You are on page 1of 10

02638762/98/$10.00+0.

00 q Institution of Chemical Engineers Trans IChemE, Vol 76, Part A, February 1998

CONDENSATION HEAT TRANSFER FUNDAMENTALS


J. W. ROSE
Department of Engineering, Queen Mary and West eld College, University of London, London, UK

he paper gives an outline and discussion of those aspects of condensation heat transfer theory which are relatively well understood. For free convection condensation, the Nusselt approximations (neglect of inertia, convection and surface shear stress) are discussed separately. The effects of interphase matter transfer and variable wall temperature, not considered by Nusselt, are also discussed. Complications arising in forced convection condensation, where the condensate lm is signi cantly affected by surface shear stress, are outlined. The present status of dropwise condensation theory and measurements is also brie y reviewed. Keywords: condensation; lmwise; dropwise; interface resistance

INTRODUCTION Many factors in uence heat-transfer coef cients during condensation. The condenser surfaces may be wetted by the condensate, when lm condensation (the normal mode) occurs, or non-wetted when dropwise condensation occurs. The vapour may be quiescent or moving across the condensing surface at signi cant velocity. The condensate and vapour ows may be laminar or turbulent. The vapour may comprise more than one molecular species and not all species may condense. Condensate from higher or upstream surfaces will generally impinge on lower or downstream surfaces (inundation) and thereby affect the heat transfer. The problems encountered in practical condensers are too complex to permit detailed and accurate modelling, and condenser design generally incorporates a substantial amount of idealization and empiricism. Even for condensation of a pure (single constituent) vapour, there exist dif cult or intractable problems such as those associated with geometry (3-D ow with free-stream vapour velocity aligned neither with gravity nor the condensing surface), turbulence, calculation of vapour-condensate interface shear stress, irregularity (rippling) of vapour-condensate interface and inundation. Before these problems can be properly addressed, it is necessary to have a good appreciation of those aspects which are better understood, namely gravitycontrolled laminar lm condensation with simple geometry and, to a lesser extent, forced convection condensation and dropwise condensation. The Nusselt1 models for laminar free convection lm condensation on a vertical plate and horizontal tube are outlined and the Nusselt approximations (neglect of inertia, convection and surface shear stress) are discussed individually. The effects of interface resistance, arising from interphase matter transfer, and variable wall temperature, not considered by Nusselt, are also discussed. As is well known, the Nusselt model is remarkably accurate in the practical ranges of the relevant parameters. The forced convection condensation problem is less straightforward in that the surface shear stress evidently cannot be neglected and in typical practical circumstances neither of the extreme
143

(high and low condensation rate) approximations for the surface shear stress is generally valid. Moreover, typically the condensate ow is neither gravity- nor shear stressdominated and, in the case of the tube, vapour boundary layer separation adds signi cant complication. Approaches to the laminar forced-convection condensation problem are also discussed. Finally, in view of recent reports of promising new techniques for promoting dropwise condensation of steam, dropwise condensation theory and measurements are also brie y reviewed. THE NUSSELT THEORY The problem considered is shown in Figure 1. Nusselt analysed the case of condensation of a pure, quiescent, saturated vapour on an isothermal plate and on an isothermal horizontal cylinder. He neglected shear stress from the vapour at the surface of the condensate lm as well as inertia/acceleration and convection in the lm. Equilibrium was assumed at the condensate-vapour interface so that the temperature at the outer surface of the condensate lm was taken as the vapour temperature. Condensate properties were assumed independent of temperature and, for the case of the horizontal tube, the lm thickness was assumed small in comparison with the tube radius. For the vertical at plate and with these assumptions, equating the net downward force on an element of the condensate lm at distance x down the surface and having width dy and height dx to zero (rather than mass acceleration) gives:
2 u D q m 2+g q y

=0

(1)

Here the vertical pressure gradient term (equal to pressure gradient in the remote vapour) omitted by Nusselt, has been included and leads to the buoyancy term with D q rather than the condensate density q ; this only affects the result for condensation near the critical point. Integration of equation (1) twice with boundary conditions of zero velocity at the wall and zero velocity gradient at the outer surface of the lm gives the velocity pro le across the lm. The product of

144

ROSE

velocity with condensate density and plate width may be integrated to give the mass ow rate in the lm. When the increment in mass ow rate over dx is equated to the condensation rate on the surface of the element (the latter written in terms of the heat ux obtained by assuming pure conduction across the lm equation) a simple differential equation for lm thickness d in terms of distance down the plate is obtained. This may be integrated to give the local and mean heat ux. The result for the mean Nusselt number is
NuL
where J

and GrL

=q

3 D q gL / g

(4)

a L

qL =k D T

2 GrL 3 J

1/ 4

(2)

=k

D T / g hfg

(3)

When the heat ux rather than D T is regarded as uniform (independent of height), the result for the mean Nusselt number is identical to equation (2) except that the constant q replaces the mean value q (in NuL ) and the mean D T replaces the constant D T. In the case of the horizontal tube, the same approach, with the additional assumption that the condensate lm thickness is small compared with the tube radius, leads to the following differential equation for the lm thickness dz 4 sin h z cos h - 2 = 0 (5) dh + 3 where the dimensionless lm thickness q D q ghfg 4 z= d (6) g k dD T
The solution of equation (5) subject to the condition that z is nite at h = 0, or by symmetry dz/dh = 0, is
2 z = 4/ 3 sin h
h

sin1/ 3 h dh
0

(7)

as may be readily veri ed by differentiation. Using equations (6) and (7), the local heat ux q = k D T / d may be written q

q D q ghfg k 3 D T 3 g d

1/ 4

- 1/ 4

- 1/ 4

sin

1/ 3

h
0

sin

1/ 3

h dh
(8)

The mean heat ux up to angle h is then given by qh


where w (h ) =
1 1 21/ 4 h
h
h
0

=h

qdh
0

q D q ghfg k 3 D T 3 g d

1/ 4

w (h )

(9)

sin1/ 3 h
1/ 4

= dh

sin1/ 3 h dh
0

=
so that qh

4 1 21/ 4 3 h

3/ 4

sin
0

1/ 3

h dh

(10)

=
h

q D q ghfg k 3 D T 3 g d

1/ 4

4 21/ 4 3

3/ 4

h
0

sin

1/ 3

h dh
(11)
3/ 4

When

= p we obtain the mean ux for the tube


q D q ghfg k 3 D T 3 g d
1/ 4 p

4 q = 1/ 4 2 3p
Figure 1. Condensation on a vertical plate and horizontal tube.

sin1/ 3 h dh
0

(12) Trans IChemE, Vol 76, Part A, February 1998

CONDENSATION HEAT TRANSFER FUNDAMENTALS

145

and
p

sin1/ 3 h dh
0

=2

7/ 3

p 2 / [C (1/ 3)]3

(13)

where C is the gamma function. Thus q


where K and
a

=K

q D q ghfg k 3 D T 3 g d
1/ 2

1/ 4

(14)

= (8/ 3)(2p =
q D T
a d

) [C (1/ 3)]-

9/ 4

= 0.728 018
1/ 4

(15)
Figure 2. Condensation on a horizontal tube. Dependence of lm thickness on angle.

q D q ghfg k = K g dD T

(16)
3 1/ 4

Nu =

=K

q D q ghfg d g k D T

(17)

It is interesting to note that Nusselt used planimetry to evaluate the integral in equation (7) to obtain values of z as a function of h . These were used to obtain the local heat ux q

k D T d

q D q ghfg k 3 D T 3 g d

1/ 4

z- 1/ 4

(18)

and hence, again using planimetry, the mean heat ux for the tube q

condenser tube, it is strictly necessary to calculate the local heat-transfer rate from vapour to coolant using a value for the coolant-side heat-transfer coef cient and taking account of two dimensional heat transfer in the tube wall. Separate solutions are needed for all condensing uids, vapour and coolant temperatures, coolant-side heat-transfer coef cients and tube diameters and thermal conductivities. The procedure is described by Honda and Fujii3 . Memory and Rose4 noted that measured wall temperature pro les are closely approximated by cosine curves (see Figure 3). The vapour-to-surface temperature difference in this case may be written
(22) D T = D T(1 - A cos h ) where A is a constant which may take values between 0 (when the temperature difference is uniform around the tube, i.e. the Nusselt case) and unity (the extreme case where the temperature difference is zero at the top of the tube). Treating the problem otherwise, as in the Nusselt solution, the following differential equation is obtained for the lm thickness dz 4 2(1 - A cos h ) (23) + 3 z cot h =0 dh sin h The solution of equation (23) with the condition that z

=p

qdh
0

(19)

It is remarkable that Nusselt s results, obtained by multiple planimetry, give a value of 0.725 for the constant K, an error less than 0.5%. Before leaving the Nusselt solution, the situation near the bottom of the tube warrants comment. Here the lm thickness predicted by the theory increases to in nity as h approaches p and the assumption that d is small compared with the tube radius is invalid. In the case of the Nusselt problem with an isothermal surface, when D T is uniform, this is not too serious since the erroneous heat ux values near the bottom of the tube where the lm is thick are small and make relatively small contribution to the total heat transfer and hence to the mean heat ux for the tube. For the case of uniform heat ux considered by Fujii et al.2 , the mean value of D T is evaluated to obtain the mean heat-transfer coef cient, otherwise using the Nusselt assumptions. Here
q D T= p k
p

d dh
0

(20)

when, as may be seen from Figure 2, the contribution to the integral in equation (20) from the erroneous values as h approaches p , is signi cant. Fujii et al.2 obtained
Nu = 0.695
q D q ghfg d 3 g k D T
1/ 4

(21)
Figure 3. Surface temperature variation around a horizontal tube. Condensation of ethylene glycol (see Memory and Rose 4 ). The broken lines are cosine ts.

In practice neither uniform D T nor uniform q is found and, in order to determine total or mean heat-transfer rate for a
Trans IChemE, Vol 76, Part A, February 1998

146

ROSE et al.5 ) that the appropriate mean temperature at which the viscosity should be evaluated is T*

remains nite at the top of the tube is


h

2 sin1/ 3 h dh

z=

4/ 3 - 2 sin h 4/ 3

3A

= (3/ 4)T + (1/ 4)T


w
y

(27)

sin

h
1/ 4 3

(24)

The density and thermal conductivity may be taken at the arithmetic mean of the wall and vapour temperature. INTERFACE TEMPERATURE DROP

De ning the dimensionless mean heat ux q*

=q

g d q D q ghfg k 3 D T

(25)

we have, from the de nition of z and for radial conduction across the lm q*

= (1 - A cos h

)z-

1/ 4 *

(26)

Figure 4 shows the variation of q around the tube for various values of A. It is seen that for values of A greater than zero the heat ux at rst increases, where the effect of the increasing value of D T outweighs that of increasing lm thickness, before passing through a maximum and decreasing to zero at the bottom of the tube where the lm thickness becomes in nite. The area under the curves, which is proportional to the mean heat ux, is found to be constant to the fourth signi cant gure for all values of A so that, in all cases, the mean Nusselt number is given by equation (17) with mean values of both q and D T, and the leading constant is 0.7280. On this basis it is suggested that, unless a conjugate vapour-to-coolant solution is carried out, the Nusselt equation should yield an accurate mean heattransfer coef cient when the wall temperature is nonuniform and, in particular, should be better than equation (21) for the uniform heat ux case. Nusselt s neglect of the temperature dependence of the viscosity, density and thermal conductivity of the condensate is also not too serious, and the variation of properties across the condensate lm is usually small with the exception, in some cases, of viscosity. Even in the case of viscosity, it is seen from equations (2) and (17) that the heat-transfer coef cient depends only on the quarter power of viscosity. By using the approximation that reciprocal viscosity is linear in temperature, and otherwise adopting the Nusselt approach, it is readily shown (see Mayhew

As noted above, Nusselt assumed that the temperature at the outer surface of the condensate lm was equal to that of the vapour. Only at equilibrium, when there is no net condensation, are the condensate surface and vapour temperatures equal. Although the interphase matter transfer problem cannot yet be said to be fully understood, theoretical expressions for the temperature drop between the vapour and condensate surface at the interface are available. Apart from other uncertainties, theoretical models incorporate a condensation coef cient de ned as the fraction of vapour molecules striking the condensate surface which remain in the liquid phase. It was earlier thought that the condensation coef cient might take values less than 0.01. In this case the predicted interface temperature drop could be signi cant in comparison with the temperature drop across the condensate lm. However, it is generally now thought that the condensation coef cient is close to unity, in which case the interface temperature drop is generally negligible. A notable exception is the case of liquid metals. Figure 5 shows experimental data for condensation of mercury. The temperature drop across the condensate given by the Nusselt theory in this case is in the range 0.10.5 K at the lowest vapour temperature and 2.5 5.5 K at the highest vapour temperature. It is evident that for liquid metals the Nusselt theory cannot be used without allowance for interphase mass-transfer resistance; in fact, in many cases, the temperature drop across the condensate lm is negligible in comparison with the interface temperature drop.

Figure 4. Condensation on a horizontal tube. Dependence of dimensionless heat ux on angle for cosine surface temperature distributions (see Memory and Rose4 ).

Figure 5. Condensation of mercury on a vertical plate at various vapour temperatures (Niknejad and Rose 23 ).

Trans IChemE, Vol 76, Part A, February 1998

CONDENSATION HEAT TRANSFER FUNDAMENTALS

147

CORRECTIONS TO NUSSELT MODEL Effect of Vapour Shear Stress As the condensate lm falls, it is retarded by the `stationary vapour and consequently the lm will be thicker than when vapour shear stress is neglected, as in the Nusselt theory. To incorporate vapour shear stress in the Nusselt model, the condition of zero velocity gradient at the outer edge of the lm is replaced by continuity of velocity and shear stress at the interface and the condensate problem must be solved simultaneously with the momentum equation for the vapour:
2 U U U U + V =m 2 x y y subject to the conditions in the remote vapour
y

(28)

Nusselt number decreases to an asymptotic value 2- 1 /2 of the Nusselt value with increasing J. Comparison of equation (35) with a more accurate solution where the interface velocity and shear stress are determined by simultaneous solution of the condensate and vapour momentum equations (see Maekawa and Rose6 ) is shown in Figure 7. It is seen that the more accurate solution involves the additional parameter R(= q g / q v g v )1/ 2 . Noting that in practice R is generally larger than 50, it is evident that the approximate solution is very accurate. Then, referring to Figure 6, and noting that in practice the upper limit of J is around 0.01, it is clear that the effect of vapour shear stress in free convection condensation is very small. Effects of Inertia and Convection in the Condensate Film

and at the interface u


u g y
m

U!

and

U 0 for y ! y !

(29)

=U =u
=
g
y

(30)
(31) (32)

These effects may be considered by treating the condensate lm using the boundary-layer equations for the condensate lm rather than the Nusselt approximations. Thus, for the at plate case, we have

= -q

U y
y

It may be noted that strictly it is the velocities and shear stresses tangential to the condensate surface which should be equated and the condensation mass ux is more accurately given by dd m = -q V + q U (33) dx Since the lm thickness increases slowly with distance down the plate (1/4 power dependence) these approximations should not lead to appreciable error. Boundary layer theory indicates that, for in nite condensation rate, the surface shear stress is given by
y y

u m + =0 x y 2 u u g D q m u u +m = q + y2 x y
T T u + m x y

(36)
(37)
(38)

T = j y2
2

= - mu = - (q/ h
i

fg

)ui

(34)

With this approximation it is no longer necessary to consider the vapour momentum equation and the Nusselt theory with the surface shear stress given by equation (34) yields
Nux Nux,Nu

J+4 4(J + 1)

1/ 4

(35)

Equation (35) is shown in Figure 6 where it is seen that the

expressing the conservation of mass, momentum and energy in the lm. It is seen that when the inertia or acceleration terms on the LHS of equation (37) and the convection terms on the LHS of equation (38) are omitted, the equations are those of Nusselt. Numerical solutions of these equations have been obtained by Sparrow and Gregg7 , Koh et al.8 , Chen9 and Fujii10 . These yielded virtually identical results except at low condensate Prandtl number where the Nusselt numbers obtained by Sparrow and Gregg, who did not consider surface shear stress, are higher. As discussed by Maekawa and Rose6 , when only the inertia effects are included the normalized Nusselt number depends only on the parameter J and the solution for this case is shown in Figure 8. Since, as noted above, the upper limit of J is around 0.01, it is clear that inertia effects are unimportant. Higher (erroneous) values of J may be found in the literature for condensation of liquid metals. These are due to temperature drops in the vapour, resulting from the

Figure 6. Effect of vapour shear stress.

Figure 7. Effect of vapour shear stress.

Trans IChemE, Vol 76, Part A, February 1998

148

ROSE

Chen9 has given the approximate equations


Nu NuNu
Nu NuNu
Nu NuNu

=
=
=

1 + 0.68H + 0.02HJ 1 + 0.85J - 0.15HJ

1/ 4

(43)
1/ 4

1 + 0.68PrJ + 0.02PrJ 2 1 + 0.85J - 0.15PrJ 2

(44)
1/ 4

1 + 0.68H + 0.02H 2 / Pr 1 + 0.85H / Pr - 0.15H 2 / Pr

(45)

Figure 8. Effect of inertia.

presence of non-condensing gas, or signi cant interface temperature drops being included in D T, the temperature drop across the condensate lm. When only convection effects are included, i.e. the boundary-layer energy equation is used with the Nusselt momentum equation, the normalized Nusselt number depends only on the parameter H(= cP D T / hfg ) as indicated in Figure 9. Since in practice H does not exceed a value of around 0.1, it is clear that convection is also unimportant. For the general case where both inertia and convection are included and the boundary-layer equations are solved subject to boundary conditions of zero velocity at the wall, continuity of velocity and shear stress at the interface, uniform wall and vapour temperatures, results of the form Nu/ NuNu or, since Pr Nu/ NuNu or Nu/ NuNu

= w (J , H) = H/ J = w (J , Pr) =n
(H , Pr)

(39)

(40) (41) (42)

are obtained. As indicated by Maekawa and Rose6 , the result obtained when only inertia effects are included corresponds to the general case when the condensate Prandtl number is zero and the result when only convection effects are included corresponds to the general case in the limit of in nite condensate Prandtl number.

which summarize the numerically-obtained results to within 1% to well beyond the practical ranges of J and H. As noted above, when the convective terms on the LHS of equation (38) are omitted, we have the Nusselt equation which leads to a linear temperature distribution across the condensate lm and consequently no account is taken of the addition to the wall heat ux due to condensate subcooling; omission of the convection terms automatically precludes subcooling. Approximate corrections of the Nusselt theory to include the effect of condensate subcooling have been given. A simple approach, retaining the Nusselt linear temperature pro le, indicates that the speci c latent heat in the Nusselt expression should be replaced by hfg + (3/ 8)cP D T . When the boundary-layer energy equation is used, as in Chen9 , the heat ux, and hence the heattransfer coef cient and Nusselt number, are evaluated from the temperature gradient in the condensate at the wall so that condensate subcooling is included. The temperature gradient and heat ux (equal to the condensation mass ux multiplied by the latent heat) at the outer surface of the condensate lm are smaller. It is interesting to note that a more accurate subcooling correction by Rohsenow11 gives hf g + 0.68cP D T to replace hf g . This is the same as equation (43) with J = 0, i.e. Chen s solution when convection but not inertia effects are taken into account. It should be noted that when using equations (43)(45), no adjustment to hf g should be made. The case of the horizontal tube has been treated by Sparrow and Gregg12 and Chen9 and is less straightforward because true similarity only exists near the top of the tube. However, based on the at plate results, inertia and convection effects are expected to be small also for the tube. This is shown by Chen9 who shows that the pro les are quite similar for angles up to 1508 and, since there is relatively little heat transfer on the lower part of the tube where the lm becomes thick, concludes that equations (43)(45) may also be used for the mean Nusselt numbers for the tube. Chen s equations give a correction factor to the Nusselt theory which include both the negative effects of vapour shear stress and condensate inertia and the positive effects of convection and subcooling. FORCED CONVECTION CONDENSATION The simplest case, namely condensation with vapour ow parallel to a horizontal, isothermal plate has been studied by Cess13 and Koh14 ; gravity is not involved and the vapour shear stress, which is responsible for the condensate motion, evidently cannot be neglected. When the Nusselt approach is used for the condensate lm, i.e. inertia and convection are ignored, and with continuity of velocity and shear stress at the condensate surface (assumed parallel to the plate), Trans IChemE, Vol 76, Part A, February 1998

Figure 9. Effect of convection.

CONDENSATION HEAT TRANSFER FUNDAMENTALS together with the assumption that the vapour free-stream velocity is much greater than the condensate surface velocity, the following results are obtained: For condensation rate approaching zero:

149

x Nux Re- 1/ 2
where G

= 0.436G-

1/ 3

(J !

0)

(46) (47)

= RJ = 0.5 = 0.332
(J !

and for condensation rate approaching in nity

x Nux Re- 1/ 2

(48)
Figure 11. Shear stress on condensate surface for condensation on a horizontal plate (after Cess 13 ) the broken lines are the high and low condensation rate asymptotes.

Equation (46) is obtained when the surface shear stress is given by


2 2 s / q v U Re1/,x v

(49) (50a) should be replaced by 2/3 and 1/3 respectively (see Mayhew et al.5 and Fujii10 ). Figure 11 illustrates the dependence of surface shear stress on G. It may be noted that, unlike the case of free convection condensation, the simple asymptotic expression for the surface shear stress given by equation (50) is not generally valid in the practical range of G (0.0110). However, owing to its simplicity, this has been widely used, particularly in treating the practically more important case of the horizontal tube. It may be seen from Figure 11 that this will underestimate the shear stress and therefore give increasingly conservative results at low values of G as may be seen from equation (48) in Figure 10. For condensation on a vertical plate or horizontal tube, not only is neither asymptotic expression for the surface shear stress valid, but gravity also has to be taken into account. This adds no dif culty in principle (except in the case of vapour up ow when instability of the condensate lm may arise from the opposing forces of gravity and vapour shear stress). When equation (50) is used for the vapour shear stress it is found that

and equation (48) when s

= mU
2 x s / q U Re1/ 2

which leads to

= J/ 2 = G/ 2

(50b) (50c)

or
2 s / q v U)Re1/,x2 v

(note that q v and g v cancel in equation (50c)). In the general case

x Nux Re- 1/ 2

= w (G) = 0.45

(51)

For the practical range of G (0.0110) an approximate expression for equation (51) has been given by Fujii and Uehara15 :

x Nux Re- 1/ 2

1.2 + G- 1

1/ 3

(52)

which is shown in Figure 10 and is in good agreement with the numerical solution. It may be seen from the above that the lm thickness (= x/Nux ) varies as x1 /2 so the approximation that the condensate surface is parallel to the plate is less accurate than for the free convection case for the vertical plate. When the condensate lm is treated on the basis of the boundary-layer equations so that the inertia and convection terms are included, it is found that inertia effects terms are not important and convection is only important at high values of G and R, which rarely occur in practice. In these cases the results given by equation (52) are conservative. To allow for variable viscosity in forced convection, the factors 3/4 and 1/4 in the reference temperature equation (27)

NuRe- 1/ 2
where F
1

= w (F)
2

(53) (54)

= J - (Gr/ Re )

Figure 10. Solution for forced convection condensation on a horizontal plate (after Cess 13 ).

F is sometimes written Pr/Fr H. This notation is somewhat misleading since cP , and hence Pr, are not involved in the problem and the fact that F measures the relative importance of gravity (through Gr) and vapour velocity (through Re2 ) is less apparent. Roughly speaking, gravity effects begin to dominate when F > about 1 and vapour shear stress becomes more important than gravity when F <about 1. Unfortunately, F = 1 lies roughly in the centre of the practical range. For condensation on a horizontal tube, the problem is further complicated by other factors (variation of surface shear stress and pressure around the tube and vapour boundary layer separation). Results of various investigators are discussed by Rose16 . Figure 12 shows representative results for condensation of steam on a horizontal tube obtained by somewhat different approaches. Sugawara et al.17 evaluated the shear stress at the condensate surface as in the case of single-phase ow, i.e. they neglected the effect of condensation (equivalent to equation (49) for the at plate case); the slope discontinuity in Figure 12 shows the point of vapour boundary-layer separation beyond which the surface shear

Trans IChemE, Vol 76, Part A, February 1998

150

ROSE

Figure 12. Forced convection condensation of steam on a horizontal tube with vertical vapour down ow predictions of various models (after Fujii et al.15 ).

stress was set to zero. Shekriladze and Gomelauri18 used the opposite extreme (in nite condensation rate) to approximate the surface shear stress (equivalent to equation (50) for the at plate case). As noted in the at plate case, both of the extreme shear stress approximations are conservative i.e. they underestimate the shear stress and hence also the heattransfer coef cient. Fujii et al.15 attempted to evaluate the surface shear stress correctly using the interface conditions equations (3032) (where y is measured radially outwards). Fujii et al. used an approximate quadratic pro le for the velocity distribution across the vapour boundary layer; with this approximation the radial velocity gradient in the vapour is always positive so that boundary layer separation is precluded in this approach and the curve in Figure 12 has no slope discontinuity. For most of the tube surface it is seen that the approach of Fujii et al. yields the highest heattransfer coef cient, as expected. DROPWISE CONDENSATION Dropwise condensation, which may occur when the condensing surface is not wetted by the condensate, was rst noted by Schmidt et al.19 but only became better understood in the 1960s. The fact that the heat-transfer coef cient during dropwise condensation of steam is much higher than for lm condensation has continued to stimulate interest in this topic. Early experimental data were widely scattered and not until 1964 was there a successful theory of dropwise condensation heat transfer. Here the focus is on what is known. The broad facts are that dropwise condensation has only been obtained with a few relatively high surface tension uids notably steam, a few organic uids and mercury. Except in the case of mercury, metal surfaces, in the absence of impurities, are wetted by the condensate and lm condensation ensues; non-wetting agents (`promoters ) are needed to promote dropwise condensation without themselves offering signi cant thermal resistance and thereby offsetting the

advantage of dropwise condensation. For steam, good promoters, such as dioctadecyl disulphide, which form monomolecular layers on the condenser surface with negligible resistance to heat transfer, have been found. These have given lifetimes of thousands of hours on copper or copper-containing surfaces under clean laboratory conditions, thereby permitting experimental determination of heat-transfer coef cients for dropwise condensation. Mercury condenses in the dropwise mode on some metallic surfaces such as stainless steel but owing to its high liquid thermal conductivity, the advantage of dropwise over lm condensation is much less marked than is the case for steam. With other uids such as ethylene glycol, dropwise condensation has been obtained in the laboratory but generally with more dif culty and with shorter lifetimes than for steam. Despite the early promise of the monolayer promoters, no satisfactory method for promoting dropwise condensation industrially have yet been found. The fact that for steam the heat-transfer coef cient for dropwise condensation is known to be around ten times that for lm condensation at power station condenser pressures (and around 20 at atmospheric pressure) continues to stimulate the search for a practical means of promoting this mode of condensation. This brief review is prompted by the fact that in some recent studies heat-transfer coef cients are reported which are evidently inaccurate (as were the early measurements). Figure 13 shows reliable data from several investigations at atmospheric pressure and Figure 14 shows data for lower pressures. It is evident that the heat-transfer coef cient depends on pressure and D T or q (as shown by the curvature of the q-D T graphs). In contrast to lm condensation, the heat-transfer coef cient increases with vapour-to-surface temperature difference or heat ux reaching an almost constant value. The data for steam are summarized by the empirical equation
q kW/ m2
a

=t

0 .8

D T D T2 + 0.3 K2 K
D T K

(55)

where t is Celsius temperature, or


kW/ m2 K

=t

0 .8

5 + 0.3

(56)

Figure 13. Dropwise condensation of steam at atmospheric pressure (from Rose 22 ).

Trans IChemE, Vol 76, Part A, February 1998

CONDENSATION HEAT TRANSFER FUNDAMENTALS

151

where C1 and C2 are shape factors (for steam C1 = 2/3 and C2 = 1/2). The terms in the denominator are the conduction and interface resistances, while the numerator is the `driving force given by the vapour-to-surface temperature difference less the amount by which the vapour must be cooled before condensation can occur and the convex surface. For the distribution of drop sizes Le Fevre and Rose used f = 1 - (r/ r)
n

dependence of saturation pressure on surface curvature dominates (the vapour must be cooled below its normal saturation temperature before condensation can occur on the convex liquid surface). For the drops a little larger than the smallest, the interphase matter transfer resistance becomes important (this is responsible for the dependence of heattransfer coef cient on pressure), while for the larger drops, the conduction resistance dominates. A general theory, therefore, must include all three effects. Le Fevre and Rose obtained, for the heat ux through the base of a hemispherical drop (a good approximation for water) having radius r, 2r T D T - rq h fg q= (57) r T c + 1 Rg T 1/ 2 C1 + C2 2 k hfg q m c - 1 2p

(58)

Figure 14. Dropwise condensation of steam at sub-pressures (from Rose22 ).

It is suggested that equation (55) be used to validate equipment and techniques. In order to assess the effectiveness of new promoters, the apparatus should rst be used with a copper condensing surface and a recognized monolayer promoter to check that results conforming (to within say 1 K) with equation (55) can be obtained. Lower heat-transfer coef cients which decrease with increasing D T are caused by the presence in the steam of non-condensing gases even minute gas concentrations can give large increases in D T unless steps are taken to avoid this. There is still some debate regarding the importance or otherwise of `constriction resistance caused by nonuniformity of surface heat ux and consequent dependence of heat-transfer coef cient on conductivity of the surface material. It seems clear, however, that this effect is small except at low heat ux (see Rose20 ) and would be of little importance in a power station condenser. Outline of Theory The general approach used in the model of Le Fevre and Rose21 and outlined by Rose20,22 is to combine an expression for the heat transfer through a single drop with an expression for the distribution of drop sizes to give the mean heat ux for the condenser surface as a function of vapour-to-surface temperature difference. The range of drop sizes (typical radius of the largest drops, around 1 mm, is 106 times that of the smallest drops) is such that for the smallest, the
Trans IChemE, Vol 76, Part A, February 1998

where f is the fraction of surface area covered by drops having base radius greater than r. It is seen that equation (58) indicates that no area is covered by drops larger than the largest and that if the smallest drop had zero radius, all of the area would be covered. n is a constant (for steam n = 1/3) and r is the base radius of the largest drop. It follows from equation (58) that the fraction of the surface area covered by drops in the size range r, r + dr is r n- 1 dr A(r)dr = n (59) r r and, the mean heat ux for the surface is given by
n q= n r
r

2r D T T - rq hfg
C1 r C2 c + 1 k T + q m h2 c - 1 fg
Rg T 2p
1/ 2

rn- 1 dr

(60) The lower and upper limits of integration are r and r = C3 (r / q g)1/ 2 (62) respectively; the former is given by the radius of the smallest viable drop (owing to surface curvature smaller drops would evaporate) and the latter was obtained from dimensional analysis, the constant C3 (= 0.4) being found empirically. The integral in equation (60) can be evaluated and the result obtained given in closed form. As may be seen from Rose20,22 the results are in very good agreement with experimental measurements, predicting the correct dependence on D T and pressure. It seems probable, however, that if dropwise condensation becomes a practical proposition

= 2r

/ q hfg D T

(61)

152

ROSE
j
k m m
y

it is only likely to be used for steam. For pressures not exceeding atmospheric the simple empirical equations (55) and (56) are adequate. The theory may be used to determine heat-transfer coef cients for steam at higher pressures or for other uids.
q q r s

NOMENCLATUR E
A A(r)dr C1 C2 C3 cP d F Fr f constant in equation (22) fraction of surface area covered by drops in the size range r, r + dr constant in equation (57) constant in equation (57) constant in equation (57) isobaric speci c heat capacity of condensate diameter of tube de ned in equation (54) 2 2 U/ gx or U/ gd fraction of surface area covered by drops having base radius greater than r
k D T g hf 8
3

k / q cP thermal conductivity of condensate kinematic viscosity of condensate kinematic viscosity of vapour density of condensate density of vapour surface tension shear stress at condensate surface

REFERENCES
1. Nusselt, W., 1916, Die Ober achenkondensation des Wasserdampfes, Z Vereines Deutsch Ing, 60: 541546, 569575. 2. Fujii, T., Uehara, H. and Oda, K., 1972, Film condensation on a surface with uniform heat ux and body force convection, Heat Transfer Jpn Res, 4: 7683. 3. Honda, H. and Fujii, T., 1984, Condensation of a owing vapour on a horizontal tube numerical analysis as a conjugate heat transfer problem, J Heat Transfer, 106: 841848. 4. Memory, S. B. and Rose, J. W., 1991, Free convection laminar lm condensation on a horizontal tube with variable wall temperature, Int J Heat Mass Transfer, 34: 27752778. 5. Mayhew, Y. R., Grif ths, D. J. and Philips, J. W., 1965, Effect of vapour drag on laminar lm condensation on a vertical surface, Proc I Mech E, 180: 280287. 6. Maekawa, T. and Rose, J. W., 1997, Laminar natural convection lm condensation. (In preparation). 7. Sparrow, E. M. and Gregg, J. L., 1959, A boundary layer treatment of laminar- lm condensation, J Heat Transfer, 81: 1318. 8. Koh, J. C. Y., Sparrow, E. M. and Hartnett, J. P., 1961, The two phase boundary layer in laminar lm condensation, Int J Heat Mass Transfer, 2: 6982. 9. Chen, M. M., 1961, An analytical study of laminar lm condensation: part 1 Flat plates and part 2 Multiple horizontal tubes, J Heat Transfer, 83: 4854, 5560. 10. Fujii, T., 1991, Theory of Laminar Film Condensation, (SpringerVerlag). 11. Rohsenow, W. M., 1956, Heat transfer and temperature distribution in laminar lm condensation, Trans ASME, 78: 16451648. 12. Sparrow, E. M. and Gregg, J. L., 1959, Laminar condensation heat transfer on a horizontal cylinder, J Heat Transfer, 81: 291295. 13. Cess, R. D., 1960, Laminar- lm condensation on a at plate in the absence of a body force, Z Angew Math Phys, 11: 426433. 14. Koh, J. C. Y., 1962, Film condensation in forced convaction boundarylayer ow, Int J Heat Mass Transfer, 5: 941954. 15. Fujii, T., Uehara, H. and Kurata, C., 1972, Laminar lmwise condensation of owing vapour on a horizontal cylinder, Int J Heat Mass Transfer, 15: 235246. 16. Rose, J. W., 1988, Fundamentals of condensation heat transfer: Laminar lm condensation, JSME Int J Series 2, 31: 357375. 17. Sugawara, S, Michiyoshi, I. and Minamiyama, T., 1956, The condensation of vapour owing normal to a horizontal pipe, Proc Sixth Jap Nat Cong App Mech paper III-4, 385388. 18. Shekriladze, I. G. and Gomelauri, V. I., 1996, The study of laminar lm condensation of owing vapour, Int J Heat Mass Transfer, 9: 581591. 19. Schmidt, E, Schurig, W. and Sellschop, W., 1930, Versuche uber die Kondensation von Wasserdampf in Film- und Tropfenform, Tech Mech u Thermodynamik (Forschung Ing Wes), 1: 5263. 20. Rose, J. W., 1994, Dropwise condensation, Heat Exchanger Design Handbook Update, 2.6.5. 21. Le Fevre, E. J. and Rose, J. W., 1966, A theory of heat transfer by dropwise condensation, Proc Third Int Heat Transfer Conf, Chicago, Vol. 2: 362375. 22. Rose, J. W., 1988, Some aspects of condensation heat transfer theory, Int Communications in Heat and Mass Transfer, 15: 449473. 23. Niknejad, J. and Rose, J. W., 1981, Interphase matter transfer an experimental study of condensation of mercury, Proc R Soc Lond, A378: 305327.

G
Gr g H hfg J K L m Nu NuNu Nux Nux,Nu NuL n Pr q q q* qh R Re Rex Rev,x Rg r r r T Tv Tw T* t U U u ui V
y

q g q g
y y

1/ 2

x y z

q D q gd /g or q D q gxc3 / g 2 speci c force of gravity cPD T/hfg speci c latent heat of evaporation k D T/g hfg constant, see equation (15) height of plate condensation mass ux Nusselt number Nusselt number given by simple Nusselt theory local Nusselt number local Nusselt number given by simple Nusselt theory mean Nusselt number for plate of height L constant, see equations (58) and (59) Prandtl number local heat ux mean heat ux dimensionless heat ux, de ned in equation (25) local heat ux for tube (q g /q g )1/2 Uq d/ g or Uq x/ g Uq x/ g Uq x/ g speci c gas constant radius of tube, radius of drop base radius of largest drop radius of smallest drop temperature vapour temperature wall temperature reference temperature Celsius temperature x-wise vapour velocity vapour free-stream velocity x-wise condensate velocity x-wise interface velocity y-wise vapour velocity y-wise condensate velocity distance from top of plate or tube (measured around surface) distance normal (outward) from plate or tube dimensionless lm thickness (see equation (6))
2
y y y y

Greek letters a heat transfer coef cient a mean heat-transfer coef cient c ratio of principle speci c heat capacities of vapour Tv - Tw D T Tv - Tw D T q v-q D q d thickness of condensate lm g viscosity of condensate g viscosity of vapour h angle with vertical measured from top of tube
y

ADDRESS
Correspondence concerning this paper should be addressed to Professor J. W. Rose, Department of Engineering, Queen Mary and West eld College, University of London, Mile End Road, London E1 4NS. This paper was presented at the 5th UK National Heat Transfer Conference, held at Imperial College London 17 18 September 1997.

Trans IChemE, Vol 76, Part A, February 1998

You might also like