Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Flow Structure in the Louvered Fin

Heat Exchanger Geometry


Ralph L. Webb
Paul Trauger
Department of Mechanical Engineering,
The Pennsylvania State University,
University Park,. Pennsylvania

A flow visualization study of the louvered fin geometry commonly used in


automotive heat exchangers was performed using a dye injection technique with
10:1 scale models. The geometrical parameters, louver pitch, louver angle, and
fin pitch were varied to determine their effect on the flow structure. Tests
covered Reynolds numbers of 400-4000 based on louver pitch. Data are
presented in the form of a dimensionless flow efficiency (defined in terms of the
mean flow angle relative to the louver angle) and Reynolds number. Correlations are developed to predict the flow efficiency as a function of dimensionless
geometrical groups and Reynolds number. The flow structure is also discussed.

Keywords: compact heat exchanger, louver fin, flow visualization


INTRODUCTION
Enhanced surface geometries permit reductions in heat exchanger size and weight. Their use is of particular importance
in the automotive industry, where a premium is placed on
space, weight, and cost [1]. For the case of automobile
radiators (Fig. 1), the dominant thermal resistance generally
occurs on the airside, which may account for 80% or more of
the total thermal resistance. Therefore, a reduction in the
air-side thermal resistance will improve the heat transfer
performance of the heat exchanger. This reduction can be
achieved by several means. First, additional heat transfer
surface area can be obtained through the use of extended
surfaces. Second, heat transfer enhancement techniques can
be employed to increase the gas-side heat transfer coefficient.
This is generally accomplished through the use of features on
the extended surface. For gas-side applications, fin geometries fall into two categories: continuous and interrupted
surfaces. Continuous surfaces achieve heat transfer enhancement through the secondary flow patterns introduced by
sudden velocity changes. Examples of continuous surfaces
are wavy and dimpled plate fins. On the other hand, interrupted surfaces (Figs. 2 and 3) achieve heat transfer enhancement by the continuous growth and destruction of laminar
boundary layers on the interrupted portion of the geometry.
Examples of interrupted surfaces are the offset strip fin and
louvered fin.
Louvered fins can be manufactured by high-speed production techniques and as a result are less expensive than other
interrupted flow geometries when produced in large quantities. For this reason, they have found wide use in the
automotive industry. Louvers are generally formed by cutting
the metal and pushing out the cut elements from the plane of
the base metal [2]. Commonly, corrugated fins (shown in
Fig. 1) provide the base for the louvered surface (shown in
Fig. 2); however, there are some applications that require the
use of plate fins. Louvered surface geometries can be made in
a wide variety of shapes depending on fabrication techniques
and consWaints.

Although louvered surfaces have been in existence since


the 1950s, it has been only within the past 20 years that
serious attempts have been made to understand the flow
phenomena and performance characteristics of the louvered
fin. Kays and London [3] were the first to report heat transfer
and pressure drop data on louvered fins. However, the geometries of the test samples reported by Kays and London [3]
do not reflect present industrial designs, so the data are of
little present value. To date, the only known open literature
sources for heat transfer and pressure drop data on the
louvered fin are Davenport [4, 5] and Achalchia and Cowell
[6]. Davenport tested single-row, corrugated fin louvered
heat exchangers, and Achaichia and Cowell tested one- and
two-row louvered fin exchangers having round tube and flat
(louvered) fin geometry.
Although experimental data will always be used in the
design of heat exchangers, researcher have begun to use
analytical [7-9] and numerical [10, 11] techniques to predict
the performance characteristics of the louvered surface geometry. These methods can provide detailed understanding of
the enhancement mechanism of the louver geometry, if the
model is correct. However, results of numerical methods may
be misleading because of the assumptions made. This will
occur if the model assumes laminar flow parallel to the louver
or is incapable of predicting flow separations and the effect of
the resulting eddies on the flow structure. The objectives of
this study are to investigate the flow phenomena in the
louvered fin array and to establish the influence of velocity
and geometrical parameters on the flow structure within the
array. This information will be useful in determining the
validity of proposed analytical and numerical solutions. Also,
it is intended that the study will provide insight into improved
fin design.
LITERATURE REVIEW
Beanvals [12] appears to have been the first to conduct flow
visualization experiments on the louvered fin array. In his
1965 study, a smoke flow visualization technique was used

Address correspondenceto Professor Ralph L. Webb, Departmentof Mechanical Engineering, The PennsylvaniaState University, University1Park, PA
16802.

Experimental Thermaland Fluid Science 1991; 4:205-217


1991 by ElsevierScience Publishing Co., Inc., 655 Avenueof the Americas, New York, NY 10010

0894-1777/91/$3.50

205

206 R.L. Webb and P. Trauger


COOLANT
FIN
TUBE

Figure 1. Typical automotive radiator with louvered fins.


Cell

with 10:1 scale models. The results are only of qualitative


value, since the actual geometric dimensions and velocities
were not documented in the paper. From inspection of the
photographs included in the paper, it appears that the louver
angle (0) was approximately 30* and the louver-to-fin pitch
ratio (Lp/Fp)was approximately 0.80. His figures show that
the main flow was nearly parallel to the louvers for the
velocities tested. Prior to this work, it was speculated that
louvers acted as a surface roughness that enhanced the performance characteristics of the fin by promoting turbulence.
In 1973, Wong and Smith [13] measured heat transfer and
pressure drop data on a 5:1 scale model of a typical louvered
fin array. Comparison of the model data with those of similar
full-scale louvered arrays showed good correlation between
the model and actual cores. It was concluded that the airflow
phenomenon was similar for the full-scale and model cores.
In 1980, Davenport performed flow visualization experiments identical to those of Beauvais and demonstrated that
the flow structure within the louvered array was a function of
Reynolds number (ReLp). At low values of ReLp, the louvers
had only a slight influence on the flow structure. Thus, the
main flow stream did not pass through the louvers. However,
at high values of ReLp the flow became nearly parallel to the
louvers. Davenport 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. This information is discussed by
Achaichia and Cowell [6].
Figure 4 shows heat transfer data of Achaichia and Cowell
[6] for three louver fin geometries used on plate fins with 11
mm tube pitch. At the hightest ReLp, the data are parallel to
(but lower than) that for laminar boundary flow over a flat
plate (the "flat plate" line). At low ReLp , the data show
characteristics similar to those of laminar duct flow. The
"duct flow" line is for fully developed laminar flow in a
rectangular channel with a 3.3:1 aspect ratio. The aspect ratio
for the finned tubes is between 2.7 and 4. This behavior is
consistent with Davenport's flow observations. Since the heat
transfer performance is closely related to the flow structure,

Bose S u r f a c e

Figure 3. Cross section of the offset strip fin array.


we may infer that two types of flow structure exist within the
louvered, flat fin array:
1. Duct flow, in which the fluid travels axially through the
array, essentially bypassing the louvers
2. Boundary layer flow, in which the fluid travels parallel to
the louvers
Figure 4 shows the Stanton number (St) characteristics for
three samples, numbers 1 and 2, where the louver angles are
29* and 20", respectively. For both samples Fp/Lp = 2.63.
The louver angle affected the Stanton number only in the
transition region between the duct and boundary flow regions.
Achaichia and CoweU [11] also numerically modeled the
flow through the louvered fin array. Figure 5, taken from
their work, shows that as the Reynolds number approaches
large values, the mean flow angle u approaches the louver
angle 0 to within a few degrees. The mean flow angle a is
the integrated average value over the length directional flow
path [11]. Their analysis, which assumes a fully developed,
periodic laminar flow, supports Davenport's hypothesis concernlng the boundary layer development on the louvers at low
values of ReLp. The study shows that the flow structure is a
function of Reynolds number ReLo , the louver-to-fin pitch
ratio Lp/Fp, and the louver angle/9. Figure 5 shows that the
mean flow angle is a strong function of ReLp at low ReLp and
is independent of ReLp at high ReLp. If eddies are shed from
the louvers, because the flow is not parallel to the louvers,
1

I I 111

L = 0.81 mm

0'1 : ~
~

Line

FalLp

~ (deg)

260

29

2.66

2o

4.11

29

z
c
.

. . . . . . .

Duct FLow~

_
6')
0"01

10
SECTION

A--A

F'IN

GEOMETRY

PER

TUBE

ROW

Itlnll

100

I I Ill

1000

R e y n o l d s N u m b e r - ReLp

Figure 2. Cross section of louvered fin array exchangers of the

Figure 4. Heat transfer characteristics of several Iouvered plate

plate fin geometry.

fin geometries [6].

Flow Structure in the Louvered Fin


1.0
0.9
0.8
0.7

C~

~5

0.6

0.s
0.4
0.3
0.2
0.1
,

0.0
10

100

, ,

1000

ReLp
Figure 5. Mean flow angle dependence in a louvered fin array.
(From Achaichia and Cowell [11].)
their analysis will probably not accurately predict the effects
of these eddies on the heat transfer coefficient and friction
factor.
Recently, Howard [14] performed flow visualization experiments on a 10:1 scale model of a two-dimensional louvered
fin array using a dye injection technique. Howard observed
that flow instabilities (i.e., vortex shedding) commenced at a
Reynolds number based on fin thickness of approximately
Re t = 30 (ReLp = 900) and that the instabilities progressed
upstream (from the exit end of the array) as the velocity was
increased. He characterized the flow structure as being either
"efficient" or "inefficient." The flow was considered to be
efficient if it was approximately parallel to the louvers.

L=160

Fp=15

Conversely, the flow was considered to be inefficient if it was


predominantly axial (i.e., did not go through the louvers).
This definition is based on the assumption that an efficient
flow structure should yield a higher heat transfer coefficient
than one having the inefficient flow pattern. Howard's definitions are consistent with the definitions of "duct flow" and
"flat plate" flow shown on Fig. 4. If the flat plate flow line
of Fig. 4 persisted into the low-ReLp region, St would be
considerably higher than the experimental values.
Howard observed a transient region between inefficient and
efficient flow, just as a transition region exists from laminar
to turbulent flow. His study, which was performed for a 20*
louver angle, showed that the transition region occurred at
Lp/Fp = 0.7-0.8. The flow structure was considered to be
efficient if Lp/Fp > 0.8 and inefficient if Lp/Fp < 0.7.
Kajino and Hiramatsu [10] also conducted a flow visualization study of the louvered fin array using dye injection and
hydrogen bubble techniques. Figure 6 shows photographs of
flow through the array for a louver angle of 26* Lp/Fp =
0.67 at low Reynolds number (ReLp = 500). Boundary layers
exist on both the upper and lower surfaces of the louver, and
a laminar wake exists downstream from each louver. Flow
separation was observed on the back side of inlet louvers.
Kajino and Hiramatsu concluded that heat transfer enhancement is due to the thin boundary layers that form at the
leading edge of each louver. However, this will be true only
if the flow passes through the louvers.
Figure 7 shows two louver arrays that have the same
louver pitch but different fin pitches. In the left-hand portion
of Fig. 7, a significant fraction of the flow bypasses the
louvers. This is because the hydraulic resistance of the "duct"

Lp=lO

207

~=26

ReL=500

Figure 6. Visualization of flow in the louvered fin array [10].

208

R. L. Webb and P. Trauger


L =160

Lp = 10

e=26'

ReL = 500

LARGE FIN PITCH F p = 2 0

SUITABLE FIN PITCH F p = l O

Figure 7. Streamlines in a model of the louvered fin array [10].


flow region is substantially smaller than that for boundary
layer flow across the louvers. When the fin pitch is reduced,
as in the right-hand part of Fig. 7, the hydraulic resistance of
the "duct" is increased, so that most of the flow passes
through the louvers.
EXPERIMENTAL PROGRAM
In order for flow visualization experiments using large-scale
models to accurately simulate the flow in a full-scale array,
both geometric and dynamic similarity must be satisfied. In
addition, the turbulence levels of the model and prototype
flows should be the same. Beauvais made turbulence measurements on large-scale and full-scale geometries with air as
the fluid and found that the two turbulence levels were in fair
agreement.
Previous studies demonstrated that the geometrical parameters that are most likely to influence the flow structure in the
louvered fin array are the louver pitch Lp, louver angle 0, fin
pitch Fp, and fin thickness t. The flow visualization model
geometries selected for testing were 10:1 scale and had
Lp/Fp, Lp/t, and 0 typical of common industrial designs.
Table 1 gives the model geometries. Two louver angles (20*
and 30*) were selected. For both louver angles, the louver
pitch Lp was 15 nun and the fin pitch Fp was variable from
11 to 30 ram. This allowed for testing of 0.49 < Lp/Fp <
1.31.
The louvered arrays selected permitted study of the effect
of louver angle 0 and the louver-to-fin pitch ratio Lp/Fp on
the flow structure within the array. Each model consisted of
10 louvered sections in the spanwise direction. The range of
Reynolds numbers was approximately 400 < ReLp < 4000,
which corresponds to a frontal air velocity range of 2.8 < Ufr
< 28.0 m/s. The louvers were made from 0.635 nun thick
brass sheet were approximately 150 nun high. All louvers

had two tabs at each end (as shown in Fig. 8), which allowed
the louver to be mounted on support plates, thus preventing
the louvers from moving during the experiment. The position
of the holes in the support plates fixed the desired louver
angle. Top and base support plates were made from 3.18 mm
Lexan.
The flow visualization tests were conducted in a closed-circuit, open water channel. The channel was approximately 5
m long, 0.3 m wide, and 0.46 m deep. Figure 9 shows a
schematic plan view of the flow visualization test section.
Water was pumped from a reservoir and through the channel
using one of two pumps for different flow rate ranges. A
metal screen at the upstream end of the channel provided a
uniform water velocity. A bank of drinking straws was placed
15 mm upstream from the model to serve as a flow straightener. The flow rate was measured with two calibrated orifices
connected to a manometer. For a given flow rate, the depth of
the water in the channel was adjusted by raising or lowering a
gate at the downstream end. The height of the water channel
was adjusted to be just above the top support plate. Water
temperature was measured with a thermometer located downstream of the test section. The width of the model varied
depending on which fin pitch was being tested. A contraction

INLET DEFLECTIONLOUVER

Table 1. Model Geometries

Lp

Fp

Model

(deg)

(mm)

(ram)

(mm)

1
2

20.0
30.0

15.0
15.0

0.635
0.635

11-30
11-30

Figure 8.

=ii

.,~

llr

=-

Q,

':,

/-7

BASESUPPORT
PLATE

Construction detail of the model of the ]ouvered fin

array tested.

Flow Structure in the Louvered Fin


DYE INJECTION

209

The denominator of Eq. (1) is given by


FLOW STRAIGHTENER

D = (t/2

Figure 9. Sketch of test section used for flow visualization.


section was constructed to narrow the width of the water
channel to that of the model (not shown in Fig. 9).
Flow visualization was performed using a dye injection
technique. The dye was a powder (Best-acid azure blue from
Best Color and Chemical, Inc.) that was mixed with water.
This solution was then mixed with approximately 20% whole
milk. The fat content of the milk is presumed to reduce the
diffusion of the dye filament [15]. Dye was injected into the
flow with a hypodermic needle, which was connected to a dye
reservoir by plastic tubing. The dye reservoir was mounted
1.0 m above the channel test section, which allowed gravity
flow of the dye. The injection rate of the dye was controlled
by a small adjustable clamp placed on the tubing. Typically,
dye was injected in a straw of the flow straightener, so that it
entered at the fin pitch centedine and at one-half the fin
height (75 mm). The maximum boundary layer thickness at
the exit end of the louver array support plates was 38 mm.
The visualization tests were conducted as follows:
1. Place the appropriate model in the channel, and set up the
contraction section, if required.
2. Set the flow rate at the lowest value, and adjust the height
of the water to be just above the top support plate.
3. Introduce the dye continuously, and observe the flow.
Observations were recorded in a notebook, and sketches
of the flow streamlines were made on a schematic diagram
for the specific model geometry.
4. Increase the flow rate incrementally, and note observations until the entire Reynolds number range is covered.

- s)tano

O)

The term S exists in Eq. (2) because the flow deflection


occurs halfway across the first louver. The L e of the first
louver is the same as for the downstream louvers. The
numerator of Eq. (1) was taken from the flow pattern sketches
made during the test runs. The definition of flow efficiency
was selected on the basis of its practical nature. The flow
efficiency is equal to 1 when the flow is parallel to and
through the louvers. It is equal to zero when the flow is axial
through the array (100% duct flow).
EXPERIMENTAL RESULTS
Data were taken for six louver-to-fin pitch ratios ( L p / F p =
0.49, 0.56, 0.66, 0.79, 0.98, and 1.31) and two louver
angles (0 = 20 and 30). Data were taken over a Reynolds
number range of 400 < ReLp < 4000. Figures l l a and l l b
present flow efficiency (for all six pitch ratios) plotted against
Reynolds number for louver angles of 20 and 30 , respectively. As will be seen, the Reynolds number based on louver
pitch was found to be a better correlating parameter than that

1.0

0.9

D
[]

0.8

0.7

[]

- - 1 4 - - 1

.....

[2

0.6

g-

"

[]

Symbol
0

0.5

Lp/Fp
049
0.56
0.66
0.79
0.98
131

z~

0.4
0

0.3

02

0:20

Ol

O0
1000

3OO

Data reduction was a simple procedure. To quantify the


data, a Reynolds number (ReLp) and "flow efficiency" (~/)
were calculated. Figure 10 provides a visual definition of the
flow efficiency. The flow efficiency 71 is defined from Fig. 10

ReLp
(a)

as

N
*/ = D

1.0

actual transverse distance


ideal transverse distance

m-

(1)

,11

[3

DO &D "A ~[3A&'O ~


.
O A& @O A@

0.9
0.8

&

0 0

Symbol

Lp/Fp
0.49
0.56

0.66

0.79

[]

0.98

1.31

0.7

0.6
~"

0.5

0.4
0.3
0.2

=>

0=30"

0.i
0.0
300

I000
Re

------

IDEAL
ACTUAL

STREAMLINE
STREAMLINE

Figure 10. Definition of method used to define, the flow efficiency.

Lp

Co)
Figure 11. (a) Flow efficiency versus Reynolds number for
0 = 20". Co) Flow efficiency versus Reynolds number for 0 =
30".

210 R . L . Webb and P. Trauger


based on hydraulic diameter (RED). One can convert the data
using Re o = ReLp[2(Fp - t)/Lp].
The maximum error associated with the ReLR is +--2%, and
for measurement of dimension N in Eq. (1) it was 3 mm.
Using the 3 mm uncertainty of dimension N in Eq. (1), the
greatest uncertainty of 7/is +2.2%.
D I S C U S S I O N OF R E S U L T S
Although great care was exercised in the experiments, the
fact that the flow efficiency data were based on visual observations raises some uncertainty. However, the data were
generally checked for repeatability, and good agreement was
found.

Effect of Reynolds Number


Figures 11 a and 11 b show 7/ versus ReLp for louver angles
of 20* and 30*. Inspection of these figures reveals that the
flow efficiency increases with increasing ReLp. This occurs
up to a particular Re value, which we define as the critical
Reynolds number, Re~.p. Above Re[p, the flow efficiency
becomes independent of Reynolds number for fixed Lp/Fp.
From inspection of Figs. I 1a and 11 b, it appears that the
critical Reynolds number is independent of Lp/Fp for a
fixed louver angle. Comparison of these figures reveals that
the critical Reynolds number decreases slightly with increasing louver angle. For louver angles of 20* and 30", the
critical Reynolds number is approximately 1380 and 1200,
respectively.
As the Reynolds number increases from small to large
values, the flow pattern will transition from the duct flow
pattern (left-hand portion of Fig. 7) to approach a flat plate
flow pattern (right-hand portion of Fig. 7). Note that the
maximum flow efficiency at high ReLp is lss than 1.0 for the
smaller values of Lp/Fp.
The following empirical correlation was developed to predict the critical Reynolds number as a function of the dimensionless louver angle.
Reap = 828(0/90) -0.34

direction. This reduced hydraulic resistance results in a


greater fraction of the flow passing in the axial direction. As
Lp/Fp is increased (smaller fin spacing), the axial direction
hydraulic resistance increases, forcing more of the flow in the
louver direction. At the smallest fin spacing (Lz,/Fp = 1.31),
the hydraulic resistance in the louver direction is much
smaller than in the axial direction, so all of the flow will go in
the louver direction.

Effect of Louver Angle (0)


A comparison of Figs. 11 a and 11 b shows that increasing the
louver angle (for fixed Lp/Fp) increases the flow efficiency.
This holds true up to the critical Reynolds number (Re~p),
after which the flow efficiency is independent of louver angle.
As Lp/Fp increases (for ReLp < Re~p), the effect of louver
angle on flow efficiency decreases. Flow efficiency may vary
by as much as 30% for the two louver angles tested. For
ReLp < Reap and Lp/Fp = 1.31 (small fin spacing), the
effect of louver angle becomes very small. It should be noted
that changing the louver angle from 20* to 30* increases the
projected blockage area by nearly 50%. However, the flow
will incur a larger turning loss as the louver angle is increased.
Prediction o f Flow Efficiency
Inspection of Figs. 11 a and 11 b shows two distinct Reynolds
number regions. Therefore, separate flow efficiency correlations are required for each of these regions. These correlations are given below.

For Re > Re*p For this region, the flow efficiency is a


function of Lp/Fp alone. The following empirical correlation predicts the flow efficiency for ReLp > Reap within
___2.5%.
r/ = 0 . 9 5 ( L p / F p )

23

(4)

(3)

For Re < Re~.v In this region, ~/ is a function of 0,


where 0 is in degrees.
Effect o f Louver to Fin Pitch Ratio
From inspection of Figs. 11 a and 11 b, it is evident that the
flow efficiency increases with increasing Lp/Fp. The flow is
parallel to the louvers (~/ = 1) only if Lp/Fp >_1.31 (the
highest value tested). For Re > ReLp and 0 -- 20 , 30 , the
flow efficiency varies by nearly 30% over the range of
Lp/Fp tested (0.49 < Lp/Fp < 1.31). For Lp/Fp = 0.49
and 1.31, the asymptotic flow efficiency is 0.78 and 1.0,
respectively.
We propose the following qualitative model to explain the
flow efficiency versus Lp/Fp and ReLp relationship. The
flow must choose whether it will follow the louver direction,
the axial direction, or some intermediate angle between the
two directions. Which direction it goes will depend on the
hydraulic resistance of the flow path in the louver direction
relative to that in the axial direction. As L_/Fp is decreased
with fixed Lp, the fin spacing increases, l~owever, the flow
cross-sectional area between the louvers is unchanged. It
appears that this wider fin spacing reduces the hydraulic
resistance in the axial direction relative to that in the louver

Lp/Fp,

and ReLp. A multiple regression correlation was


performed to provide the best fit of the data in this region.
The resulting correlation is
7/ = 0.091(ReLp)39(

Lp/Fp) 0.44(0/90) 0.3

(5)

Approximately 90% of the data are correlated within


_+ 10% with this equation. The correlation is only moderately
successful in predicting the flow efficiency near the critical
Reynolds number. Also, small overprediction, for example,
6 - 8 % , will occur for Lp/Fp = 1.31, resulting in flow efficiencies greater than unity. Should the user predict ~ > 1,
one should use 71 = 1.0.
Since Eqs. (4) and (5) are empirical, we do not recommend
that they be extrapolated beyond the range of the experimental data. The range studied is 400 _< ReLp < 4000, 0.49 _<
Lp/Fp < 1.31, and 20* -< 0 -< 30*.

The Flow Structure


Figure 12a shows a flow pattern map for 0 = 30 with
and 1.31. The 17 versus ReLp data points were

Lp/Fp = 0.56

Flow Structure in the Louvered Fin


1.0
d

0.9

0.8
0.7

//

0.6

/
///

0.5

//
o

Lp/Fp
0.56

1.31

Symbol

0.4

0=30

0.3
0.2

0.i
0.0
300

1000

Rehp
(a)
Lp/Fp = 0.56,
O.

LOW ReLp

e =30 deg.
b.

ReLp> ReLp

/
!
Lp/Fp= 1.31, O =30deg.
c.

LOW ReLp

/ / / /

d.

ReLp> Re'~p

/ / / /

(b)
Figure 12. (a) Flow pattern map for 0.56 < L p / Fn < 1.31. Co)
Defined flow patterns for (a).

taken from Fig. l lb. Figure 12b shows the flow patterns
associated with the symbols a, b, c, d on Fig. 12 a. Sketches
a and b of Fig. 12 b show the flow pattern for small Lp/Fu
at small and high ReLp, respectively. Similarly, sketches c
and d show the corresponding flow patterns for high Lp/Fp
(smaller fin spacing). Purely laminar flow on the louvers and
in the wakes was observed for ReLp < 500-600 at all Lp/Fp
and at both louver angles. However, separated flows were
observed at wide fin pitches (small Lp/Fp), as discussed
below.
Sketch a of Fig. 12 b shows the flow pattern observed for
wide fin spacing at low ReLp. A recirculation zone was
observed at the inlet detection louver and at the first two full
louvers. Eddies were shed at the downstream side of the
louvers. As Re L_ is increased, the flow angle increases and
the extent of th~ separated flow decreases. At the higher
ReLp, the trailing wakes rapidly mix the dye, and it was
difficult to perceive very small separated flow zones. For
ReL~ > 1200, 7/ = 0.82, for which the flow angle is 25
compared to the 30 louver angle. We were unable to discern
flow separations, although trailing vortices may have existed.
For a smaller fin spacing, Lp/F. = 1.31, the flow can
more easily approach the louver ang(e, as shown in sketches

211

c and d of Fig. 12b. At the lowest ReLp for LplFp = 1.31,


~/ = 0.94, for which ot = 28.2 , as compared to the 30*
louver angle. No recirculation zones were observed. The
boundary layer on the louver was laminar, as were the
wakes. The laminar wake appeared to fully dissipate before
the flow reached the downstream louver. At the highest ReLp,
the flow was parallel to the louver, so no recirculation zones
were observed. However, the wake region was not laminar
and showed intense mixing of the dye.
A criterion was not developed to define the transition from
laminar to turbulent wakes. However, Figure 12 a shows that
the low values of ~/at low ReLp initiate flow separations and
thus wake instability. The wake instability does not occur
until higher values of ReLp for a smaller fin spacing (larger
Lp/Fp). The visual observations performed here show that
nonlaminar wakes can occur for ReLp < Re*p.
The numerical solution of Kajino and Hiramatsu [10]
assumed laminar flow on the louvers and in the wakes and
used the boundary layer approximations, which are unable to
predict flow separation. Hence, their model was unable to
account for the separated flow structure and the effect of
eddies observed in the present study.
Achaichia and Cowell [11] numerically solved the elliptic
form of the Navier-Stokes equations for "fully developed
flow, in the periodic sense" to predict the flow results plotted
on Fig. 5. This method is capable of predicting flow separations, but it ignores the possibility of entrance region behavior. It is interesting to compare their predicted results with
the present experimental results. Figure 13 shows the Fig. 5
predictions superimposed on the two data sets used to prepare
Figs. 11 a and 11 b. Figure 13 shows that the predicted results
agree with the asymptotic value, 7/ = 0.9. However, the data
show that the flow efficiency begins its drop at a Reynolds
number approximately 10 times higher than that predicted by
the numerical solution [11]. Our flow visualization showed
stronger flow separation and recirculation zones at the first
few louvers than occurred deeper in the array. The observed
"entrance region" behavior may partially explain the difference between the observed flow and that predicted by
Achaichia and Cowell.
The analytical modelers [7-9] assumed a laminar flow
parallel to the louvers. Flow parallel to the louvers was
observed only for Lp/Fp >_1.31 and ReLp > Reap. As previously described, it appears that the flow angle relative to

1.0

q~

. . . . . . . .

. . . . . . . .

0.9

0.8
0.7

~ / / / / I -

~,4gD*

i/

0.5

//

/
0.3

//

0,2

Legend:
0 Offi20(exp)
----

0.8

effi20O(pre)
effi30(pre)

//

o.1

Lp/Fp=2/3
10

100

1000

ReLp
F i g u r e 13. Comparison of measured flow efficiency with that
predicted by Achaichia and Cowell [11].

212

R. L. Webb and P. Trauger


VARIANT 14. OSF, AND FLAT PLATE VS. REIp

,.ooo

........
~f-louv

........

';-

/f-

~.~

"~

, , r 1.000

o.

efficiency-0.600

~'~_~,.,,.~
(ReL~ ~
0.010

. . . . . . . .

10

"

0.200
~

. . . . . .

100

oo

~
,,I

- - IO UV
=,

0.000

1000
Relp

Figure 14. Curves for j and f

versus ReLp for the louvered and


offset strip fins of the same dimensionless parameters. Curve of
7/versus Re for the louvered fin geometry.

the louvers may be related to the hydraulic resistance of the


flow path parallel to the louvers, relative to the flow path that
entirely bypasses the louvers. The boundary layer on the
louvers should be laminar for practical operating ranges. For
a given ReLp, a higher value of ~ will exist for larger values
of Lp/Fp because of the higher resistance of the axial flow
path. For a given Lp/Fp, the axial flow path may be viewed
as a rough surface. At low velocity, this flow path may be
"hydraulically smooth." One may speculate that as the flow
speed increases, the friction factor and pressure drop of the
axial flow path substantially increase, forcing more of the
flow in the louver direction.
APPLICATION TO HEAT TRANSFER
DESIGN
An objective of the study was to identify louvered surface
designs that provide improved performance. The results suggest that the highest heat transfer performance should occur
for large values of Lp/Fp. As the value of Lp/Fp is
reduced, a smaller fraction of the flow follows the louvers,
and eddies are shed from the trailing edge of the louvers.
Such eddies should yield a form drag component, which may
increase the pressure drop without significantly adding to the
heat transfer. This possibility was investigated by studying
the heat transfer and friction data of Achaichia and Cowell
[6] and then comparing these data to the corresponding data
for a related fin geometry, the offset strip fin.
Figure 14 shows curves of j and f versus ReLp for three
surface geometries:
1. The louver fin data for core no. 14 of Achalchia and
Cowell [6]
2. The offset strip fin (OSF) geometry for the same basic
dimensions as that of Achaichia and Cowell's core no. 14
3. A constant-temperature flat plate with laminar flow, whose
results are represented by j = 0.664(ReLp) -'5
The OSF (Fig. 3) geometry is sometimes described as a
"parallel louver fin." The air passes parallel to the fins in the
offset strip fin, so the concept of flow efficiency does not
apply to this geometry. Both the louver and OSF geometries
have the same dimensionless geometric variables, L p / F p =
0.507, t/Lp = 0.0455, and aspect ratio Ar = (Fp - t)/H

- 0 177 The louver fin has 0 - 22*, compared to 0 - 0


for the OSF geometry. The j and f versus Re of the offset
strip fin were predicted using the correlation of Manglik and
Bergles [ 16].
Both the louver and OSF fin geometries are intended to
provide heat transfer enhancement by periodic development
and destruction of laminar boundary layers. Their performance is compared on Fig. 14 with that for laminar flow over
a constant-temperature flat plate. Figure 14 shows that the
louver fin yields higher performance than the OSF geometry,
but both geometries fall below that of the flat plate. The slope
of the j factor for the OSF geometry closely approximates
that of the flat plate. However, the slope of the louver fin j
factor is more negative than that of the OSF. Figure 14 shows
that the j factor of the louver fin attains a maximum and then
drops as the ReLp is further reduced, contrary to the performance of the OSF geometry.
Figure 15 shows the ratios flouv/fosf and Jlouv/Josf v e r s u s
ReLp , for the louver (louv) and offset strip fins (OSF). When
the j factor of the louver fin begins to drop (Fig. 14), the
flouv/fosf ratio of the louver fin significantly increases as
Re Lp decreases Clearly, the louver fin experiences a detrimental flow phenomenon that does not occur with the offset
strip fin. We will attempt to show that the reducing flow
efficiency of the louver fin is responsible for the unfavorable
performance of the louver fin. The line labeled "flow efficiency" on Figs. 14 and 15 was calculated using Eqs. (4) and
(5).
Examination of Figs. 14 and 15 allows distinction of the
following performance differences between the louver and
OSF fins:
1. As ReLe is reduced from its highest value to the point
where Jlouv attains its maximum, flour increases faster
than fosf does as Re LP is reduced. This is apparently
because of the enhancement provided by the eddies. Note
that the flow efficiency of the louver fin decreases from
68% a t R e L p = 1000 to 35 % at Re t p = 2 0 0
2. As ReLp IS reduced to 200, an inflection point occurs on
the Jl~v curve. Below ReLp = 200, the ratio Jlo~v/Josf
decreases dramatically, and flour/fosf rapidly increases
(see Fig. 15). The flow efficiency continues to drop for
ReLp < 200, attaining ~/ = 0.18 at ReLp = 50.
Examination of the data for 15 core geometries tested by
Achalchia and Cowell [6] shows that all of the louver fin
j-factor curves display the same characteristic behavior as
that of Fig. 14. These curves show a sigmoidal shape rather
than the approximately linear shape shown by the OSF geometry. Such a sigmoidal curve contains an inflection point. It is
possible that this inflection point indicates the initiation of
significant vortex shedding. As ReLp is reduced, the intensity
of the vortex shedding apparently increases and the flow
efficiency decreases. With further reduction of ReLp, the j
factor begins to fall and the friction factor increases. Below
this ReLp, the louver fin geometry shows poor performance.
As shown by Fig. 15, this drastic behavior is initiated at
ReLp = 200, at which 7/ = 0.35. We define this as the "incipient poor performance condition" and denote the associated Reynolds number by R e ~ .
It is desirable to establish the flow condition at which the
"incipient poor performance" is initiated. We may define
this condition by fitting a straight line to the approximately
linear portion of the j-factor curve between the maximum j

Flow Structure in the Louvered Fin

VARIANT14VS.OSFWITHFLOWEFFICIENCY
.......

4.000

3.5003.000"
2.5002.000-

,I

. . . .

,~'low
efficiency

0.7

0.800

0.6
0.600

~"

0.5

KEY:

0.4

~ /

0.400

--

0.2

0.200

........

'.

100

........

1000

, ,

0ffi30*

0.0

0.000

Relp

14.2

Lp=1.0 m m

0.1

10

FPI=

- - FPIffi20O
.... FPI=33.0

0.3

j ~ / / ~ . . ~ , ~

1 .ooo.

0.500

...-'"

0.9

1.000

08

~f-louv
~-osf
~
j-louv

1.5oo-

. . . . . . . .

1.0

213

10

15

20

25

30

Ufr(m//s)

15. Friction factor and j-factor ratios (louver-to-OSF


fin) versus ReLp, and flow efficiency for the louvered fin.

16. Flow efficiency versus air frontal velocity for L p =


1.0 mm and 0 = 30*

Figure

Figure

region and the inflection point. The point at which the


experimental j factor deviates from this straight line is
indicated by the noint
labeled ReL*.
on Fig " 14 " We have
,K~1found that thts Re L. condition can be correlated with the flow
efficiency. Ten of ~chaichia and Cowell's 15 core geometries
have values of Lp/Fp within the ran,,ge of the present tests.
The value of ~ corresponding to Re L_ was predicted for the
10 geometries by using Eq. (5) Equation (5) applies because
Re~., < R e ~ , as defined, by Eq. (3). The observed values of**
Reto and the predicted values of flow efficiency (~/*) at Reto
are listed in Table 2.
The value of 7/at the "incipient poor performance" point
is in the range 35-45 %. Therefore, it can be concluded that
there is a very close relationship between the flow efficiency
7" and the incipient poor performance condition. It is recommended that the louver fin not be selected to operate at
~/< 045

(4) and (5) were used to prepare Fig. 16, which shows the
flow efficiency versus the air frontal velocity (at 20"C) with
FPI (fins/in.) shown as a parameter. This figure shows that
the flow efficiency for the 14 fins/in, geometry begins to drop
rapidly for air velocities less than 13 m/s. The figure also
indicates that lower flow efficiencies will occur for fewer than
14.2 fins/in. Below 0.45 flow efficiency, there should be
significant deterioration of the louver performance, as suggested in the previous section. This occurs at approximately 4
m / s for the 14.2 fins/in, geometry.
In general, reducing the louver pitch will increase the FPI
required to meet a specific flow efficiency. Conversely, increasing the louver pitch will reduce the required FPI to meet
a specified flow efficiency. Also, reducing the louver pitch
will increase the air frontal velocity required to meet a
specific flow efficiency and vice versa.

SIGNIFICANCE AND USEFULNESS OF


RESULTS
When the air velocity is decreased, a condition is reached at
which the performance of specific louver fin arrays will begin
to deteriorate. Equations (4) and (5) can be used to predict
the flow efficiency as a function of the louver geometry,
Ret4,, and Lp/Fp. The application of these equations is
illustrated for an actual louver geometry having a louver pitch
Lp of 1.0 mm (0.040 in.) and 30* louver angle 0. Equations
2. Value of ~/at "Incipient Poor Performance
Condition" for Achaichia and Cowell [6] Data

Table

Core
1
3
4
5
6
7
11
13
14
15

L ~ / F,,

Re~

,1"

0.69
085
0.67
0.69
0.65
0.82
0.51
0.51
0.51
0.51

25.5
255
21.5
28.5
25.5
25.5
300
28.0
22.0
22.0

210
230
200
200
145
145
185
200
195
270

0.43
0.47
0.38
0.39
0.36
0.40
0.38
0.38
0.35
0.40

CONCLUSIONS
1. A flow visualization study of the louvered fin was performed to determine the influence of the louver surface
geometry and Reynolds number on the flow structure.
2. Correlations have been developed (based on the Table 1
geometries) to predict flow efficiency as a function of
louver geometry and Reynolds number.
3. The condition of laminar flow in the wake region is
generally not realized over the Reynolds number range of
practical interest.
4. The study provides insight to understanding the requiremerits for design of more efficient louvered surface geometries.
5. Comparison of the heat transfer and friction of Achaichia
and Cowell [6] with the present correlation for ~ suggests
that the j factor will begin to drop below a log-linear
behavior at Ret. p corresponding to 0.35 < ~ < 0.45. For
Ret. p below this value, the j / f ratio is reduced.
6. Equations (4) and (5) can be used to define the air
velocity, FPI, and louver pitch combinations that will
result in low flow efficiency, and possibly low j and low
j / f performance.
This work was performed under the sponsorship of The International
Copper Association. Mr. Paul Trauger performed the flow visualization
experiments and contributed greatly to the interpretation, of the results
and preparation of this manuscript.

214

R. L. Webb and P. Trauger


NOMENCLATURE

A r aspect ratio (Fp - t ) / H , dimensionless


Cp specific heat of fluid, J/(kg K)
D ideal transverse distance traveled by streakline, Fig.
10, m
D h hydraulic diameter of fin array, m
Fp fin pitch, m
FPI fins per inch, in. - i
h heat transfer coefficient, W / ( m 2 s)
H fin height, m
L fin array depth in flow direction, m
Lp louver pitch, m
N actual transverse distance traveled by streakline, Fig.
10, m
Re~ Reynolds number based on hydraulic diameter ( =
UfrD h / ~y~,), dimensionless
ReLp Reynolds number based on louver pitch ( =
u fr L p / a p ) , dimensionless
Re t Reynolds number based on fin thickness ( = Ufrt/au),
dimensionless
Reap critical Reynolds number used in Eq. (3) ( =
UfrL p / a u ) , dimensionless
ReLp point of incipient poor performance, dimensionless
S half-length of inlet deflection louver ( = L p / 2 ) , m
St Stanton number (= h/UfrCp), dimensionless
t fin thickness, m
Ufr frontal velocity, m/s
u velocity accounting for contraction ratio ( = u f r / a ) ,
m/s

Greek Symbols
c~ mean flow angle, deg
flow efficiency, ratio of minimum flow area and frontal
area, dimensionless
0 louver angle, deg
a contraction ratio (minimum flow area-to-frontal area)
of louver array, dimensionless
v kinematic viscosity of fluid, m 2/s
REFERENCES
1. Mori, Y., and Nakayama, W., Recent Advances in Compact Heat
Exchangers in Japan, in Compact Heat Exchangers- History,
Technological Advancement and Mechanical Design Problems,
R. K. Shah, C. F. McDonald, and C. P. Howard, Eds., ASME
Syrup. Vol. HTD-Vol 10, pp. 5-16, ASME, 1980.
2. Shah, R. K., Heat Exchangers, in Handbook of Heat Transfer
Applications, 2nd ed., W. M. Rohsenow, J, P. Hartnett, and E. N.
Ganic, Eds., pp. 4-225-4-227, McGraw-Hill,New York, 1985.
3. Kays, W. M., and London, A. L., Compact Heat Exchangers, 3rd
ed., McGraw-Hill,New York, 1984.
4. Davenport, C. J., Heat Transfer and Flow Friction Characteristics
of Louvered Heat Exchanger Surfaces, in Heat Exchangers: Theory and Practice, J. Taborek, G. F. Hewitt, and N. Afgan, Eds.,
pp. 397-412, Hemisphere, Washington,D.C., 1983.
5. Davenport, C. J., Correlations for Heat Transfer and Flow Friction
Characteristics of Louvered Fin, Heat Transfer- Seattle 1983,
AIChE Symp. Ser., No. 225, Vol. 79, pp. 19-27, 1983.
6. Achalchia,A., and CoweU, T. A., Heat Transfer and Pressure Drop
Characteristics of Flat Tube and Louvered Plate Fin Surfaces, Exp.
Thermal Fluid Sci., 1, 147-157, 1988.

7. Smith, M. C., Gas Pressure Drop of Louvered Fin Heat Exchangers, ASME Paper 68-HT-27, 1968.
8. Smith, M. C., Performance Analysis and Model Experiments for
Louvered Fin Evaporator Core Development, SAE Paper No.
720078, 1978.
9. Howard, P., An Analytical Model for Heat Transfer and Friction
Characteristics of a Multi-LouveredFin Heat Exchanger, Masters
Paper, The PennsylvaniaState University, 1987.
10. Kajino, M., and Hiramatsu, M., Research and Development of
Automotive Heat Exchangers, in Heat Transfer in High Technology and Power Engineering, W. J. Yang and Y. Mori, Eds., pp.
420-432, Hemisphere, Washington,D.C., 1987.
11. Achaichia, A., and Cowell, T. A., A Finite Difference Analysisof
Fully Developed Periodic Laminar Flow in InclinedLouvre Arrays,
Proc. 2nd UK National Heat Transfer Conference, Glasgow,
Vol. 2, pp. 883-888, 1988.
12. Beauvais, F. N., An AerodynamicLook at AutomobileRadiators,
SAE Paper No. 650470, 1965.
13. Wong, L. T., and Smith, M. C., Air-Flow Phenomena in the
Louvered Fin Heat Exchanger, SAE Paper No. 730237, 1973.
14. Howard, P., PreliminaryReport on Flow VisualizationStudies on a
Two-DimensionalModel of a Louvered Fin Heat Exchanger, Penn
State Project Report, 1987.
15. Mueller, T. J., Flow Visualizationby Direct Injection, in Fluid
Mechanics Measurements, R. J. Goldstein, Ed., pp. 352-355,
Hemisphere, Washington,D.C., 1983.
16. Manglik, R. M., and Bergles, A. E., The Thermal-HydraulicDesign of the Rectangular Offset-Strip-FinCompact Heat Exchanger,
in Compact Heat Exchangers: A Festschrift for A. L. London,
R. K. Shah, A. Kraus, and D. E. Metzger, Eds., pp. 123-149,
Hemisphere, Washington,D.C., 1990.

Received December 21, 1989; revised July 25, 1990

WRITTEN DISCUSSION
Webb has presented the results of an experimental study of the
flow-directing properties of louver arrays. He has defined a
"flow efficiency" parameter as the ratio of the lateral
displacement of the flow over the first bank of louvers relative
to the ideal lateral displacement, that is, the lateral displacement that would have occurred if the flow had become fully
aligned with the louvers immediately upon entry to the array.
Correlating equations relating the experimentally determined
"flow efficiency" to Reynolds number and the array geometric parameters are presented. Webb quotes experimental data
on the performance of louvered plate fins from the literature
[6] and relates his flow efficiency parameter to the flattening of
Stanton number curves that was observed with these surfaces at
low Reynolds numbers. He aims to show how his correlating
equations can be used as predictors of curve flattening. The
paper raises a number of points that are worthy of further
discussion,
Flow Efficiency
The definition given in the paper suffers from the drawback
that the flow-alignment process can be gradual over the first
few louvers. The number of louvers required for fully
developed periodic flow is a function of Reynolds number and

Flow Structure in the Louvered Fin


the geometric parameters of the array. For example, Zhang
and Lang [17] and Antoniou et al [18] show that the fully
developed periodic condition can take five louvers to become
established. This means that, strictly speaking, "flow efficiency" as defined by Webb is valid only for an array of seven
louver rows as used in his experiments. The experimental
Stanton number data of Ref. 10 is based on samples for which
the number of louvers in a bank ranged from 10 to 18. As
recognized by Webb, it is the significant extent of this
developing flow region and the use of relatively few louvers in
his model that are principally responsible for the large
differences between his flow efficiency data and those based on
numerical analysis of the fully developed periodic situation
presented in Ref. 14. While use of the fully developed periodic
flow data has the opposite shortcoming--it ignores the developing flow region--we propose here that it is a better predictor
of behavior of these geometries.
Correlating Equations
Webb admits that his correlating equations have limited
accuracy and recommends that they not be extrapolated beyond
the experimental range of Reynolds number, that is, 4004000. However, in their subsequent application to the prediction of curve flattening, the relevant Reynolds numbers range
from 145 to 270. The implications of this are made clearer by
consideration of Fig. 1, in which the experimental data and
correlating equation for a louver angle of 20* are plotted
together.
It can be seen that the equations have limited ability to
predict likely experimental behavior at Reynolds numbers
around 200.
Flattening Prediction
As already pointed out, Webb acknowledges the numerically
determined fully developed periodic "flow efficiency" data
presented in Ref. 14. However, he did not recognize that in
Ref. 10 Achaichia and Cowell used the numerical results in
their correlating equation for the prediction of Stanton number.
Their data, which covered 15 different geometries, are
described by
Ot

St= 1.18 ~ ReLp0'58

(6)

where a is the fully developed periodic mean flow angle.


1"0

, .%

-----;,&~

c-"

identifiers

.4 ~

"6

. . . . . . .

"0---

.,..~"-% ~ " j . ~ / . ~

.....

0"~

experiment

- I

200

.----e

, = --;-0_-:--_'

/'.~.2~..'%.-"~,.,~"~/,..."~ symbol LplFp

"o-~ / ~ ~ ~ _ ~ . ~ ,

--

curve fit
I

400

1-31

0-56

(~/~9

g = 20"
I

600
Re/ll~

1(100
number

ReLp

Figure 1. Flow efficiency plotted against Reynolds number--a

comparison of Webh's experimental data with his curve fit of Eq.

(5).

215

Table 1. Incipient Poor Performance Indicators

Sample

Re~*

7/*

1
2~
3
4
5
6
7
8t
9*
l0 t
11
12 t
13
14
15

210
136
230
200
200
145
145
167
173
194
185
211
200
195
270

0.43
0.29
0.47
0.38
0.39
0.36
0.40
0.30
0.35
0.27
0.38
0.30
0.38
0.35
0.40

c~*/0
0.89
0.80
0.91
0.86
0.90
0.86
0.88
0.82
0.85
0.74
0.87
0.75
0.86
0.82
0.84

c~*/C~mx
0.95
0.92
0.96
0.94
0.94
0.93
0.93
0.94
0.95
0.94
0.95
0.93
0.95
0.94
0.95

* Omitted by Webb.
c~/0 can thus be considered a flow efficiency of similar
nature to that defined by Webb. It is a ratio of angles, whereas
Webb's parameter is a ratio of tangents of angles. The term
was deliberately introduced to the correlation procedure to
accommodate the curve-flattening behavior, and yielded an
equation valid down to a Reynolds number of 75. Achaichia
and Cowell presented the following equation to fit their
numerically obtained flow efficiency results in the region of
interest
ct = 0.936 - 243/ReLp -- 1.76 Fp/L, + 0.9950
0
0

(7)

where c~ and 0 are in degrees.


Equations (6) and (7) demonstrate that this flow efficiency is
a determinant of heat transfer behavior and that it should be
able to predict flattening.
Table 1 lists the "point of incipient poor performance"
ReLp** as identified by Webb from the Stanton number data in
Ref. 10. He omitted five of the samples because they lay
outside his experimental geometric range. All the samples have
been included here, and values of ReLp** are identified where
necessary. The values of "incipient poor performance flow
efficiency" as defined by Webb are given in the third column.
The fourth column lists the values of ct*/0 at which the
curves have started to flatten off, as calculated from Eqs. (6)
and (7). This parameter appears to be a somewhat better
predictor than ~/*. a*/O values cover the range 0.74-0.91 (a
ratio of 1.23) for all 15 samples, while ,/* goes from 0.35 to
0.47 (a ratio of 1.34) for the samples covered by Webb.
Nonetheless, two samples in particular show significant
deviation from the 0.8-0.9 range within which most of the
samples lie. Consideration of the significance of Eqs. (6) and
(7) leads to a better curve-flattening predictor.
For any particular geometry, the mean flow angle approaches a limiting value Ctm~x as Reynolds number is increased. Equations (6) and (7) mean that the Reynolds number
at which the limiting value is reached is also the Reynolds
number at which the straight-line region of the log St/log ReLp
curve begins. The appropriate parameter for the identification
of curve flattening is therefore c~*/C~m~xrather than ot*/0. This

216

Ralph L. Webb

is confirmed by consideration of the values presented in the


final column of Table 1. [C~m~is calculated by setting ReLp to
infinity in Eq. (7).] The values all lie within the range 0.920.96. The mean is 0.94 with a standard deviation of 0.01.
These values confirm that Eqs. (6) and (7) can be used for
predicting Stanton number curve flattening. Their validity
covers the wide range of Lp/Fp values of the experimental data
in Ref. 10, despite the fact that they extend beyond the range
used in the numerical determination of the fully developed
periodic flow angles. What is more, the method arrives at a
limiting value that matches with common sense; one would
expect to detect flattening once the curve had dropped to less
than 95 % of its extrapolated straight-line value.
It is worth noting that although the narrow range of a*/am~x
values suggest that the parameter is a very precise predictor of
flattening, the curves of t~*/C~ax against Reynolds number
have small slope as they approach unity, and for any geometry
the predicted value of critical Reynolds number will be quite
sensitive to the choice of critical parameter value. On the other
hand, neither can the precise point at which an experimental
curve starts to deviate from a straight line be identified with
great accuracy.
Conclusion

Webb has identified that it can be useful to know the value of


Reynolds number at which a louvered fin begins to demonstrate deterioration in performance due to changes in the flow
deflection characteristics. He has proposed the use of experimental data for developing flow for this purpose. It is
recognized that developing and fully developed flow behavior
of louver arrays are interrelated, and as such, by suitable
manipulation, information on either should be able to predict
Stanton number curve flattening. However, it is suggested here
that fully developed periodic data [14] lead to a better
predictor, principally because of the sounder physical basis of
the equations involved.
REFERENCES
17. Zhang, H., and Lang, X., The Experimental Investigationof Oblique
Angles and Interrupted Plate Lengths for Louvered Fins in Compact
Heat Exchangers, Exp. Thermal Fluid Sci., 2, 100-106, 1989.
18. Antoniou, A. A., Heikal, M. R., and Cowell, T. A., Measurements
of Local Velocity and Turbulence Levels in Arrays of Louvered Plate
Fins, Heat Transfer 1990, 4, 105-110, 1990.
T. A. Cowell and A. Achaichia
Department of Mechanical and Production Engineering
Brighton Polytechnic
Brighton BN2 4GJ, UK

AUTHOR'S REBUTTAL
I am appreciative of the additional material provided by Drs.
Cowell and Achaichia (C & A). The motivations for the
present work were the photographs of Kajino and Hiramatsu
[10], which showed that the flow does not necessarily follow
the louver angle, and the numerical predictions of Cowell and

Achaichia [6]. My initial objective was to experimentally


establish the actual flow pattern in a typical louver array for a
range of geometric parameters and Reynolds numbers. If the
test sections were geometrically scaled to the louver pitch
dimensions used in radiator cores, the airflow depth would be
27 mm for Lp = 1.5 mm, and 18 mm for Lp = 1.0 mm. These
flow-depth dimensions are within those currently used for onerow cores. As stated in the paper, the dye was injected into the
flow-straightener straws upstream from the test section. The
measured lateral deflection of the dye was used to define the
mean flow angle over the water flow distance. I believe that the
empirical correlations developed are reasonably accurate for
the range of test conditions.
C & A use the parameter a/O rather than my "flow
efficiency" 7. These terms are related by the simple expression
tan a
=
=rl tan 0 0

(8)

where 0 is the louver angle and c~ is the mean flow angle.


Either expression is valid to define the characteristics of the
mean flow. The approximation ~ = c~/0 is valid within 3 % for
0 = 20* for 0.2 < 71 < 0.9.
The C & A predicted values are for fully developed periodic
flow. However, the present experimental data are for an array
typical of those used in 1-row radiators. It is not expected that
"fully developed" flow would prevail in such arrays 18-27
mm deep. Hence, we question the applicability of the C & A
predictions to actual 1-row radiators. Further, C & A have not
validated their predicted "fully developed" values by experiments. Figure 10 in the present paper shows substantial
disagreement between their predicted t~/0 and the flow
efficiency measured in the present work.
I was intrigued by the maxiumum value shown by the Staton
number data of C & A and thought that the dropoff of flow
efficiency at low Reynolds numbers might be responsible for
the behavior. Hence, I endeavored to explain why this
behavior exists. The flow efficiency is an indicator of the
fraction of the flow that is parallel to the louvers. A flow
efficiency less than 1 means that part of the flow bypasses the
louvers and flows in the "duct region" between the louvered
passages. It is reasonable to expect that if the flow fraction
passing over the louvers decreases, the j factor should also
decrease. I sought to show that the dropoff of the j factor can
be related to the flow efficiency. I was able to relate the
approximate value of the ReLp associated with the "incipient
poor performance" to the flow efficiency. This is presented in
Table 2.
In their discussion, C & A present their Table 3 as a
proposed alternative to Table 2. The value or*/0 is approximately equal to the flow efficiency, as shown by Eq. (8) above.
However, comparison of Table l columns 3 and 4 show that
the flow efficiency (from the present correlation) is substantially smaller than the values shown in column 4. We conclude
that the predicted values of ct*/O, based on the C & A
numerical solution, does not agree with the results of the
present experiments.
C & A propose that the value tx*/'Vr~ (column 5 of Table 1)
is a better predictor of the "point of incipient poor performance." This means that when the flow angle falls only 5%
below its maximum (limiting) value, the j-factor curve will
drop as ReLpis reduced. This is contrary to the findings of this

Flow Structure in the Louvered Fin


work. Consider, for example, core 4 in Table 2. The louver
angle is 21.5 , which is very close to the 20 data plotted in
Fig. 8a. At Lp/Fp = 0.66, one reads ~/m~ = 0.88. The
experimental value of R e ~ (point of incipient poor performance) is 200 (Table 3). Extrapolation of the Lo/Fp data points
in Fig. 8a to ReLp = 200 suggests that 7/will be less than 0.5
(the correlation gives ,/ = 0.38). Using the C & A criterion of
a*/Otm~ = 0.95 would predict the incipient point of poor
performance at ~/ = 0.95 x 0.88 = 0.84, at which ReLp =

217

1300. This value is substantially higher than the ReLp at which


the j factor begins to drop.
I agree that a more precise correlation than Eq. (5) is desired
for prediction of ~ with ReLp < Re*p. However, for the range
of Lp/Fp shown in Table 2 (0.51-0.82), Eq. (5) does a pretty
good job.
R. L. Webb

You might also like