Sunteți pe pagina 1din 9

International Journal of Heat and Mass Transfer 55 (2012) 1496–1504

Contents lists available at SciVerse ScienceDirect

International Journal of Heat and Mass Transfer


journal homepage: www.elsevier.com/locate/ijhmt

Analysis of heat-transfer performance of cross-flow fin-tube heat exchangers


under dry and wet conditions
Cheen Su An, Do Hyung Choi ⇑
Department of Mechanical Engineering, Korea Advanced Institute of Science and Technology, Daejeon 305-701, Republic of Korea

a r t i c l e i n f o a b s t r a c t

Article history: A three-dimensional analysis procedure for the detailed phenomenon in a fin-tube heat exchanger has
Received 16 June 2011 been developed and applied to predict the heat/mass transfer characteristics of the wave-fin heat
Accepted 7 August 2011 exchangers. The continuity, Navier–Stokes and energy equations together with the species equation for
Available online 29 November 2011
the air–vapor mixture are solved in a coupled manner, so that the inter-dependence between the temper-
ature and the humidity can be properly taken into account, by using the SIMPLE-type finite volume
Keywords: method. Having validated the procedure, calculations have been carried out for various frontal-velocity
Wave-fin heat exchanger
and inlet-humidity conditions. It has been shown that the flow characteristics, such as the temperature
Dehumidifying condition
Fin efficiency
and humidity fields, along with the local heat flux and the condensation rate, can be successfully cap-
3D analysis tured. The numerical results reveal that the existing correlations considerably underestimate the fin effi-
ciency especially for multi-row heat exchangers. For dehumidifying cases, the sensible heat-transfer rate
seems insensitive to the inlet-humidity change. The rate changes mostly in the narrow band of partially
wet regime between 25% and 40% of inlet relative humidity. The range of the frontal velocity that gives
the best performance for various numbers of rows is also estimated. The analogy between the heat and
mass transfer on the fin surface is also examined.
Ó 2011 Elsevier Ltd. All rights reserved.

1. Introduction louvers, etc., however, one needs to have detailed local information
which normally is beyond the reach of experimental study. CFD
It is of practical importance to determine the fin efficiency accu- techniques may provide the accurate and detailed heat-transfer
rately in estimating the overall heat transfer of a heat exchanger. characteristics of heat exchangers. Tsai et al. [8], Perotin and Clodic
The solution generally is not known for various fin types and con- [9] and Tao et al. [10] carried out three-dimensional conjugate
figurations, and numerous mathematical approximations have heat-transfer analyses for dry condition and obtained the local
been suggested in the literature. Most such studies, however, are heat-transfer characteristics and/or the heat-transfer enhancement
based on one-dimensional analysis together with the assumption for various fin types. Few in the literature, however, provide a
that the fin-to-air heat-transfer coefficient is uniform. Improved methodology that takes the water condensation/evaporation into
methods have been suggested [1–7] by taking the variations of account. The rate of mass transfer may be estimated by the
the heat-transfer coefficient or the air temperature into consider- Chilton–Colburn [11] analogy assuming that the mass transfer
ation by approximating them as functions of the distance from can be decoupled from the heat transfer. Although this gives a
the tube. Haung and Shah [2] carried out comparative study of straight-forward way of predicting the mass-transfer rate from
these methods for a simple fin shape, but no clear conclusion as the known heat-transfer performance, it may not be appropriate
to which is most accurate. Besides, all these approaches show lim- as the condensation and the temperature are interdependent as
ited success as the heat-transfer coefficient is substantially larger seen in the psychrometric chart. Recently, Comini et al. [12]
near the leading edge of the fin compared to the rest of the region developed an analysis procedure, in which the cooling air is treated
and make it impossible to model by simple analytic functions. as two-component air–vapor mixture, and successfully estimated
Estimation of heat transfer under dehumidifying condition is the temperature and condensate distributions, and the
even more complex. Various effects of design parameters have heat-transfer coefficient for rectangular-finned heat exchangers.
been explored. To tackle the problems of practical interest such The objective of this study is to develop a fully coupled
as optimization of fin pitch, wave depth, positions of slits or three-dimensional heat/mass-transfer analysis procedure under
dry and wet conditions. The procedure is then applied to analyze
⇑ Corresponding author. Tel.: +82 42 350 3018; fax: +82 42 350 3210. the detailed heat/mass transfer characteristics of a wave fin of
E-mail address: d-h-choi@kaist.ac.kr (D.H. Choi).
complex shape and the multi-row heat-exchanger performance.

0017-9310/$ - see front matter Ó 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijheatmasstransfer.2011.10.055
C.S. An, D.H. Choi / International Journal of Heat and Mass Transfer 55 (2012) 1496–1504 1497

Nomenclature

A heat exchange area [m2] Qmax maximum heat-transfer rate [W]: ma(ia,in  ir,in)
Aa total surface area [m2] Rv gas constant of air [J/kg K]
Afr frontal area [m2] Re Reynolds number
Amin minimum free flow area [m2] T temperature [°C; K]
Cp specific heat [J/kg °C] t fin thickness [m]
D diffusion coefficient [m2/s] ui flow velocity in xi direction [m/s]
dh hydraulic diameter [m] vfr frontal velocity [m/s]
do outer tube diameter [m] w humidity [kg/kg air]
ha, hs air-side and sensible heat-transfer coefficients [W/m2 K]
hm mass-transfer coefficient Greek symbols
ia enthalpy of humid air [J/kg] gf fin efficiency
ir enthalpy of saturated water vapor at mean air tempera- l viscosity [N s/m2]
ture [J/kg] q density [kg/m3]
k thermal conductivity [W/m K] r standard deviation
L length of the fin [m]
N number of rows of heat exchanger Subscripts
p pressure [Pa] a air
PT, PL transversal and longitudinal tube pitch [m], (see Fig. 1) b fin base
qlm latent heat of water [J/kg] f fin
q00 heat flux [W/m2] s saturated state
Q overall heat-transfer rate [W] t tube
Qn heat-transfer rate of nth row [W] w fin surface

These numerical results may help evaluate the commonly invoked Energy:
data reduction practices: The fin efficiency is estimated and com- @ ^ ¼ 1 @ 2 T^
^j TÞ
ðu ð6Þ
pared with existing correlations based on various approximations. @ ^xj Re  Pr @ ^xj @ ^xj
The effects of inlet relative humidity, and the heat and mass trans-
fer analogy are also examined and/or assessed. Species:
2
@ ^ @ w
^j wÞ ¼ D
ðu ð7Þ
2. Equations and solution procedure @ ^xj @ ^xj @ ^xj
where the carets denote dimensionless variables and
The humid air can be taken as a mixture of air and water vapor.
Since the mass fraction of the water vapor is very low, it has little ef- q p T  Tt
q^ ¼ ; p^¼ ; T^ ¼ ;
fect on the air flow. Taking air as the carrier fluid and water vapor as a qin qin u2max T in  T t
dilute species, the laminar humid air flow can be described by the ð8Þ
q umax dh lC p ^ D
equations of continuity, momentum, energy, and species transport. Re ¼ in ; Pr ¼ ; D¼
l k umax dh
The effects of water film on the overall heat transfer or the thermal
resistance for the condensate of the droplet shape are known to be Here ui is the velocity component in the xi direction, p the pressure,
very small and thus are neglected in the present analysis. T the temperature, w the humidity, D the mass diffusivity, q the
Introducing the hydraulic diameter dh and the characteristic density, l the dynamic viscosity, k the thermal conductivity, and
velocity umax as the characteristic length and velocity scales: Cp the specific heat. The flow is incompressible, but the density
variation due to the temperature or the humidity is taken into
4Amin L
dh ¼ ð1Þ consideration as was done in O’Connell [13].
Aa
p
Afr q¼ P ð9Þ
umax ¼ v fr ð2Þ RT k ðm k =M k Þ
Amin
where mk and Mk are the mass fraction and the molecular weight of
with species k, respectively. The saturated humidity, which is identified
by the saturation line in the psychrometric chart, may be described
Amin ¼ MinðA1 ; 2A2 Þ ð3Þ
by the following formula:
where A1 and A2 are the areas between two adjacent tubes shown in ps ðT w Þ
Fig. 1 and vfr the frontal velocity, the dimensionless governing equa- ws ðT w Þ ¼ ð10Þ
qRv T w
tions can be written as
Continuity: The fin configuration of the heat exchanger is taken to be that of the
@ðq
^u ^j Þ experimental settings [14] and the boundary conditions for air and
¼0 ð4Þ
@ ^xj water temperature, and humidity are also matched to those of the
experiment. The schematic of the heat exchanger under consider-
Momentum:
ation together with its dimensions is shown in Fig. 2. It may be
@ @p ^ 1 @2u ^i
^i u
ðu ^j Þ ¼  þ ð5Þ assumed periodic in the transverse direction and it suffices to
@ ^xj @ ^xi Re @ ^xj @ ^xj consider only half of a unit module indicated in the figure. The
1498 C.S. An, D.H. Choi / International Journal of Heat and Mass Transfer 55 (2012) 1496–1504

PL

A2

Air flow PT A1
do

(a)

Fig. 2. Schematic of the heat exchanger and detailed fin shape: (a) heat exchanger,
(b) wave-type fin.
(b)
Fig. 1. Schematic of a fin-tube heat exchanger: (a) tube arrangement, (b) fin broken
q00latent ¼ qqlm Drw  n
^ ð13Þ
into sectors.
The humidity on the fin surface is equal to the saturation humidity,
calculation domain is chosen to include the fin and the flow passage ws, but assumes the value at the inlet if the local air has not reached
between the fins, and is extended to 6do upstream and 20do down- the saturation condition:
stream of the heat exchanger in the streamwise direction to capture ww ¼ min½ws ðT w Þ; win  ð14Þ
the entrance and exit regions properly. Using the second-order
upwind scheme for discretization of the convective terms, the gov- To properly account for the discontinuity in the heat flux between
erning Eqs. (4)–(7) are solved by the SIMPLE-type finite volume the air and fin sides as noted in Eq. (12), the air-side and fin-side do-
method, FLUENT, with the following boundary conditions: mains are solved successively by using each other’s results as
The velocity and the temperature along with the humidity are boundary conditions: Eqs. (11) and (14) are applied at the air-side
prescribed at the inlet boundary while the gradient in the flow interface while Eq. (12) is imposed on the fin side. These interface
direction is assumed negligible at the outlet boundary. No slip con- conditions are updated after each iteration and are determined as
dition on the solid surface and the coolant-tube temperature are part of the solution. The sensible heat exchanged may then be esti-
specified. The symmetry or periodic condition is imposed at the mated by integrating the heat flux into the air over the surface:
respective boundaries. Z 
@T 
At the fin–air interface, the temperature and the heat flux on Q sensible ¼ ka dA ð15Þ
Aa @nair
the fin and air sides should match, respectively, as:
where ka is the thermal conductivity of the air.
Tjfin ¼ Tjair ð11Þ The fin efficiency, defined as the ratio of actual heat-transfer
rate, Qactual, to the ideal maximum rate, Qideal, can be obtained.
 
@T  @T  Q actual
kf ¼ k þ q00latent ð12Þ
@nfin @nair
a
gf ¼ ð16Þ
Q ideal
where the latent heat flux q00latent is determined from the mass-trans- This concept is useful in data reduction and also in estimating the
fer rate on the surface: actual heat transfer from the more readily obtainable ideal heat
C.S. An, D.H. Choi / International Journal of Heat and Mass Transfer 55 (2012) 1496–1504 1499

transfer, in which the temperature variation in the fin is ignored. number due to the heat transfer in the heat exchanger amounts to
Since the fin material is uniform and the fin is sufficiently thin, 6.5% for the single-row case at the lowest frontal velocity. The Pra-
the air-side heat-transfer characteristic is the single most signifi- ndtl number variation from inlet to outlet is not as significant and
cant factor in the fin-tube heat exchangers. The exact solutions remains close to 0.72. The heat-transfer coefficient for a single
are known only for the limited fin types and configurations. It is module is first calculated by solving the fin–air conjugate heat-
thus, in general, obtained from a simple 1D analysis under the transfer problem. To account for the temperature variation of the
assumption that the heat-transfer coefficient for the fin surface is air and the coolant through the heat exchanger, the amount of heat
constant and the ambient temperature variation is negligible. For transfer for all such modules is estimated by using the tube-by-
fin-tube heat exchangers, Schmidt’s approximation [1], in which tube algorithm [20] with the fixed heat-transfer coefficient for
the polygon-shaped fin area is approximated by an equivalent cir- each row. The results are in remarkably good agreement with the
cular disk, is widely used. Improved results may be obtained by tak- measurement for all conditions. The slightly less satisfactory re-
ing account of the varying air temperature [2], fin thickness [3], sults for the three-row cases may be attributed to the shortcom-
non-uniform thermal conductivity [5,6], variable heat-transfer coef- ings inherent in the tube-by-tube method. The accuracy may be
ficients [7], etc. Threlkeld [15], McQuiston [16], and Wu and Bong improved by carrying out the three-dimensional calculation for
[17] proposed the fin efficiency under fully wet conditions. Wu each fin segment with varying coolant and air temperatures. No at-
and Bong also suggested a partially-wet 1D overall fin efficiency. tempt has been made to do this, though. The excellent agreement
Recently, Pirompugd et al. devised a reduction scheme for the mul- confirms that the procedure is capable of accurately predicting the
tiple row fin-and-tube heat exchangers by subdividing the region heat-transfer performance of the heat exchanger.
into many tiny segments and analyzed the heat/mass transfer char- In subsequent calculations, it is seen that the spatial density
acteristics under dehumidifying conditions [18,19]. The fin effi- variation of the air–vapor mixture due to the temperature change
ciency approximations under both dry and wet conditions are in the field is as large as ±4.2%, while those of viscosity and specific
compared with that obtained in the present study. heat turn out to be smaller than ±1%. The variation of the thermal
conductivity is about ±3%, but shows negligible effects on the re-
3. Results and discussion sults. This confirms that the present approach of treating the den-
sity as variable and keeping the transport properties constant is
Taking advantage of periodicity and symmetry as pointed out in appropriate. In the leading-edge region, the hydraulic and thermal
the previous section, only one half of the unit fin module is consid- boundary layers are developing and the heat transfer is very active.
ered. Approximately 200,000 cells are used to fit the domain of The local heat transfer in the vicinity of the leading edge becomes
each row as shown in Fig. 3. Additional 10,200 and 17,000 grid cells high and the fin temperature varies rapidly. This phenomenon is
are distributed for the upstream and downstream regions of the clearly visible in the temperature and heat-flux distributions for
fin, respectively. the cooled case at 1 m/s frontal velocity in Fig. 5. The leading-edge
The calculation is performed for the inlet air temperature of area is the most effective heat-transfer region of the fin. The flow
21 °C for the dry case and 27 °C with 43% humidity for the wet accelerates between the coolant tubes due to the narrowing flow
case, while the tube temperature is fixed at 45 °C and 6 °C for the passage; this results in the high heat flux on the fin surface there.
dry and wet cases, respectively. To validate the procedure de- In the certain region behind the tube, the air temperature becomes
scribed above, the normalized overall heat-transfer rate under lower than the adjacent fin temperature and the negative heat
dry and wet conditions for heat exchangers with various row transport is observed. This peculiar behavior may be explained as
numbers is compared with the experimental data of Youn [14] in follows: the air temperature behind the coolant tube is directly af-
Fig. 4. The Reynolds number based on the inlet air condition ranges fected by the low tube temperature while the fin surface below is
from 125 to 625 and 124 to 495 for dry and wet cases, respectively, kept relatively warm by the ambient air away from the spot. To see
for the frontal velocity range tested. The variation of the Reynolds the heat-transfer behavior more clearly, the heat- and mass-flux

Fig. 3. Sample grid for the periodic computational domain.


1500 C.S. An, D.H. Choi / International Journal of Heat and Mass Transfer 55 (2012) 1496–1504

layer, the fully saturated region in the first row decreases as the
flow velocity increases. For subsequent rows, however, the air is
sufficiently cooled to the point that the entire region becomes fully
saturated.
The condensation occurs as the air temperature dips below the
saturation temperature. The condensation in the leading edge re-
gion depends very much on whether the saturation condition is
reached there. Once past the developing region, the condensation
and the heat flux move in unison as seen in Fig. 6 for the cross-
section A1–A10 . The trend continues to the second and third rows.
The flow reaches the saturation condition earlier when the frontal
velocity is lower; the larger condensation rate follows.
One of the assumptions frequently adopted in the fin-efficiency
approximation is axisymmetry. This assumption usually does not
hold in real situations, as the fin temperature along three different
circular paths (L1, L2, L3) for the air velocity of 1 m/s in Fig. 9 exhib-
its. It is evident that the axisymmetry deteriorates as the circle gets
farther away from the tube: along the path close to the tube, L1,
the dimensionless temperature varies little, smaller than 0.03,
while the variation along L3 is 0.16. The flow appears to return
to symmetry about the tubes for the second and third rows as
the flow properties varies little in the streamwise direction. From
these results, coupled with the humidity variation in the flow
direction that needs to be accounted for in the analysis, one-
dimensional fin efficiency approximations or simple modified
models thereof that only take account of the geometric change in
the fin shape, may introduce significant error when applied to
the fins of complex shape.
The fin efficiency under both dry and wet conditions computed
in the present study is depicted in Fig. 10. The results of the exist-
ing fin-efficiency approximations are also shown in the figure for
comparison. The fin-efficiency curve is constructed by varying
the frontal velocity. Under dry conditions shown in Fig. 10(a), the
fin efficiency becomes higher as the number of rows increases. It
is because that the fin–air temperature difference becomes smaller
towards the downstream end and so does the temperature varia-
tion. This tends to make the actual heat transfer close to that of
the ideal heat-transfer rate and pushes the fin efficiency up close
to unity. For fins with large flow depth, the fin efficiency approxi-
mation underestimates the value up to 10–15%. The flow depth
needs to be small for the fin efficiency approximation to remain va-
lid. Among various approximations, the Han and Lefkowitz method
underestimates the results most. The heat-transfer coefficient is
very high near the leading edge of the fin and the fin efficiency is
Fig. 4. Overall heat-transfer rate for various heat exchangers: (a) dry condition, likely to be underestimated if it is based on the average heat-trans-
(b) wet condition. fer coefficient. The sector method and the methods of Schmidt, and
Huang and Shah agree quite well with one another but the values
are appreciably smaller than the present results. The fin tempera-
distributions along the A1–A10 cross-section for two different fron- ture variation is much less in the second and the third row than in
tal velocities are shown in Fig. 6. The heat flux is very high near the the first row, i.e., the fin efficiency depends less on the distance
leading edge in the developing region. It is natural to observe that a from the tube there. The close agreement of Huang and Shah,
higher frontal velocity gives a larger overall heat flux. The slightly which takes account of the variation of ambient air temperature
irregular behavior in the heat-flux distribution may be attributed in the transverse direction, with other methods seems to imply
to the crests or troughs of the wave fin that makes the flow that the effects of transverse mixing is insignificant.
accelerate or decelerate. The wall-shear-stress distribution in the For the fin efficiency under wet condition shown in Fig. 10(b),
symmetry planes, A1–A10 , A2–A20 , and A3–A30 in Fig. 7, clearly elu- the earlier studies [15–17] are again seen to give significantly low-
cidates this phenomenon. The shear stress exhibits the local max- er values. The coolant temperature in those studies is slightly dif-
ima or minima which coincide with the crests or the troughs of the ferent from that of the present, 7 °C vs. 6 °C. The discrepancy is
wave fin; the heat flux tends to move with the wall shear stress for attributed to the assumptions made in devising the approximation
obvious reasons. formulae. If the fully wet condition is met, i.e., the condensation
The procedure is capable of predicting the humidity field about occurs from the fin leading edge, it is reported that the fin effi-
the heat exchanger. The saturation surface moves gradually out- ciency varies only slightly with the relative humidity at the inlet
ward from the fin surface as the air gets cooler as it travels down- [17–19]. It is seen in the figure that, when the inlet humidity
stream. Fig. 8 shows the development of the saturation region in changes from 60% to 100%, the fin efficiency by Wu and Bong
the symmetry plane for various frontal velocities, i.e., 0.5, 1.0, [17] results in the difference of up to 6% while that in the present
and 2.0 m/s. Similar to the behavior of the hydraulic boundary study 2.8%. For multi-row heat exchangers, the difference becomes
C.S. An, D.H. Choi / International Journal of Heat and Mass Transfer 55 (2012) 1496–1504 1501

^ (b) heat-transfer rate q00


Fig. 5. Temperature and local heat-transfer rate for vfr = 1.0 m/s: (a) temperature T, .
q00av

smaller as hs increases, i.e., with increasing frontal velocity, while of the pressure drop through the heat exchanger DP and the frontal
the opposite is true for the single-row heat exchanger. The insen- velocity vfr:
sitive behavior of fin efficiency to the inlet relative humidity is
due to the fact that the flow is in near fully wet condition. To see ðdriving powerÞ
¼ DP  v fr ½W=m2  ð17Þ
how the fin efficiency varies from the dry to wet condition, calcu- ðunit face areaÞ
lations have been carried out with varying inlet relative humidity
and the results are plotted in Fig. 11. This is for the single-row case Naturally, for both dry and wet conditions, a single-row heat ex-
with three different inlet velocities. The fin efficiency remains changer is superior to the multi-row heat exchangers when the
fairly constant when the condition is either dry or fully wet. It var- power is low. As the power increases, however, the performance
ies mostly in the narrow region of partially wet regime between of the two- or three-row heat exchangers surpasses that of the
25% and 40% of inlet relative humidity. This band of partially wet single-row heat exchanger at point A indicated in the figure. Even-
regime becomes narrower as the frontal velocity decreases. Similar tually, the performance of the three-row heat exchanger comes out
behavior, but with narrower partially wet regime, is also observed to be on top at relatively high power beyond point B. To find the
for two- and three-row heat exchangers. optimal range of operation for a given number of rows, the variation
For heat exchangers with multiple rows, the total heat-transfer in the tube temperature must also be considered using, for example,
rate increases with the number of rows. Unfortunately, that also the tube-by-tube method. This is not pursued in the present study
accompanies higher pressure loss. From the perspective of wall- to keep the analysis simple.
shear stress depicted in Fig. 7, the shear stress becomes periodic As one of the many design considerations, it may be useful to
immediately after the leading-edge region of the first row. The point out that, at the crossover point, B in Fig. 12(b), for two-
pressure drop for the heat exchanger, therefore, is expected to in- and three-row heat exchangers under the wet condition, the
crease almost linearly with the number of rows. The performance frontal velocities are 1.32 and 1.14 m/s, respectively. Similarly,
of each row for various conditions is summarized in Table 1. In fact, for single- and two-row exchangers, the frontal velocities at the
the heat-transfer rate for the second or third row is much smaller crossover point are 0.63 and 0.50 m/s. These numerical results
than that of the first row. The lower the frontal velocity, the less may be vital in designing the heat exchangers.
effective the heat-transfer rate of the second and third rows. The effect of inlet humidity on the sensible heat transfer is pre-
To assess the performance of the multi-row heat exchangers, sented in Fig. 13. The temperature change between the inlet and
the heat-transfer rate is plotted against the power consumed in the outlet relative to that of the dry condition is plotted against
Fig. 12. The driving power per unit face area is given by the product the frontal velocity in the figure for two different inlet-humidity
1502 C.S. An, D.H. Choi / International Journal of Heat and Mass Transfer 55 (2012) 1496–1504

(a) (a)

(b)

(b)

Fig. 6. Heat-flux and condensation-rate distributions along A1–A10 section for two
different frontal velocities: (a) vfr = 0.5 m/s, (b) vfr = 2.0 m/s.

conditions. This compares the sensible heat-transfer drop from


that of the dry condition for various numbers of rows. The drop
is most pronounced, above 20%, for the single row heat exchanger
at the highest frontal velocity. This may be elucidated as follows: Fig. 7. Shear-stress distribution on the fin surface along the symmetry planes for
the air is most humid in the first row and thus the latent heat two different frontal velocities: (a) vfr = 0.5 m/s, (b) vfr = 2.0 m/s.
transfer becomes largest. This results in the smallest sensible heat
transfer there. Increasing frontal velocity makes the moisture flux
in the air stream larger and also reduces the sensible heat transfer.
Naturally the drop becomes less significant as the frontal velocity
decreases and/or the number of rows increases. This observation
is in line with the experimental study of Wang et al. [21] but the
present analytical procedure provides much finer details as to
how the performance is affected.
Let us now examine the analogy between the heat- and mass-
transfer on the fin surface. The following Pearson product–moment
correlation coefficient Cor(hs, hm), which lies between 0 and 1 with
1 indicating the perfect dependence, is useful in quantifying the
degree of linear dependence of the two parameters.
E½ðhs  Eðhs ÞÞðhm  Eðhm ÞÞ
Corðhs ; hm Þ ¼ ð18Þ
rH rM
where E denotes the area average of the argument over the fin sur-
face and rH, rM the standard deviation of hs, hm, respectively:
h n oi1=2
rH ¼ E ðhs  Eðhs ÞÞ2 ð19Þ Fig. 8. Saturation lines in the A1–A10 section of the first row for various frontal
velocities.
C.S. An, D.H. Choi / International Journal of Heat and Mass Transfer 55 (2012) 1496–1504 1503

Fig. 9. Fin surface temperature along various circular paths.

Fig. 11. Effects of inlet relative humidity on fin efficiency for various inlet velocities.

(a)
Table 1
Heat-transfer rates for the second and third rows relative to the first row.

vfr [m/s] Dry Wet


Q2/Q1 Q3/Q1 Q2/Q1 Q3/Q1
0.50 0.1444 0.02133 0.1785 0.03186
1.00 0.3462 0.1104 0.3803 0.1364
2.00 0.5752 0.2517 0.6125 0.2901

row. In the leading-edge region, the condensation rate is relatively


low compared to the heat flux as seen in Fig. 6. This is presumably
due to the dissimilarity of the developing profiles of temperature
and humidity and/or the fact that the flow is not fully saturated
in the leading-edge region. This tendency is well supported by the
significantly small value of the correlation for the first row in the ta-
ble. If the upstream 25% of the first row is treated separately, the
(b) correlation for this region is about 0.502, for the higher frontal
velocity case, whereas that for the remaining 75% is 0.968, higher
than that of the second row. One may deduce from this result that
for fins with louvers or slits, as they have much larger overall devel-
oping area compared to the wave fins, the heat and mass transfer
analogy would be less accurate.

4. Conclusions

A numerical procedure that predicts the overall heat-transfer


rate of the heat exchanger has been developed. The three-dimen-
sional continuity, momentum, energy, and species transport equa-
tions coupled with the condensation model at the interface are
solved by using the finite volume method of SIMPLE type to obtain
the heat-transfer rate of multi-row cross-flow fin-tube heat
Fig. 10. Fin efficiency against the heat-transfer rate: (a) dry condition, (b) wet
exchangers under both dry and wet conditions. The details of the
condition. heat-transfer characteristics including the local condensation rate
are seen to be accurately captured. The optimal range of frontal
h n oi1=2 velocity is also identified for one-, two-, and three-row heat
rM ¼ E ðhm  Eðhm ÞÞ2 ð20Þ exchangers.
The fin efficiency concept is found to underestimate the heat-
transfer rate noticeably for all cases. This is attributed to the sim-
The results are presented in Table 2. It is shown that the two are plified assumptions adopted in approximating the fin efficiency.
more strongly related in the second and third rows than in the first Although the results may be improved by making modifications
1504 C.S. An, D.H. Choi / International Journal of Heat and Mass Transfer 55 (2012) 1496–1504

Table 2
(a) Pearson product–moment correlation factor between the local heat- and mass-
transfer rates.

vfr [m/s] Row Correlation factor


0.5 First 0.8321
Second 0.9384
Third 0.9138
2.0 First 0.7513
Second 0.9536
Third 0.9451

to the formulae, the improvement is limited and leaves much to be


desired in accurately estimating the fin efficiency, especially, for
multi-row heat exchangers. For dehumidifying cases, the overall
sensible heat-transfer rate seems insensitive to the inlet humidity
change if the fully wet condition is met. The fin efficiency varies
mostly in the narrow region of partially wet regime between 25%
and 40% of inlet relative humidity.
(b) A quantitative assessment of the heat- and mass-transfer anal-
ogy by using the product–moment correlation shows that the anal-
ogy holds almost perfectly for most of the fin surface except the
small developing entrance region. The analogy may still be useful
and accurate when the inlet air is in near saturation state.

References

[1] T.E. Schmidt, Heat transfer calculations for extended surfaces, Refrig. Eng. 49
(1949) 351–357.
[2] L.J. Huang, R.K. Shah, Assessment of calculation methods for efficiency of
straight fins of rectangular profile, Int. J. Heat Fluid Flow 13 (1992) 282–293.
[3] W. Lau, C.W. Tan, Errors in the one-dimensional heat-transfer analysis in
straight and annular fins, J. Heat Transfer 95 (1973) 549–551.
[4] J.B. Aparecido, R.M. Cotta, Improved one-dimensional fin solutions, Heat Trans.
Eng. 11 (1990) 49–59.
[5] H.M. Hung, F.C. Appl, Heat transfer of thin fins with temperature dependent
thermal properties and internal heat generation, J. Heat Transfer 89 (1967)
155–162.
[6] H. Barrow, J. Mistry, D. Clayton, Numerical and exact mathematical analyses of
two-dimensional rectangular composite fins, Heat Transfer 2 (1986) 367–372.
[7] L.S. Han, S.G. Lefkowitz, Constant cross-section fin efficiencies for nonuniform
surface heat-transfer coefficients, ASME Paper 60-WA-41, 1960.
Fig. 12. Heat-transfer rate vs. power consumption for various numbers of rows: [8] S.F. Tasi, T.W.H. Sheu, S.M. Lee, Heat transfer in a conjugate heat exchanger
(a) dry condition, (b) wet condition. with a wavy fin surface, Int. J. Heat Mass Transfer 42 (1999) 1735–1745.
[9] T. Perrotin, D. Clodic, Thermal-hydraulic CFD study in louvered fin-and-flat-
tube heat exchangers, Int. J. Refrig. 27 (2004) 422–432.
[10] Y.B. Tao, Y.L. He, J. Huang, Z.G. Wu, W.Q. Tao, Numerical study of local heat
1 transfer coefficient and fin efficiency of wavy fin-and-tube heat exchangers,
Int. J. Thermal Sci. 46 (2007) 768–778.
[11] T.H. Chilton, A.P. Colburn, Mass transfer (absorption) coefficients-prediction
0.95 from data on heat transfer and fluid function, Ind. Eng. Chem. 26 (1934) 1183–
1187.
[12] G. Comini, C. Nonino, S. Savino, Numerical evaluation of fin performance under
(ΔTin-out )wet / (ΔTin-out )dry

dehumidifying conditions, J. Heat Transfer 129 (2007) 1395–1402.


0.9 [13] J.P. O’Connell, J.M. Prausnitz, B.E. Poling, Properties of Gases and Liquids, fifth
ed., McGraw-Hill, 2001.
[14] B. Youn, Private Communication, Samsung Electronics Co., 2004.
[15] J.L. Threlkeld, Thermal Environmental Engineering, Prentice-Hall, Inc., New
0.85 York, 1970.
[16] F.C. McQuiston, Fin efficiency with combined heat and mass transfer, ASHRAE
Trans. 81 (1975) 350–355.
[17] G. Wu, T.-Y. Bong, Overall efficiency of a straight fin with combined heat and
0.8 mass transfer, ASHRAE Trans. 100 (1994) 367–374.
[18] W. Pirompugd, S. Wongwises, C.-C. Wang, Simultaneous heat and mass
RH = 60 % N=1 transfer characteristics for wavy fin-and-tube heat exchangers under
0.75 N=2 dehumidifying conditions, Int. J. Heat Mass Transfer 49 (2009) 132–143.
RH = 100 % N=3 [19] W. Pirompugd, C.C. Wang, S. Wongwises, A fully wet and fully dry tiny circular
fin method for heat and mass transfer characteristics for plain fin-and-tube
heat exchangers under dehumidifying conditions, J. Heat Transfer 129 (2006)
0.7 1256–1267.
0.5 1 1.5 2 [20] P.A. Domenski, Computer modeling and prediction of performance of an air
source heat pump with a capillary tube, Ph.D. Dissertation, The Catholic
vfr [m/s] University of America, Washington D.C., 1989, 62–93.
[21] C.-C. Wang, Y.-T. Lin, C.-J. Lee, Heat and momentum transfer for compact
Fig. 13. Sensible heat transfer relative to the dry case for two inlet-humidity louvered fin-and-tube heat exchangers in wet conditions, Int. J. Heat Mass
conditions and varying row numbers. Transfer 43 (2000) 3443–3452.

S-ar putea să vă placă și