer, J.P. (2011). Two-phase flow in inclined Proceedings of 7th Int Heat Transfer Conference, 6–10
tubes with specific reference to condensation: a review. Int. September, 1982, Munich, Germany, vol. 5, 125–131.
J. Multiphase Flow 37: 845–859. McQuillan, K.W. and Whalley, P.B. (1985). A comparison
Lips, S. and Meyer, J.P. (2012). Stratified flow model for between flooding correlations and experimental flooding
convective condensation in an inclined tube. Int. J. Heat data for gas-liquid flow in vertical circular tubes. Chem.
Fluid Flow 36: 83–91. Eng. Sci. 40 (8): 1425–1440.
Liu, T.Q., Mu, C.F., Sun, X.Y., and Xia, S.B. (2007). Meisenburg, J., Boarts, R.M., and Badger, W.L. (1935). The
Mechanism study on formation of initial condensate influence of small concentration of air in steam on the
droplets. AIChE J 53 (4): 1050–1055. steam film coefficient of heat transfer. Trans. Am. Inst.
Longo, G.A. (2009). R410A condensation inside a commercial Chem. Eng. 31: 622–637. Quoted in Ge et al. (2016).
brazed plate heat exchanger. Exp. Thermal Fluid Sci. 33 (2): Merte, H. (1973). Condensation heat transfer. In: Advances in
284–291. Heat Transfer (eds. T.F. Irvine and J.P. Hartnett), 180–272.
Longo, G.A. (2011). The effect of vapour super-heating on Cambridge, MA: Academic Press.
hydrocarbon refrigerant condensation inside a brazed plate Meyer, J.P., Dirker, J., and Adelaja, O.K. (2014). Condensation
heat exchanger. Exp. Thermal Fluid Sci. 35 (6): 978–985. heat transfer in smooth inclined tubes for R134a at
Longo, G.A., Righetti, G., and Zilio, C. (2015). A new different saturation temperatures. Int. J. Heat Mass Transf.
computational procedure for refrigerant condensation 70: 515–525.
inside herringbone type brazed plate heat exchangers. Int. Michael, A.G., Rose, J.W., and Daniels, L.C. (1989). Forced
J. Heat Mass Transf. 82: 530–536. convection condensation of steam on a horizontal
Lyulin, Y., Marchuk, I., Chikov, S., and Kabov, O. (2011). tube – experience with vertical downflow of steam. J. Heat
Experimental study of laminar convective condensation of Transf. 111: 792–797.
pure vapour inside an inclined circular tube. Microgravity Minkowycz, W.J. and Sparrow, E.M. (1966). Condensation
Sci. Technol.: 439–445. heat transfer in the presence of non-condensables,
72 2 Heat Transfer During Condensation

interface resistance, superheating, variable properties, and during condensation inside smooth, helical micro-fin, and
diffusion. Int. J. Heat Mass Transf. 9: 1125–1144. herringbone tubes. Int. J. Refrig. 30: 609–623.
Minkowycz, W.J. and Sparrow, E.M. (1969). The effect of Othmer, D.F. (1929). The condensation of steam. Indian
superheating on condensation heat transfer in a forced Chem. Eng. 21: 57–583.
convection boundary layer flow. Int. J. Heat Mass Transf. Owen, R.G., Sardesai, R.G., Smith, R.A., and Lee, W.C. (1983).
12: 147–154. Gravity controlled condensation on low integral-fin tubes.
Mirmov, N.I. and Yemelyanov, Y.V. (1976). Coefficient of heat AIChe. Symp. Ser. 75: 415–428.
transfer for ammonia condensers. Heat Transfer Sov. Res. 8 Palen, J. and Yang, Z.H. (2001). Reflux condensation flooding
(1): 50–51. prediction: review of current status. Trans. IChemE 79 (A):
Miropolskiy, Z.L., Shneerova, R.I., and Ternakova, L.M. 464–469.
(1974). Heat transfer at superheated steam condensation Park, J.E., Vakili-Farahani, F., Consolini, L., and Thome, J.R.
inside tube. In: Proceedings of theFifth International. Heat (2011). Experimental study on condensation heat transfer
Transfer Conference, Tokyo, vol. 3, 246–249. in vertical minichannels for new refrigerant R1234ze(E)
Misra, B. and Bonilla, C.F. (1956). Heat transfer in versus R134a and R236fa. Exp. Thermal Fluid Sci. 35:
condensation of metal vapors: mercury and sodium up to 442–454.
atmospheric pressure. Chem. Eng. Prog. Symp. Ser. 52 (18): Rahman, M.M., Kariya, K., and Miyara, A. (2018). An
7–21. experimental study and development of new correlation
Miyara, A., Nonaka, K., and Taniguchi, M. (2000). for condensation heat transfer coefficient of refrigerant
Condensation heat transfer and flow pattern inside a inside a multiport minichannel with and without fins. Int.
herringbone-type microfin tube. Int. J. Refrig. 23: 141–152. J. Heat Mass Transf. 116: 50–60.
Mizoshina, T., Ito, R., Yamashita, S., and Kamimua, H. Razavi, M.D. and Damle, A.S. (1978). Heat transfer
(1977). Film condensation of superheated vapor in a coefficients for turbulent filmwise condensation. Trans.
vertical tube. Heat Transfer Jpn. Res. 6 (4): 92–101. Inst. Chem. Eng. 56 (2): 81–85.
Mohseni, S.G., Akhavan-Behabadi, M.A., and Saeedinia, M. Rifert, V.G., Barabash, P.A., Solomakha, A.S. et al. (2018).
(2013). Flow pattern visualization and heat transfer Hydrodynamics and heat transfer in a centrifugal film
characteristics of R-134a during condensation inside a evaporator. Bulg. Chem. Commun. 50 ((Special issue K)):
smooth tube with different tube inclinations. Int. J. Heat 49–57.
Mass Transf. 60: 598–602. Rohsenow, W.M. (1956). Heat transfer and temperature
Moser, K.W., Webb, R.L., and Na, B. (1998). A new equivalent distribution in laminar film condensation. Trans. ASME
Reynolds number model for condensation in smooth tubes. 78: 1645–1648.
J. Heat Transf. 120: 410–416. Rohsenow, W.M. (1973a). Condensation. In: Handbook of
Mozafari, M., Akhavan-Behabadi, M.A., Qobadi-Arfaee, H., Heat Transfer (eds. W.M. Rohsenow and J.P. Hartnett),
and Fakoor-Pakdaman, M. (2015). Condensation and 12-1–12-33. New York: McGraw-Hill.
pressure drop characteristics of R600a in a helical Rohsenow, W.M. (1973b). Film condensation in liquid metals.
tube-in-tube heat exchanger at different inclination angles. In: Progress in Heat and Mass Transfer, vol. 7 (ed. O.E.
Appl. Therm. Eng. 90: 571–578. Dwyer), 469–484. London: Pergamon Press.
Mu, C., Pang, J., Lu, Q., and Liu, T. (2008). Effects of surface Rohsenow, W.M., Webber, J.H., and Ling, A.T. (1956). Effect
topography of material on nucleation site density of of vapor velocity on laminar and turbulent film
dropwise condensation. Chem. Eng. Sci. 63 (4): 874–880. condensation. Trans. ASME: 1637–1643.
Mudawar, I. (2017). Chapter five - flow boiling and flow Rose, J.W. (2002). Dropwise condensation theory and
condensation in reduced gravity. Adv. Heat Tran. 49 (2017): experiment: a review. P. I. Mech. Eng. A-J. Pow. 216:
225–306. 115–128.
Murata, K. and Hashizume, K. (1992). Prediction of Roth, J.A. (1962). Wright-Patterson AFB. Report
condensation heat transfer coefficients in horizontal ASD-TDR-62-738. Quoted in Sukhatme and Rohsenow
integral-fin tube bundles. Exp. Heat Transfer 5: 115–130. (1964).
Nicol, A.A. and Gacesa, M. (1970). Condensation of steam on Rudy, T.M. and Webb, R.L. (1985). An analytical model to
a rotating vertical cylinder. J. Heat Transf. 92 (1): 144–151. predict condensate retention on horizontal integral-fin
Nusselt, W. (1916). Die Oberflachenkondensation des tubes. J. Heat Transf. 107 (2): 361–368.
Wasserdampfes. Zeitscrift Ver. Deutsche Ing. 60 (541): 569. Rufer, C.E. and Kezios, S.P. (1966). Analysis of two-phase one
Olivier, J.A., Liebenberg, L., and Thome, J.R. (2007). Heat component stratified flow with condensation. J. Heat
transfer, pressure drop, and flow pattern recognition Transf. 88 (3): 265–275.
References 73

Saffari, H. and Naziri, V. (2010). Theoretical modeling and Shah, M.M. (2016a). Prediction of heat transfer during
numerical solution of stratified condensation in inclined condensation in inclined tubes. Appl. Therm. Eng. 94:
tubes. J. Mech. Sci. Technol. 24: 2587–2596. 82–89.
Sajjan, S.K., Kumar, R., and Gupta, A. (2015). Experimental Shah, M.M. (2016b). A new correlation for heat transfer
investigation during condensation of R-600a vapor over during condensation in horizontal mini/micro channels.
single horizontal integral-fin tubes. Int. J. Heat Mass Int. J. Refrig. 64: 187–202.
Transf. 88: 247–255. Shah, M.M. (2016c). Comprehensive correlations for heat
Sakhuja, R.K. (1970). Effect of superheat on film transfer during condensation in conventional and
condensation of potassium. Sc.D. thesis. Mechanical mini/micro channels in all orientations. Int. J. Refrig. 67:
Engineering Department of MIT Cambridge, MA. Quoted 22–41.
in Rohsenow (1973a). Shah, M.M. (2018a). Improved general correlation for heat
Salimpour, R.S., Shahmoradi, A., and Khoeini, D. (2017). transfer during gas-liquid flow in horizontal tubes. J.
Experimental study of condensation heat transfer of Therm. Sci. Eng. Appl. 10: 051009-1–051009-7.
R-404A in helically coiled tubes. Int. J. Refrig. 74: 584–591. Shah, M.M. (2018b). General correlation for heat transfer to
Sardesai, R.G., Owen, R.G., Smith, R.A., and Lee, W.C. (1983). gas-liquid flow in vertical channels. J. Therm. Sci. Eng.
Gravity controlled condensation on a horizontal low-fin Appl. 10: 061006-1–061006-9.
tube. IChemE Symp. Ser. 75: 415–428. Shah, M.M. (2019a). Improved correlation for heat transfer
Sarraf, K., Launay, S., and Tadrist, L. (2016). Analysis of during condensation in conventional and mini/micro
channels. Int. J. Refrig. 98: 222–237.
enhanced vapor desuperheating during condensation
Shah, M.M. (2019b). Prediction of heat transfer during
inside a plate heat exchanger. Int. J. Therm. Sci. 105:
condensation in non-circular channels. Inventions 2019
96–108.
(4): 31. https://doi.org/10.3390/inventions4020031.
Schlager, L.M., Pate, M.B., and Bergles, A.E. (1990).
Shah, M.M., Mahmoud, A.A., and Lee, J. (2013). An
Performance predictions of refrigerant-oil mixtures in
assessment of some predictive methods for in-tube
smooth and internally finned tubes – Part II: design
condensation heat transfer of refrigerant mixtures, Paper #
equations. ASHRAE Trans. 96 (1): 170–182.
DE-13-004 presented at ASHRAE meeting in Denver, CO,
Schrage, R.W. (1953). A Theoretical Study of Interphase Mass
June 2013, ASHRAE Transactions, 119 (2): 38–51.
Transfer. New York: Columbia University Press.
Shao, D.W. and Granryd, E. (1995). Heat transfer and pressure
Shaffrin, E.G. and Zisman, W.A. (1960). J. Phys. Chem. 64:
drop of HFC134a-oil mixtures in a horizontal condensing
519. Quoted in Merte (1973).
tube. Int. J. Refrig. 18 (8): 524–533.
Shah, M.M. (1975). Visual observations in an ammonia
Shekriladze, I.G. and Gomelauri, V.I. (1966). Theoretical
evaporator. ASHRAE Trans. 81 (1).
study of laminar film condensation of flowing vapour. Int.
Shah, M.M. (1976). A new correlation for heat transfer during
J. Heat Mass Transf. 9: 581–591.
boiling flow through pipes. ASHRAE Trans. 82 (2): 66–86. Shen, B. and Groll, E.A. (2005). A critical review of the
Shah, M.M. (1979). A general correlation for heat transfer influence of lubricants on the heat transfer and pressure
during film condensation inside pipes. Int. J. Heat Mass drop of refrigerants – part II: lubricant influence on
Transf. 22: 547–556. condensation and pressure drop. HVAC&R Res. 11 (4).
Shah, M.M. (1981). Heat transfer during film condensation in Shiraishi, M., Kikuchi, K., and Yamanishi, T. (1981).
tubes and annuli, a literature survey. ASHRAE Trans. 87 Investigation of heat transfer characteristics of a two-phase
(1): 1086–1105. closed thermosyphon. In: Advances in Heat Pipe
Shah, M.M. (2009). An improved and extended general Technology (ed. D.E. Reay), 95–104.
correlation for heat transfer during condensation in plain Shon, B.H., Jung, B.H., Kwon, J.I. et al. (2018). Characteristics
tubes. HVAC&R Res. 15 (5): 889–913. on condensation heat transfer and pressure drop for a low
Shah, M.M. (2013). General correlation for heat transfer GWP refrigerant in brazed plate heat exchanger. Int. J.
during condensation in plain tubes: further development Heat Mass Transf. 122: 1272–1282.
and verification. ASHRAE Trans. 119 (2). Short, G.D. (2017). Assessment of Lubricants for Ammonia
Shah, M.M. (2014). A new flow pattern based general and Carbon Dioxide Refrigeration Systems. Technical
correlation for heat transfer during condensation in Paper #5, 2017 IIAR Natural Refrigeration Conference and
horizontal tubes. Proceedings of the 15th International Heavy Equipment Expo, San Antonio, TX.
Heat Transfer Conference, IHTC-15, Kyoto, Japan, 10–15 Siddique, M.S., Golay, M.W., and Kazimi, M.S. (1993). Local
August, 2014. heat transfer coefficients for forced-convection
74 2 Heat Transfer During Condensation

condensation of steam in a vertical tube in the presence of Sukhatme, S.P. and Rohsenow, W.M. (1966). Heat transfer
a noncondensable gas. Nucl. Technol. 102: 386–402. during film condensation of a liquid metal vapor. J. Heat
Soliman, H.M. (1986). The mist-annular transition during Transf. 88: 19–28.
condensation and its influence on the heat transfer Sur, B. and Azer, N.Z. (1991). Effect of oil on heat transfer and
mechanism. Int. J. Multiphase Flow 12: 277–288. pressure drop during condensation of Refrigerant-113
Soliman, M., Schuster, J.R., and Berenson, P.J. (1968). A inside smooth and internally finned tubes. ASHRAE Trans.
general heat transfer correlation for annular flow 97 (1): 365–373.
condensation. J. Heat Transf. 90: 267–276. Taylor, R.G. and Krishna, R. (1993). Multicomponent Mass
Son, C. and Lee, H. (2009). Condensation heat transfer Transfer. New York: Wiley.
Tepe, J.B. and Mueller, A.C. (1947). Condensation and
characteristics of R-22, R-134a and R-410A in small
subcooling inside an inclined tube. Chem. Eng. Prog. 43 (5):
diameter tubes. Heat Mass Transf. 45: 1153–1166.
267–278.
Sparrow, E.M. and Gregg, J.L. (1959a). Laminar condensation
Thome, J.R. (1998). Condensation of fluorocarbon and other
heat transfer on a horizontal tube. J. Heat Transf. 81 (4):
refrigerants: a state-of-the-art review. Final report prepared
291–295.
for Air-Conditioning and Refrigeration Institute (ARI).
Sparrow, E.M. and Gregg, J.L. (1959b). A theory of rotating
Quoted in Shen and Groll (2005).
condensation. J. Heat Transf. 81: 113–120.
Thome, J.R., El Hajal, J., and Cavallini, A. (2003).
Sparrow, E.M. and Hartnett, J.P. (1961). Condensation on a Condensation in horizontal tubes. Part 2: new heat transfer
rotating cone. J. Heat Transf.: 101–102. model based on flow regimes. Int. J. Heat Mass Transf. 46:
Sparrow, E.M., Minkowycz, W.J., and Saddy, M. (1967). 3365–3387.
Forced convection condensation in the presence of Thonon, B. (1995). Design method for plate evaporators and
non-condensables and interfacial resistance. Int. J. Heat condensers. In: Proceedings of 1st Int. Conf. Process
Mass Transf. 10: 1829–1845. Intensification for Chemical Industry, 149–155. BHR Group
Spencer, D.L. and Ibele, D.E. (1966). Laminar film Conference Series Publication.
condensation of a saturated and superheated vapor on a Tichy, J.A., Macken, N.A., and Duval, W.M.B. (1984). An
surface with a controlled temperature distribution. In: experimental investigation of heat transfer in
Proceedings Third International Heat Transfer Conference, forced-convection-condensation of oil-refrigerant
7–12 August, Chicago, USA, 337–347. Begell House. mixtures. ASHRAE Trans. 90 (1).
Sreepathi, L.K., Bapat, S.L., and Sukhatme, S.P. (1996). Heat Traviss, D.P., Baron, A.B., and Rohsenow, W.M. (1973).
transfer during film condensation of R-123 vapour on Forced convection condensation inside tubes: heat transfer
horizontal integral-fin tubes. J. Enhanc. Heat Transfer 3: equation for condenser design. ASHRAE Trans. 79 (1):
147–164. 157–165.
Stein, R.P., Cho, D.H., and Lambert, G.A. (1985). Umur, A. and Griffith, P. (1965). Mechanism of dropwise
Condensation on the underside of a horizontal surface in a condensation. J. Heat Transf. 87 (1965): 275–282.
closed vessel. In: Multiphase Flow and Heat Transfer, vol. Varma, V.C. (1977). A study of condensation of steam inside a
tube. MSc thesis. Rutgers University, New Jersey.
47 (eds. V.K. Dhir, J.C. Chen and O.C. Jones). New York:
Vierow, K.M. and Schrock, V.E. (1991). Condensation in a
ASME HTD.
natural circulation loop with noncondensable gases. Part I.
Stephan, K. (1992). Heat Transfer in Condensation and
Heat transfer. In: Proceedings of the International
Boiling. Berlin: Springer Varlag. (English language edition).
Conference on Multiphase Flow, 183–186. Tsukuba, Japan:
Subbotin, V.I., Ivanovskii, M.N., Sorokin, V.P., and Chulkov,
International Atomic Energy Agency (IAEA).
B.A. (1964). Heat transfer during the condensation of
Wallis, G.B. (1969). One-Dimensional Two-Phase Flow. New
potassium vapour. Teplofiz. Vysok. Temper 2 (4): 616.
York: McGraw-Hill.
Quoted in Necme and Rose (1976). Walt, J.V.D. and Kroger, D.G. (1972). Heat transfer during
Sugawara, S. and Katsuta, K. (1966). Fundamental study on condensation of saturated and superheated Freon 12. In:
dropwise condensation. In: Proceedings of the Third Progress in Heat and Mass Transfer, vol. 6 (eds. G. Hetsroni,
International Heat Transfer Conference, Chicago, vol. 2, S. Sideman and J.P. Hartnett), 75–98. London: Pergamon
354. New York: ASME. Press.
Sukhatme, S.P. and Rohsenow, W.M. (1964). Heat transfer Wang, B.X. and Du, X.Z. (2000). Study on laminar film-wise
during film condensation of a liquid metal vapor. condensation for vapor flow in an inclined
Massachusetts Institute of Technology. Cambridge, MA. small/mini-diameter tube. Int. J. Heat Mass Transf. 43:
Report MIT-2995-1. 1859–1868.
References 75

Wang, H.S. and Honda, H. (2003). Condensation of Xing, F., Xu, J., Xie, J. et al. (2015). Froude number dominates
refrigerants in horizontal microfin tubes: comparison of condensation heat transfer of R245fa in tubes: effect of
prediction methods for heat transfer. Int. J. Refrig. 26: inclination angles. Int. J. Multiphase Flow 71: 98–115.
452–460. Xio, J. and Hrnjak, P. (2018). Flow regime map for
Wang, H.S. and Rose, J.W. (2005). A theory of film condensation from superheated vapor. 17th International
condensation in horizontal noncircular section Refrigeration and Air Conditioning Conference at Purdue
microchannels. J. Heat Trans.-T ASME 127 (10): 1096–1105. (9–12 July 2018), West Lafayette, Indiana, USA. https://
Wang, H.S. and Rose, J.W. (2006). Film condensation in docs.lib.purdue.edu/iracc/1910.
horizontal microchannels: effect of channel shape. Int. J. Xu, H., Sun, Z., Gu, H., and Li, H. (2016). Forced convection
Therm. Sci. 45: 1205–1212. condensation in the presence of noncondensable gas in a
Wang, H.S. and Rose, J.W. (2011). Theory of heat transfer horizontal tube; experimental and theoretical study. Prog.
during condensation in microchannels. Int. J. Heat Mass Nucl. Energy 88: 340–351.
Transf. 54 (11–12): 2525–2534. Yan, Y.Y., Lio, H.C., and Lin, T.F. (1999). Condensation heat
Webb, R.L. (1998). Convective condensation of superheated transfer and pressure drop of refrigerant R-134a in a plate
vapor. ASME J. Heat Transfer 120: 418–421. heat exchanger. Int. J. Heat Mass Transf. 42 (6): 993–1006.
Webb, R.L. and Murawski, C.G. (1990). Row effect for R-11 Yanniotis, S. and Kolokotsa, D. (1996). Experimental study of
condensation on enhanced tubes. J. Heat Transf. 112: water vapour condensation on a rotating disc. Int.
768–776. Commun. Heat Mass Transfer 23: 721–729.
Webb, R.L., Rudy, T.M., and Kedzierski, M.A. (1985). Yu, J. and Koyama, S. (1998). Condensation heat transfer of
Prediction of the condensation coefficient on horizontal pure refrigerants in microfin tubes. In: Proceedings of
integral-fin tubes. J. Heat Transf. 107 (2): 369–376. International Refrigeration Conference at Purdue University,
Welch, J.F. and Westwater, J.W. (1961). Microscopic study of 325–330. Purdue, Indiana, USA, West Lafayette: Purdue
dropwise condensation. In: International Developments in University.
Heat Transfer, Part II, 302–309. New York: ASME. Yu, J., Chen, J., Li, F. et al. (2018). Experimental investigation
Wellsandt, S. and Vamling, L. (2005). Prediction method for of forced convective condensation heat transfer of
flow boiling heat transfer in a herringbone microfin tube. hydrocarbon refrigerant in a helical tube. Appl. Therm.
Int. J. Refrig. 28: 912–920. Eng. 129: 1634–1644.
Wilcox, S.J. and Rohsenow, W.M. (1970). Film condensation Yunxiao, Y. and Li, J. (2015). Experimental investigation on
of potassium using copper condensing block for precise heat transfer coefficient during upward flow condensation
wall-temperature measurement. J. Heat Transf. 92C: 359. of R410A in vertical smooth tubes. J. Therm. Sci. 24 (2):
Williams, P.E. and Sauer (1981). Condensation of of 155–163.
refrigerant-oil mixtures on horizontal tubes. Int. J. Refrig. 4 Zeinelabdeen, M.I.M., Kamran, M.S., and Briggs, A. (2018).
(4): 209–222. Comparison of empirical models with an experimental
Wongwises, S. and Polsongkram, M. (2006). Condensation database for condensation on banks of tubes. Int. J. Heat
heat transfer and pressure drop of HFC-134a in a helically Mass Transf. 122: 765–774.
coiled concentric tube-in-tube heat exchanger. Int. J. Heat Zhang, J., Martin, R.K., Ommen, T. et al. (2019).
Mass Transf. 49: 4386–4398. Condensation heat transfer and pressure drop
Wurfel, R., Kreutzer, T., and Fratzscher, W. (2003, 2003). characteristics of R134a, R1234ze(E), R245fa and
Turbulence transfer processes in a diabatic and condensing R1233zd(E) in a plate heat exchanger. Int. J. Heat Mass
film flow in inclined tube. Chem. Eng. Technol. 26: 439–448. Transf. 128: 136–149.

You might also like