You are on page 1of 13

Heat Mass Transfer (2008) 44:11271139

DOI 10.1007/s00231-007-0346-4

ORIGINAL

Air-side performance of louver-finned flat aluminum heat


exchangers at a low velocity region
Nae-Hyun Kim Jin-Pyo Cho

Received: 19 May 2007 / Accepted: 12 September 2007 / Published online: 25 October 2007
 Springer-Verlag 2007

Abstract The heat transfer and pressure drop characteristics of heat exchangers having louver fins were
experimentally investigated. The samples had small fin
pitches (1.01.4 mm), and experiments were conducted up
to a very low frontal air velocity (as low as 0.3 m/s). Below
a certain Reynolds number (critical Reynolds number), the
fall-off of the heat transfer coefficient curve was observed.
The critical Reynolds number was insensitive to the louver
angle, and decreased as the louver pitch to fin pitch ratio
(Lp/Fp) decreased. Existing correlations on the critical
Reynolds number did not adequately predict the data. The
heat transfer coefficient curves crossed over as the Reynolds number decreased. Possible explanation is provided
considering the louver pattern between neighboring rows.
Different from the heat transfer coefficient, the friction
factor did not show the fall-off characteristic. The reason
was attributed to the form drag by louvers, which offsets
the decreased skin friction at low Reynolds numbers. The
friction factor increased as the fin pitch decreased and the
louver angle increased. A new correlation predicted 92% of
the heat transfer coefficient and 94% of the friction factor
within 10%.

List of symbols
A
heat transfer area (m2)
cp
specific heat (J /kg s)
Cr
heat capacity ratio, dimensionless [Eq. (6)]
D
ideal transverse distance (m)

N.-H. Kim (&)  J.-P. Cho


Department of Mechanical Engineering, University of Inchon,
#177 Dohwa-Dong, Nam-Gu, Inchon 402-749, South Korea
e-mail: knh0001@incheon.ac.kr

Dh
FD
Fe
Fp
f
fi
H
h
j
k
Ll
Lp
m_
N
NTU
Pf
Pr
Q
ReLp
ReLp
ReDh
S1
S2
t
T
Tp
tf
U
Vmax

hydraulic diameter (m)


depth of fin array in flow direction (m)
flow efficiency, dimensionless [Eq. (18)]
fin pitch (m)
airside friction factor, dimensionless [Eq. (15)]
tube-side friction factor, dimensionless [Eq. (10)]
fin height (m)
heat transfer coefficient (W/m2 K)
Colburn j factor, dimensionless [Eq. (14)]
thermal conductivity (W/m K)
louver length (m)
louver pitch (m)
mass flow rate (kg /s)
actual transverse distance (m)
number of transfer units, dimensionless [Eq. (7)]
fin pitch (m)
Prandtl number, dimensionless
Heat transfer rate (W)
Reynolds

 number based on Lp, dimensionless
Vmax Lp
m
critical Reynolds number, dimensionless
tube-side Reynolds number based on
Dh, dimensionless
Non-louvered inlet and exit fin length (m)
Re-direction louver length (m)
tube wall thickness (m)
temperature (K)
tube pitch (m)
fin thickness (m)
overall heat transfer coefficient (W/m2 s)
maximum airside velocity (m/s)

Greek symbols
a
louver angle (degree)
b
flow angle (degree)

123

1128

e
DP
g
go
q
m
r

Heat Mass Transfer (2008) 44:11271139

thermal effectiveness, dimensionless [Eq. (5)]


pressure loss (Pa)
fin efficiency, dimensionless [Eq. (12)]
surface efficiency, dimensionless [Eq. (11)]
density (kg/m3)
kinematic viscosity (m2/s)
contraction ratio of the cross-sectional area,
dimensionless

Subscripts
c
heat exchanger core
exp experimental
i
tube-side
in
inlet
f
fin
m
mean
max maximum
min minimum
o
airside
out
outlet
pred predicted
t
tube

1 Introduction
Fin-and-tube heat exchangers have been widely used as
condensers or evaporators in a household air-conditioning
system. In the forced convective heat transfer between air
and refrigerant, the controlling thermal resistance is on the
air-side. To improve the air-side performance, rigorous
efforts have been made, which include a usage of high
performance fins and of small diameter tubes, etc. However, fin-and-tube heat exchangers have inherent shortcomings such as the contact resistance between fins and
tubes, the existence of a low performance region behind
tubes, etc. These short-comings may be overcome if fins
and tubes are soldered, and low profile flat tubes with high
performance fins are used. Brazed aluminum flat-tube heat
exchangers with louver fins could be the choice, although
the price could be higher due to additional brazing process.
Flat tube heat exchangers have been used as condensers of
automotive air conditioning units for more than ten years,
and they are replacing fin-and-tube condensers of the residential air-conditioning units. The possibility of replacing
the residential fin-and-tube heat exchangers by flat tube
heat exchangers has been studied by Webb and Jung [1].
They showed that, for the same air-side thermal capacity,
the flat-tube geometry requires less than half the heat
exchanger volume compared with the fin-and-tube counterpart. The advantage of flat-tube heat exchangers has
further been studied by Webb and Lee [2]. They compared
the thermal performance of flat tube condenser having

123

866 fins per meter with that of the fin-and-tube condenser


having 7.0 mm round tubes and 1,024 fins per meter. The
flat tube condenser was shown to reduce the material up to
50%. One major challenge for the application of the flat
tube condenser to the residential air-conditioner is the
reduced frontal air velocity. The frontal air velocity of the
residential system is much lower (0.51.5 m/s) than that of
the automotive system (25 m/s), where most of the flat
tube heat exchanger technology is based. The thermal
performance of the louver fin is known to significantly
decrease as the flow velocity decreases.
Although louvered surfaces have been used since the
1950s, the flow characteristics have not been unveiled until
the pioneering work by Davernport [3]. He showed that,
through flow visualization study, the flow did not pass
through the louvers at low Reynolds numbers. At high
Reynolds numbers, however, the flow became nearly parallel to the louvers. He speculated that, at low air velocities,
the developing boundary layers on adjacent louvers
became thick enough to effectively block the passage,
resulting in nearly axial flow through the array. Achaichia
and Cowell [4] further confirmed that, through heat transfer
tests on flat tube heat exchangers having louvered plate
fins, the Stanton numbers approached those of the duct flow
at sufficiently low Reynolds numbers. At high Reynolds
numbers, the data were parallel to those of the laminar
boundary layer for a flat plate. Two types of flow were
identified within the louvered plate fin arrayduct
directed flow and louver directed flow. The amount of
either flow depended on the louver geometry such as fin
pitch, louver pitch, louver angle as well as Reynolds
number. Webb and Trauger [5] conducted flow visualization experiments using large scale models and introduced
the flow efficiency, which implies the ratio of the louver
directed flow to the total flow. The flow efficiency
increased as the Reynolds number increased or the fin pitch
decreased. The role of vortex shedding in louver geometry
has been investigated by DeJong and Jacobi [6]. The
shedding started at the last downstream louver, and moved
upstream as the Reynolds number increased. The critical
Reynolds number, where the vortex shedding started,
decreased as the louver angle increased and the fin pitch
decreased.
Significant efforts have also been devoted to numerically analysis to obtain detailed flow structures over the
louver array. Achaichia and Cowell [7] modeled one louver
in the fully developed region assuming periodic boundary
conditions. Their results confirmed the existence of duct
directed flow at low Reynolds numbers and louver
directed flow at high Reynolds numbers. Atkinson et al. [8]
conducted time-dependent two- and three-dimensional
numerical study. The results showed vortex shedding from
the trailing edges. Three-dimensional model predicted the

Heat Mass Transfer (2008) 44:11271139

1129

data better than the two-dimensional model. Tafti et al. [9]


investigated the flow transition in a louver fin geometry by
conducting time-dependent two-dimensional analysis. The
initial instability was observed at the exit louver, and
moved upstream as the Reynolds number increased. The
findings were in close agreement with the experimental
results by DeJong and Jacobi [6].
Following the pioneering study by Davernport [10], many
investigations have been made on the air-side heat transfer
and pressure drop characteristics of louver finflat tube heat
exchangers. The previous experimental studies are summarized in Table 1. Davernport [10] provides data on 32
one-row samples, where louver dimensions were systematically varied (0.94 Lp/Pf 2.24, 8 a 36) for two
louver lengths (12.7 and 7.8 mm). The Reynolds number
ranged 300 ReLp 4,000. However, it should be noted that
the fin corrugation pattern of Davernport [10] is not common. Their fin geometry had adjacent fins inclined at a
significantly larger angle to each other (Z-shape) than is the
case with the more common pattern (U or V-shape). Davernport [11] provides j and f correlations based on the data.
Sunden and Svantesson [12] tested 6 one-row samples
having 0.26 Lp/Pf 0.91, 14 a 34 for 100 ReLp 700,
and proposed j and f correlations based on their own data.
Webb and Jung [1] provides data for 6 one-row samples
having 0.48 Lp/Pf 1.0, a = 30 for 100 ReLp 2,000.
Chang and Wang [13] tested 27 samples having 0.60 Lp/Pf
0.85, a = 28 for 100 ReLp 800, and reported that
Shanoun and Webb [14] analytic model predicted the heat
transfer and friction data successfully. Significant amount of
data have been provided by Kim and Bullard [15] (45
samples, 0.26 Lp/Pf 0.91, 14 a 34) for 100 ReLp
500. The heat transfer coefficient increased as the louver
angle increased. However, above a critical louver angle, the
heat transfer coefficient decreased with the increase of the
louver angle. The friction factor continuously increased with
the increase of the louver angle. Multiple regression correlations were proposed based on their own data. Achaichia
and Cowell [4] tested 15 samples having 0.24 Lp/Pf 0.85,
22 a 30 for 30 ReLp 1,000, and proposed j and f

correlations from their own data. Achaichia and Cowells


sample had louvered plate fins on flat tubes, different from
the traditional flat tube geometry, which has folded louver
fins between flat tubes.
The literature survey reveals that most of the studies
have been conducted for the Reynolds number (based on
louver pitch) larger than 100. Only Achaichia and Cowells
[4] samples were tested up to Reynolds number of 30. For a
household air-conditioner, the frontal air velocity is
between 0.5 and 1.5 m/s with corresponding Reynolds
number between 60 and 200 (for 1.7 mm louver pitch). It
has been revealed by many investigators that, as the Reynolds number decreases, more flow tends to be ductoriented, which significantly decreases the heat transfer
coefficient. This situation may be remedied if the fin pitch
is decreased or louver pitch is increased. Smaller the fin
pinch (or lager the louver pitch), more flow is expected to
follow the louver. Table 1 reveals that most of the previous
studies have been conducted at small values of Lp/Pf. The
Lp/Pf of Achaichia and Cowell [4] is less than 1.0. The
present samples have small fin pitches (1.01.4 mm) and
large louver length (1.7 mm) with corresponding Lp/Pf
from 1.21 to 1.70. In addition, in the present study,
experiments were conducted at frontal air velocities as low
as 0.3 m/s (with the corresponding Reynolds number as
small as 40). The effect of fin pitch and louver angle has
also been investigated.

2 Experiments
2.1 Heat exchanger samples
A total of 12 heat exchangers having three different fin
pitches (1.0, 1.2, 1.4 mm) and four different louver angles
(15, 19, 25, 27) were tested. The louver pitch was fixed
at 1.7 mm. The samples consisted of 24 steps of louver fins
brazed to flat tubes as illustrated in Fig. 1. The height and
width of the samples were 254 and 400 mm, respectively.
The tube-side was circuited in a serpentine fashion with

Table 1 Summary of previous studies on the air-side performance of flat-tube heat exchangers having louver fins
Investigators

LP (mm)

Fp (mm)

a (degree)

LP/Fp

ReLp

Davernport [10]

1.53.0

1.01.6

836

0.942.24

3004,000

Achaichia and Cowell [4]

0.81.4

1.73.3

2230

0.240.85

301,000

Sunden and Svantesson [12]

0.81.5

1.52.0

1434

0.260.91

100700

Webb and Jung [1]

1.01.4

1.42.1

30

0.481.0

1002,000

Chang and Wang [13]

1.31.9

1.82.2

28

0.600.85

100800

Kim and Bullard [15]

1.7

1.01.4

1529

1.211.70

100500

This study

1.7

1.01.4

1527

1.211.70

301,000

123

1130

Heat Mass Transfer (2008) 44:11271139

2.2 Test apparatus and procedures

Fig. 1 Schematic drawing of the test sample

two tubes per pass. With this circuitry, tube-side flow was
maintained turbulent. Maintaining turbulent flow in the
tube-side is important because the tube-side thermal
resistance needs to be minimized for an accurate assessment of the airside heat transfer coefficient. In addition to
this, tube-side flow mal-distribution problem, which might
exist for a multiple tube configuration, was eliminated.
Dimensional details of the flat tube and the louver fin are
provided in Fig. 2 and Table 2.

20.0
1.4

R0.6

0.4
R1

2.0

0.5
Dh: 1.307mm

A
Ll H
TP

FD
LP

A schematic drawing of the apparatus is shown in Fig. 3.


It consists of a suction-type wind tunnel, water circulation
and control units, and a data acquisition system. The
apparatus is situated in a constant temperature and
humidity chamber. The airside inlet condition of the heat
exchanger is maintained by controlling the chamber
temperature and humidity. The inlet and outlet dry and
wet bulb temperatures are measured by the sampling
method as suggested in ASHRAE Standard 41.1 [16]. A
diffusion baffle is installed behind the test sample to mix
the outlet air. The waterside inlet condition is maintained
by regulating the flow rate and the outlet temperature of
the constant temperature bath situated outside of the
chamber. Both the air and the water temperatures are
measured by pre-calibrated RTDs (Pt-100X sensors).
Their accuracies are 0.1 K. The water flow rate is
measured by a mass flow meter, whose accuracy is
0.0015 l/s. The airside pressure drop across the heat
exchanger is measured using a differential pressure
transducer. The air flow rate is measured using a nozzle
pressure difference according to ASHRAE Standard 41.2
[17]. The accuracy of the differential pressure transducers
is 1.0 Pa. The wind tunnel is equipped with multiple
nozzles, and an appropriate one is selected depending on
the air velocity.
During the experiment, the water inlet temperature
was held at 45C. The chamber temperature was maintained at 21C with 60% relative humidity. Experiments
were conducted varying the frontal air velocity from 0.3
to 3.5 m/s. The energy balance between the airside and
the tube-side was within 2% for the air velocity larger
than 1.0 m/s. It increased to 5% at the air velocity of
0.3 m/s. All the data signals were collected and converted by a data acquisition system (a hybrid recorder).
The data were then transmitted to a personal computer
for further manipulation. An uncertainty analysis was
conducted following ASHRAE Standard 41.5 [18], and
the results are listed in Table 3. The major uncertainty on
the friction factor was the uncertainty of the differential
pressure measurement (10%), and the major uncertainty on the heat transfer coefficient (or j factor) was
that of the tube-side heat transfer coefficient (10%).
The uncertainties decreased as the Reynolds number
increased.

FP

2.3 Data reduction


S1

S2

Fig. 2 Geometric dimensions of the test sample

123

The total heat transfer rate used for the calculation of airside heat transfer coefficient was obtained from the
mathematical average of Qo and Qi.

Heat Mass Transfer (2008) 44:11271139

1131

Table 2 Geometric dimensions of test samples


Samples

FD (mm)

a (degree)

FP (mm)

LP (mm)

Ll (mm)

H (mm)

TP (mm)

S1 (mm)

S2 (mm)

20

15

1.0

1.7

6.4

8.15

2.0

1.82

1.0

20

19

1.0

1.7

6.4

8.15

2.0

1.82

1.0

20

25

1.0

1.7

6.4

8.15

2.0

1.82

1.0

20

27

1.0

1.7

6.4

8.15

2.0

1.82

1.0

20

15

1.2

1.7

6.4

8.15

2.0

1.82

1.0

20

19

1.2

1.7

6.4

8.15

2.0

1.82

1.0

20

25

1.2

1.7

6.4

8.15

2.0

1.82

1.0

20

27

1.2

1.7

6.4

8.15

2.0

1.82

1.0

20

15

1.4

1.7

6.4

8.15

2.0

1.82

1.0

10

20

19

1.4

1.7

6.4

8.15

2.0

1.82

1.0

11
12

20
20

25
27

1.4
1.4

1.7
1.7

6.4
6.4

8.15
8.15

2.0
2.0

1.82
1.82

1.0
1.0

Fig. 3 Schematic drawing


of the apparatus

123

1132

Heat Mass Transfer (2008) 44:11271139

2
fi 1:58 lnReDh;i  3:28 :

Table 3 Experimental uncertainties


Parameter

Max. Uncertainties

Temperature

0.1 K

Differential pressure

1 Pa

Water flow rate

1.5 106 m3/s

ReDc

2%

10%

12%

Q Qo Qi =2

where Qo and Qi are heat transfer rates of air and water


sides, respectively.
Qo m_ o cp;o To;out  To;in

Qi m_ i cp;i Ti;in  Ti;out :

The UA value was obtained from the effectiveness and


NTU method assuming unmixed-unmixed cross flow [19].

e 1  exp


NTU0:22 
expCr NTU 0:78  1
Cr


4

where
e Q=Qmax

_ p min =mc
_ p max :
Cr mc

The UA value was obtained from the following


equation.
_ p min NTU:
UA mc

1
1
1
t



go ho Ao UA hi Ai kt At

where go is the overall surface effectiveness of the louver


fin, t is the thickness of the tube wall, and the subscripts
o and i stand for the air and tube sides, respectively.
Note that the tube-side heat transfer area, Ai includes the
internal web surfaces. The tube-side heat transfer
coefficient, hi was evaluated from the Gnielinski [20]
semi-empirical correlation.

hi


ki
ReDh;i  1; 000Pri fi =2
p
Dh;i 1:0 12:7 fi =2Pr 2=3  1
i

where

123

During the experiment, the tube-side Reynolds number


was maintained at 6,500, which was the maximum value
obtainable from the present experimental set-up. Due to the
small hydraulic diameter of the flat tube, it was very hard to
increase the tube-side Reynolds number within permissible
pressure loss. At the tube-side Reynolds number 6,500, the
tube-side thermal resistance was within 5% of the total
thermal resistance.
The surface efficiency go was obtained from Eq. (11).

go 1 

Af
1  g:
Ao

11

The fin efficiency is given by Schmidt [21] as

tan hml
ml

12

where
s


2ho
tf
m
1
kf t f
FD :

13

H
l  tf
2
The heat transfer coefficient is traditionally presented as
the Colburn j factor.

ReLp

The airside heat transfer coefficient ho was then


calculated by subtracting the water-side and wall
resistances from the total thermal resistance.

10

Vmax Lp
m
ho

qo Vmax cpo

Pr 2=3
o

14
15

where Vmax is the maximum velocity in the core of the heat


exchanger. The Vmax is obtained when all the flow passes
through louvers. All the fluid properties were evaluated at
an average air temperature. The core friction factor was
calculated from the measured pressure drop.
"


Ac qm 2DPqin
qin
2
f

K

1

r


2

1
c
Ao qin qm Vmax 2
qout

q
1  r2  Ke in :
16
qout
In Eq. (16), Kc and Ke are coefficients for pressure loss
at the inlet and outlet of the heat exchangers, and were
evaluated at ReDh = ? from Figs. 4, 5 of Kays and London
[22].

Heat Mass Transfer (2008) 44:11271139

1133

= 15

= 19

1.0 mm
1.2 mm

1.0 mm
1.2 mm
1.4 mm

0.1

0.1

0.01
30

100

1000

0.01
30

2000

100

1000

2000

Re Lp

Re Lp

a) = 15

b) = 19

1
o

= 25

= 27

1.0 mm
1.2 mm
1.4 mm

1.0 mm
1.2 mm
1.4 mm

0.1

0.1

0.01
30

100

1000

2000

0.01
30

100

1000

Re Lp

Re Lp

c) = 25

d) = 27

2000

Fig. 4 The j and f factors of present samples showing the effect of fin pitch
Fig. 5 Illustration of the flow
efficiency

FLOW

IDEAL STREAMLINE
ACTUAL STREAMLINE

3 Results and discussions


Figure 4 shows the heat transfer coefficients and friction
factors for four different louver angles. Each graph contains data for three different fin pitches except for 15
louver angle where fin pitch 1.4 mm data are missing. The
sample was spoiled during preparation. Figure 4 shows two
distinct regimes. At high Reynolds numbers, the j factor
increases as the Reynolds number decreases. At low
Reynolds numbers, however, it decreases as the Reynolds

number decreases. This fall-off of the j factor has also been


reported by Achaichia and Cowell [4, 7]. As noted by many
investigators [3, 4, 5], more flow tends to follow the duct as
the Reynolds number decreases, which results in the
decrease of the heat transfer coefficient. Davernport [3]
conjectured that, at low Reynolds numbers, the laminar
boundary layer on louvers becomes sufficiently thick to
effectively block off the gaps between adjacent louvers.
The critical Reynolds numbers at which the fall-off of j
factor curve occurs are listed in Table 4. The critical

123

1134

Heat Mass Transfer (2008) 44:11271139

Table 4 The critical Reynolds numbers predicted by Cowell et al. [22] and Webb and Trauger[5]
Investigators

Fp = 1.0 mm (Lp/Fp = 1.70)


15

19

25

Fp = 1.2 mm (Lp/Fp = 1.42)


27

15

19

25

Fp = 1.4 mm (Lp/Fp = 1.21)


27

15

19

25

27

Present heat transfer test

%140

%140

%140

%140

%100

%110

%130

%130

%70

%70

%70

%70

Cowell et al. [22]

328

258

196

182

332

261

198

183

336

264

200

184

Webb and Trauger [5]

1,522

1,405

1,280

1,247

1,522

1,405

1,280

1,247

1,522

1,405

1,280

1,247

Reynolds number has also been investigated by Cowell


et al. [23]. They defined the critical Reynolds number as
the one obtained when the flow angle (b) has dropped to
95% of its maximum value (bmax). The bmax is the limiting
value of b at high Reynolds numbers, and was obtained
from their previous numerical results [7]. The flow angle b
is illustrated in Fig. 5. The resulting equation for the critical Reynolds number obtained by Cowell et al. [23] is
ReLp 4; 860=0:936  1:76=Lp =FP 0:995a

17

where a is the louver angle. The critical Reynolds number


has also been identified by Webb and Trauger [5] through
the flow visualization test. They noted that the flow angle is
dependent on the louver geometry and the Reynolds
number. Then, they introduced the flow efficiency to
correlate the data.
Fe

N
D

18

where N is the actual transverse distance and D is the ideal


transverse distance. The flow efficiency is illustrated in
Fig. 5. If all the flow passes the through the louvers, it would
be displaced the lateral distance D. Actual flow path is
displaced a lateral distance N. Webb and Trauger noted that,
below a certain Reynolds number, the flow efficiency
decreased as the Reynolds number decreased. Above the
value, which they defined as the critical Reynolds number,
the flow efficiency was independent of the Reynolds number.
Webb and Trauger developed the following equation for the
critical Reynolds number
ReLp 828a=900:34 :

19

Note that the critical Reynolds number by Eq. (19) is


only dependent on the louver angle.
The critical Reynolds numbers have been calculated
from both correlations, and they are summarized in
Table 4. Table 4 reveals that the critical Reynolds numbers
obtained from present j factors are insensitive to the louver
angle, and decrease as the Lp/Fp decreases. When the
present data are compared with the predictions by Cowell
et al. [23] and Webb and Trauger [5] correlations, a large

123

discrepancy is noted. Predicted values reveal strong


dependency on the louver angle and relative independency
on Lp/Fp. The present data show an opposite trendstrong
dependency on Lp/Fp and relative independency on the
louver angle. Note that the present louver fin has Lp/Fp
values from 1.21 to 1.70, which lie outside of the applicable range of Webb and Trauger (Lp/Fp up to 1.31) and
Cowell et al. (Lp/Fp up to 1.0) correlations. Table 4 shows
that the Webb and Trauger correlation predicts much larger
ReLp values compared with the present heat transfer results
or the Cowell et al. predictions. The reason may partly be
attributed to the small number of louver rows of Webb and
Traugers samples. Their samples consisted of five rows of
louvers. Recent study by Springer and Thole [24] and
Beamer et al. [25] revealed that, as fewer rows of louvers
are used, the flow is forced to become duct-directed
because of the end-wall effects. Duct-directed flow will
delay the fall-off of the j factor curve. Springe and Thole
[24] recommended samples of more than 19 rows for a
proper flow visualization test.
Kim and Kim [26] conducted flow visualization tests for
two louver geometries having Lp/Fp larger than 1.0 (Lp/Fp
= 1.0 and 1.4, a = 27). Tests were conducted in a water
tunnel using four times scaled-up models. The test section
comprised of 14 louver arrays. The flow efficiencies were
obtained from the flow visualization results, and they are
reproduced in Fig. 6. This figure shows that the flow efficiency decreases slightly as the Reynolds number
decreases for ReLp larger than 100. Below ReLp of 100,
however, it drops significantly, suggesting the critical
Reynolds number of approximately 100. This value is in
approximate accordance with the critical Reynolds numbers obtained from the heat transfer tests (first row of
Table 4). Figure 6 also shows that the flow efficiency
slightly increases as Lp/Fp increases. Also shown in the
graph are the predictions by Cowell et al. [23] and Webb
[27] correlation. Webb [27] updated the Webb and Trauger
[5] flow efficiency correlation. Figure 6 shows that Cowell
et al. correlation generally predicts the trend of the data,
although some overprediction is noticed. Webb correlation
reasonably predicts the data above the critical Reynolds
number. However, the fall-off of the data is not predicted.
Figure 4 shows that the effect of fin pitch is prominent at
low Reynolds numbers, and decreases as the Reynolds

Heat Mass Transfer (2008) 44:11271139

1135

1.0
0.9

Fe

0.8
Webb

0.7

Cowell et al.

0.6

Lp/Fp=1.0
Lp/Fp=1.4

0.5
0.4

50

100 150 200 250 300 350 400 450 500 550

Re Lp

Fig. 6 Flow efficiency of the louver geometries having large values


of Lp/Fp

number increases. In addition, one can clearly notice the


cross-over of j factor curves. It has been shown by Webb
and Trauger [5] that the flow efficiency increases as the fin
pitch decreases, which will yield a high heat transfer
coefficient. This is true for high Reynolds numbers. Figure 4 shows that, for high Reynolds numbers, the j factor
increases as the fin pitch decreases. For low Reynolds
numbers, however, the trend is reversed. The j-factor
decreases as the fin pitch decreases. This is an unexpected
result. Careful review of Achaichia and Cowell [4] data,
however, reveals the same trend. Their sample 3 (Fp =
1.65 mm, Lp = 1.4 mm) yielded lower j factor than that of
sample 6 (Fp = 2.15 mm, Lp = 1.4 mm). Then, why the jfactor decreases as the fin pitch decreases? Figure 7 shows
louver patterns for three different fin pitches at 27 louver
angle. For a given louver angle, different fin pitch yields
different transverse gap between the upstream and downstream louvers, as illustrated in Fig. 7. For the fin pitch
1.0 mm, louvers on the lower row almost align with those
on the upper row yielding very small gap between louvers.
This gap increases as the fin pitch increases. It is known

Fp

= 27

1.0 mm

1.2 mm

1.4 mm

Fig. 7 A graph showing louver patterns for the louver angle 27

that louvers enhance the heat transfer through repeated


growth of laminar boundary layer, followed by its dissipation in the wake region [28]. If the thermal and the
velocity boundary layers from the upstream louver are not
fully dissipated in the wake region, the heat transfer performance of the downstream louver will be deteriorated.
For the fin pitch of 1.4 mm, the gap between neighboring
louvers appears wide enough that the flow approaching the
downstream louver is not likely to be affected by the wake
from the upstream louver. For the free-stream velocity of
1.0 m/s, the boundary layer thickness at the trailing edge of
the present louver (width 1.7 mm) is approximately
0.9 mm. For the smallest fin pitch of 1.0 mm, where there
is virtually no gap, the downstream louver will be buried in
the boundary layer from the upstream louver. In this case,
the thermal performance will be deteriorated. This may
explain the effect of fin pitch at low Reynolds numbers.
With increasing Reynolds number, the boundary layer gets
thinner, and the downstream louver will eventually be
situated outside of the boundary layer even with the
smallest fin pitch. Then, the heat transfer coefficient will
increase as the fin pitch decreases.
Different from the j factor, the friction factor curve does
not show the fall-off characteristic. It continuously
increases as the Reynolds number decreases. For high
Reynolds numbers, most of the flow will be aligned with
louvers, and the skin friction will constitute most of the
pressure drop. For low Reynolds numbers, however, most
of the flow will follow the duct. In this case, a form drag by
louvers should be added to the skin friction. This may
explain the continual increase of the friction factor with
decreasing Reynolds number. The form drag does not
contribute to the heat transfer from louvers. Figure 4 shows
that the friction factor increases as the fin pitch decreases.
Same trend has been reported by Achaichia and Cowell [4]
and Chang and Wang [13]. As noted by Achaichia and
Cowell, more flow will be louver-directed for decreased fin
pitch, which will lengthen the travel distance of the flow,
and eventually increase the friction factor.
In Fig. 8, the data are re-plotted for different fin pitches.
Each graph shows the effect of louver angle. Figure 7
shows that the louver angle has minor influence on the j
factor for Fp = 1.4 mm. For smaller fin pitches, the j factor
increases as the louver angle increases. It has been shown
through flow visualization [5] that the flow efficiency
increases as the louver angle increases, which will yield a
high heat transfer coefficient. The effect of louver angle is
likely to decrease as the fin pitch increases. For a very large
fin pitch, for example, the flow will be duct-directed, and
louvers will behave like roughness elements on a surface.
In this case, the effect of louver angle will be negligible.
For Fp = 1.0 mm, j factors cross over at the Reynolds
number of approximately 140. Below the Reynolds

123

1136

Heat Mass Transfer (2008) 44:11271139


1

F p = 1.0mm
o

19
o
27

15
o
25

0.1

0.01
30

100

1000

2000

Re Lp

a) Fp = 1.0 mm
1
F p = 1.2mm
o

19
o
27

15
o
25

0.1

0.01
20

100

1000

2000

Re Lp

b) Fp = 1.2 mm
1

F p = 1.4mm
o

25

19
o
27

0.1

downstream louvers decreases as the louver angle increases, yielding virtually no gap at 27 louver angle. In this
case, the downstream louver will be buried in the boundary
layer from the upstream louver, which will decrease the
heat transfer coefficient. Figure 8 shows that the friction
factor increases as the louver angle increases. Same trend
has been reported by Achaichia and Cowell [4] and Chang
and Wang [13]. As noted by Achaichia and Cowell, more
flow will be louver-directed for increased louver angle,
which will lengthen the travel distance of the flow, and
eventually increase the friction factor.
The literature shows several correlations which predict
the heat transfer coefficients and friction factors of louver
finned heat exchangers. The present data are compared
with the predictions by Davernport[11], Achaichia and
Cowell [4], Sunden and Svantessen [12], Chang and Wang
[29], Chang et al. [30] and Kim and Bullard [15] correlations, and the results are shown in Fig. 10. The statistical
data are summarized in Table 5. The analytical model by
Sahnoun and Webb [14] was also assessed. For the heat
transfer coefficient, Achaichia and Cowell [4] correlation
and Sahnoun and Webb model [14] reasonably predict the
data except for the low Reynolds number range. At low
Reynolds numbers, the data are overpredicted. The reason
may be attributed to the extrapolation of the correlation or
the model. The Achaichia and Cowell correlation is
applicable for the Reynolds number larger than 150 as
reported by themselves. The Sahnoun and Webb model
includes the flow efficiency correlation, which was developed from the flow visualization data by Webb and Trauger
[5]. The flow visualization test was conducted in a water
tunnel for the Reynolds number range 400  ReLp  4,000.
Kim and Bullard [15] correlation overpredicts the
ReLp \ 100 data, and underpredicts the ReLp [ 100 data.
Figure 10 shows that the Davernport correlation highly

0.01
30

100

1000

2000

Fp = 1.0mm

15

Re Lp

c) Fp = 1.4 mm
19

Fig. 8 The j and f factors of present samples showing the effect of


louver angle
25

number, the sample having lower louver angle yield higher


heat transfer coefficient. This result appears, at first,
uncommon. Careful review of the louver pattern, however,
provides a clue to this uncommon behavior. In Fig. 9,
louver patterns for different louver angles are provided.
Figure 9 shows that the gap between upstream and

123

27

Fig. 9 A graph showing louver patterns for the fin pitch 1.0 mm

Heat Mass Transfer (2008) 44:11271139

1137

Fig. 10 The present data


compared with existing
correlations

jexp/jpred

Davenport[11]
Achaichia and Cowell[4]
Sunden and Svantesson[12]
Chang and Wang[27]
Sahnoun and Webb[14]
Kim and Bullard[15]
Present study

0
0

200

400

600

800

1000

1200

Re Lp

a) jexp/jpred
5
Davenport[11]
Achaichia and Cowell[4]
Sunden and Svantesson[12]
Chang and Wang[27]
Sahnoun and Webb[14]
Kim and Bullard [15]
Present study

fexp/fpred

0
0

200

400

600

800

1000

1200

Re Lp

b) fexp/fpred
underpredicts the data, and the Chang and Wang correlation overpredicts the data. Note that the present samples
have very small fin pitches, which most of the previous
correlations do not cover. Figure 10 shows all the correlations underpredict most of the present friction factors
except for the Achaichia and Cowell [4] correlation, which
overpredicts the data.
Since none of the existing correlations adequately predicted the data, an attempt was made to develop a new
correlation. Figure 4 shows that, for ReLp [ 150, the
present j factor increases as the Reynolds number decreases
and the fin pitch decreases. The trend is reversed for
ReLp 150. Thus, different correlations were developed for
the corresponding Reynolds number ranges. A multiple
regression procedure was carried out to correlate the data.
The potentially significant variables are a flow variable
ReLp and louver fin parameters (a, Fp, Lp, etc.). The final
correlation is as follows:

ReLp [ 150 :

j 0:705Re0:477
Lp

 a 0:271  L 0:155
p

90

Fp
20

ReLp \150 :

j 0:0311Re0:183
Lp

 a 0:0475  L 1:25
p

90

Fp
21

entire Reynolds number :

 a 0:493
f 8:42Re0:560
Lp
90
 0:535
Lp

:
22
Fp

Note that, for the j factor, the exponents of the Reynolds


number and (Lp/Fp) have different signs for different
Reynolds number ranges. The present correlations are
compared with the data, and the results are listed in

123

1138

Heat Mass Transfer (2008) 44:11271139

Table 5 Comparison of existing correlations with the present data


Investigators

10
(%)

30
(%)

50
(%)

100
(%)

(3)

Standard
deviation (%)

(4)
Davernport [11]
jexp/jpred

3.8

9.1

fexp/fpred

0.0

0.5

24.2
1.08

100.0

35.8

12.4

58.3

Achaichia and Cowell [4]


jexp/jpred

58.4

88.7

96.2

100.0

19.3

fexp/fpred

0.0

0.0

44.7

100.0

145.0

Sunden and Svantesson [12]


jexp/jpred
28.0
81.7
fexp/fpred

3.8

93.6

100.0

27.3

28.0

77.4

100.0

27.5

Sahnoun and Webb [14]


jexp/jpred

74.2

87.1

91.9

100.0

24.3

fexp/fpred

0.5

2.2

6.5

76.3

45.4

65.1

88.2

100.0

58.2

28.0

72.0

100.0

27.6

Chang and Wang[27]


jexp/jpred

0.0

Chang et al. [28]


fexp/fpred

5.9

Kim and Bullard [15]


jexp/jpred

5.3

85.6

95.8

100.0

23.0

fexp/fpred

23.0

64.7

96.2

98.4

21.4

jexp/jpred

91.9

100.0

100.0

100.0

6.7

fexp/fpred

93.5

100.0

100.0

100.0

5.6

This study

Table 5. The present correlation predicts 92% of the heat


transfer coefficient and 94% of the friction factor within
10%. Figure 10 shows that the prediction is excellent
even for the low Reynolds number range.

4 Conclusions
In this study, the heat transfer and pressure drop characteristics of heat exchangers having louver fins were
experimentally investigated. The samples had small fin
pitches (1.01.4 mm), and experiments were conducted up
to very low frontal air velocities (as low as 0.3 m/s). Listed
below are the major findings.
(1)

(2)

The critical Reynolds number, where the fall-off of j


factor curve occurs, is insensitive to the louver angle,
and decreases as the Lp/Fp decreases. Existing
correlations on the critical Reynolds number do not
adequately predict the data.
For high Reynolds numbers, j factors increase as the
fin pitch increases. For low Reynolds numbers,
however, the trend is reversed yielding cross-over of
j factor curves. Possible explanation is provided
considering the louver pattern at different fin pitches.

123

(5)

The j factor increases as the louver angle increases.


At the small fin pitch (Pf = 1.0 mm), however, the
cross-over of j factor curves occurred.
Different from the j factor, the friction factor does not
show the fall-off characteristic. The reason may be
attributed to the form drag by louvers, which offsets
the decreased skin friction at low Reynolds numbers.
The friction factor increases as the fin pitch decreases
and the louver angle increases.
A new correlation predicts 92% of j factors and 94%
of f factors within 10%.

References
1. Webb RL (1992) Air-side performance of enhanced brazed aluminum heat exchangers. ASHRAE Trans 98(2):391410
2. Webb RL, Lee H (2001) Brazed aluminum heat exchangers for
residential air-conditioning. J Enhanc Heat Transf 8:114
3. Davernport CJ (1980) Heat transfer and fluid flow in louvered
triangular ducts. PhD thesis, Lanchester Polytechnic, UK
4. Achaichia A, Cowell TA (1988) Heat transfer and pressure drop
characteristics of flat tube and louvered plate fin surfaces. Exp
Therm Fluid Sci 1:147157
5. Webb RL, Trauger PE (1991) Flow structure in the louvered fin
heat exchanger geometry. Exp Therm Fluid Sci 4:205217
6. DeJong NC, Jacobi AM (2003) Localized flow and heat transfer
interactions in louvered fin arrays. Int J Heat Mass Transf
46:443455
7. Achaichia A, Cowell TA (1988) A finite difference analysis of
fully developed periodic laminar flow in inclined louvered arrays.
In: Proceedings of second UK national heat transfer conference.
Glassgow, pp 883888
8. Atkinson KN, Drakulic R, Heikal MR, Cow ell TA (1998) Two
and three dimensional numerical models of flow and heat transfer
over louvered fin arrays in compact heat exchangers. Int J Heat
Mass Transf 41:40634080
9. Tafti DK, Wang G, Lin W (2000) Flow transition in a multilouvered fin array. Int J Heat Mass Transf 43:901919
10. Davernport CJ (1983) Heat transfer and flow friction characteristics of louvered heat exchanger surfaces, in Heat Exchangers:
Theory and Practice, Hemisphere Pub, pp 387412
11. Davernport CJ (1983) Correlation of heat transfer and flow friction characteristics of louvered fin. AIChE Symp Ser 79:1927
12. Sunden B, Svantessen J (1992) Correlation of j and f factors for
multi-louvered heat transfer surfaces. In: Proceedings of third UK
national heat transfer conference, pp 805811
13. Chang YJ, Wang CC (1996) Air-side performance of brazed
aluminum heat exchangers. J Enhanc Heat Transf 3(1):1528
14. Sahnoun A, Webb RL (1992) Prediction of heat transfer and
friction for the louver fin geometry. J Heat Transf 114:893900
15. Kim MH, Bullard CW (2002) Air-side thermal hydraulic performance of multi-louvered fin aluminum heat exchangers. Int J
Refrigeration 25:390400
16. ASHRAE Standard 41.1 (1986) Standard method for temperature
measurement, ASHRAE
17. ASHRAE Standard 41.2 (1987) Standard method for laboratory
air-flow measurement, ASHRAE
18. ASHRAE Standard 41.5 (1975) Standard measurement guide,
engineering analysis of experimental data, ASHRAE
19. Incropera FP, Dewitt DP (1990) Fundamentals of heat and mass
transfer, 3rd edn. Wiley, London

Heat Mass Transfer (2008) 44:11271139


20. Gnielinski V (1976) New equations for heat and mass transfer in
turbulent pipe flows. Int Chem Eng 16:359368
21. Schmidt TE (1949) Heat transfer calculations for extended surfaces. J ASRE Refrig Eng 4:351357
22. Kays WM, London AL (1984) Compact heat exchangers, 3rd
edn. McGraw-Hill Pub, New York
23. Cowell TA, Heikal MR, Achaichia A (1995) Fluid flow and heat
transfer in compact louvered fin surfaces. Exp Therm Fluid Sci
10:192199
24. Springer ME, Thole KA (1998) Experimental design for flow
field studies of louvered fins. Exp Therm Fluid Sci 18:258269
25. Beamer HE, Ghosh D, Bellows KD, Huang LJ, Jacobi AM (1998)
Applied CFD and experiment for automotive compact heat
exchanger development, SAE 980426

1139
26. Kim NH, Kim HJ (2007) Flow efficiency in louvered fin geometries having large louver pitch to fin pitch ratio. Int J Air Cond
Refrig (submitted)
27. Webb RL (1991) Letters to the editor. Exp Therm Fluid Sci 4:374
28. DeJong NC, Jacobi AM (1997) An experimental study of flow
and heat transfer in parallel plate arrays: local, row by row and
surface average behavior. Int J Heat Mass Transf 40(6):1365
1378
29. Chang YJ, Wang CC (1997) A generalized heat transfer correlation for louver fin geometry. Int J Heat Mass Transf 40(3):533
544
30. Chang YJ, Hsu KC, Lin YT, Wang CC (2000) A generalized
friction correlation for louver fin geometry. Int J Heat Mass
Transf 43:22372243

123

You might also like