Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Proceedings of IMECE2008

2008 ASME International Mechanical Engineering Congress and Exposition


Proceedings
November 2-6, 2008, Boston, of IMECE2008
Massachusetts, USA
2008 ASME International Mechanical Engineering Congress and Exposition
October 31-November 6, 2008, Boston, Massachusetts, USA

IMECE2008-66113

COMPARISON OF PLATE FIN COMPACT HEAT EXCHANGER PERFORMANCE

Ibrahim Khalil, Ahmad Abu Heiba and Robert Boehm


Center for Energy Research
UNLV Box 454027
Las Vegas, NV 89154-4027

ABSTRACT
Plate fin heat exchangers (PFHE) are characterized by very
close temperature approaches and high thermal effectiveness,
large heat transfer area per unit volume, low weight per unit
transfer and possibility of heat exchange between many process
streams. These advantages are only limited by operating fluid
temperatures and pressures.
The main target of this paper is to study the performance of
plate fin compact heat exchangers and to provide full
explanation of previous comparison methods of compact heat
exchanger surfaces (plain, strip, louvered, wavy, pin, perforated
and vortex) used in plate fin compact heat exchangers. We Figure 1.b Strip Fin Surface (Hall, 2003)
generalize these methods to identify the advantages and
disadvantages of each type of geometry (more than sixty
geometries studied) based on required size, entropy generation,
pumping power, weight, and cost. The effect of using different
surfaces on each side of the heat exchanger and design
recommendations are also discussed.

INTRODUCTION
A plate-fin heat exchanger (PFHE) is a type of compact
exchanger that consists of a stack of alternate flat plates called
parting sheets and corrugated fins brazed together as a block.
Different types of surfaces may be used as cores of plate fin
compact heat exchangers such as plain, strip, louvered, wavy, Figure 1.c Louvered Fin Surface (Hall, 2003)
pin, perforated, and vortex. All these surfaces are shown in
Fig.1-a to Fig. 1-f.

Figure 1.a Plain Fin Surface (Hall, 2003)


Figure 1.d Wavy Fin Surface (Hall, 2003)

1 Copyright © 2008 by ASME


and broader design considerations for surface size, shape and
weight.
Shah and Sekulic (2003) presented the procedures of rating
and sizing problems. They used the mean temperature
difference of the fluid on each side of the heat exchanger in
order to calculate the fluid properties assuming the uniformity
of thermo-physical properties.
Sekulic (2005) offered a very clear methodology for
calculating core dimensions of a compact heat exchanger. He
considered the analytical complexity of implemented
calculations. The most intricate basic flow arrangement
situation in a single pass configuration would be a crossflow in
Figure.1.e Pin Fin Surface (Hall, 2003) which fluids do not mix orthogonally to the respective flow
directions. Calculations were executed using an explicit step-
by-step routine based on a set of known input data and this is
provided in the problem formulation.
Tagliafico (1996) provided a comparative study of entropy
generation of many surfaces scaled by that of a reference
configuration (a parallel-plate channel). This considered
irreversibility analyses as an important factor in determining
the operating costs of the heat exchanger.
Fiebig (1995) provided a comprehensive study of the use
of vortex generators in either tube fin or plate fin compact heat
exchangers. He compared the performance of transverse vortex
generators and longitudinal vortex generators by describing the
mechanism of heat transfer enhancement due to using vortex
generators. A comparison was made of the performance of high
Figure 1.g Perforated Fin Surface (Stevens, 2001)
performance surfaces (louvered, strip) used in plate type
compact heat exchangers. Jacobi and Shah (1995) discussed the
recent progress of vortex-induced heat transfer enhancement.
This included the theoretical basis for passive and active
implementation.
Jacobi and Shah (1998) studied the behavior of air flow in
complex heat exchangers passages with a focus of boundary
layer development with turbulence, both span wise and stream
wise. Each of these flow features is discussed for the plain,
wavy, and interrupted passages found in contemporary heat
Figure 1.f Vortex Generator Surface (Brockmeier, exchanger design.
1993) Hall (2003) discussed air-cooled compact heat exchanger
design using published data from Kays and London (1984), and
Kays and London (1984) obtained the relation between this contains measured heat transfer and pressure drop data for
Colburn factor (j) and mean friction factor (f) for each compact a variety of circular and rectangular passages.
heat exchanger surface. Selection of particular surfaces to be
applied from the many available is the first design step. NOMENCLATURE
The selection criteria for these surfaces are discussed in
Shah (1978). A general method for comparison of compact heat
A 2 2
transfer surfaces has been recently proposed by Cowell (1990). Total transfer area of one side of exchanger, ft or m
This method provides statements of the relative merits of a Plate thickness, ft or m
different heat transfer surfaces by comparing relative pumping Ab 2 2
powers and relative hydraulic diameters. Nunez (1999) Base plate area, ft or m
developed a thermal hydraulic model that represents the Ac 2 2
Free-flow area of one side, ft or m
relationship between pressure drop, heat transfer coefficient and
Af 2 2
exchanger volume. Total fin area on one side, ft or m
Hesselgreaves (2001) has attempted to provide a treatment Afr 2
Frontal area of one side, ft or m
2
that goes beyond dimensionless design data information. In
Ar The ratio of fin area to total heat transfer area on one
addition to the basic design theory, he includes descriptions of
side
industrial compact heat exchangers, thermodynamic analysis

2 Copyright © 2008 by ASME


b Plate spacing, ft or m f Fin Efficiency
Cc Flow stream capacity rate of cold side fluid
Ch Flow stream capacity rate of cold side fluid o Total surface effectiveness
cp Specific heat at constant pressure
 Ratio of free-flow to frontal area of one side of
exchanger
Dh Hydraulic diameter of any internal passage, ft or m
f Mean friction factor  Dynamic viscosity, Pa s
fo Mean friction factor for reference surface  Density, kg/m3
G Flow stream mass velocity, kg/m2 s
h Convective heat transfer coefficient, W/ m2 K OBJECTIVE OF THE STUDY
j Colburn factor The objective of this work focuses on three primary tasks.
jo Colburn factor for reference surface The first is sizing the compact heat exchanger core to specify
the core dimensions for different high performance surfaces
k Thermal conductivity, W/mK
(more than sixty five surfaces), the second is calculating the
Kc Contraction loss coefficient at entrance pumping power required, and the final one involves
Ke Expansion loss coefficient at exit determining the entropy generation in order to make the
L Flow length on one side, ft or m selection of the best surface more apparently for certain design
l Fin length from root to center, ft or m constraints.
Lstack No flow length, ft or m
m Correction factor for friction factor
SIZING PROCEDURES
N Entropy factor
The procedure suggested by Sekulic (2005) was
n Correction factor for Colburn factor generalized in this paper to obtain the core dimensions of more
nf Number of fins per meter than sixty-five surfaces. This procedure follows a somewhat
NTU Number of heat transfer units modified sizing procedure derived from the routine put forth by
Nu Nusselt number Shah and Sekulic (2003).
The design procedure for a sizing problem features two
Pcc Pumping power on cold side, W distinct parts of the calculation. The first one delivers the
magnitude of the thermal size of the core, expressed as a
Phh Pumping power on hot side, W product of the overall heat-transfer coefficient and the heat-
transfer area UA based on "effectiveness number of heat-
Pr Prandtl number transfer units" method. The second part of the calculation is
Q Heat transfer rate, W devoted to the determination of actual overall physical
Re Reynolds number dimensions of the core.
This feature of the calculation is only one aspect of the
Reynolds number for reference surface design methodology that ultimately leads to an iterative
Re o
calculation sequence. The main reason for an iterative
rh Hydraulic radius, ft or m
procedure is the constraint imposed on pressure drops.
Sp Magnitudes of pressure drops must be obtained from the
Entropy generation rate per unit exchanger length
hydraulic part of the design procedure. The hydraulic design
St Stanton number part of the procedure cannot be decoupled from the thermal
Ti Inlet temperature, K part.
The following are the key equations in Sekulic (2005)
Tw Wall temperature, K procedure:
Overall thermal conductance, W/m2 K
(1)
U UA  NTU  C1
V Total exchanger volume, ft3, m3
Where: C1 is the heat capacity of the weaker fluid (lower heat
VG Vortex generator
capacity)
W Mass flow rate, kg/s
12
β Ratio of total area on one side to volume between plates Gc   j  p  2 pin o  m  (2)
   23
ΔP Pressure drop on one side, kPa  f  pin  NTU Pr  c
 Fin thickness, ft or m
 Exchanger Effectiveness 12
(3)
G h   j  p  2 p in o  m 
   23 
 f  p in  NTU Pr  h

3 Copyright © 2008 by ASME


Re c  G c D h,c (4) m C  NTU
p c (19)
Ac  Ah 
c U
Re h  Gh Dh ,h (5) The assumption of using the same geometry on both sides
h will make the equation (18) easier to solve to obtain the value
NTU c C * of overall heat transfer coefficient ( U ).
Tw  Th , ref  Tc,ref
NTU h (6)
NTU c C *
Ac , c  m c (20)
1 Gc
NTU h
n Ac , h  m h (21)
j c ,corr  j c  T w 
 (7) Gh
T 
 c , ref 
14
The free-flow areas on the cold fluid side and hot fluid
   sides are determined from the definition of the mass velocity.
n  0 .3   log  T w  (8)
10  
  T c , ref 
 c   h  b D h (22)
n 8b  a 
j h,corr  j  Tw 
  jh (9)
h 
 Th, ref  Equation (21) was obtained from core geometry relations
for same geometries on both sides (Kays and London year).
The hot fluid experiences cooling conditions. The flow
regime is in the laminar region. Therefore, the exponent n = 0 Afr,c 
Ac ,c (23)
(Shah and Sekulic, 2003, table 7.12, p.531) is used. c

m A fr , h  Ac ,h (24)
 T  (10) h
f c ,corr  fc  w  , m  0.1
 Tc ,ref 
 
m
 T  (11) The frontal area on the cold fluid side and the hot fluid side
f h ,corr  f h  w 
 Th ,ref  is determined from the relation between porosity and free-flow
 
area.
hc  jc ,corr
Gc c p.c (12)
Prc2 3 Lc  Dh Ac (25)
The heat-transfer coefficients for the cold fluid (hc) and the 4 Ac ,c
hot fluid (hh) are determined from the definition of the Colburn The fluid flow length on the cold or on the hot side
factor. represents the principal core dimension in this direction.

hh  jh ,corr
Gh c p.h (13) Lh  Dh Ah (26)
Prh2 3 4 Ac , h

 f , c  tanh( ml) c (14)


ml c L stack  A fr ,c (27)
Lh

 f , h  tanh( ml) h (15) The core dimension in the third direction (no flow length)
ml h
can be calculated by using the frontal area of either the cold
fluid or the hot fluid.
The thermal conductivity of the fin is assumed to be 200 After calculating all dimensions and areas, pressure drop
W/mK for an alloy at the given temperature in Equations (14) through the core must be calculated to check for pressure drop
and (15). constraints.
Af (16)
 o ,c  1  (1   f ,c )  p  G c2  L i     (28)
A    (1    K c )  f
2
 2 i  1  (1   2  K e ) i 
 p i  c 2( p in  in ) c  rh  m  o   o 
c
A (17)
 o , h  1  (1   f ,h )
Af
The relative pressure drop calculations require
(18) determination of both entrance and exit pressure loss
1 coefficients Kc, Ke. These coefficients can be determined from
U   1  Ac Ah  Kays and London (1984, Figures 5.4, 5.5 p.113-114). The
 
 ( o h) c ( o h) h 

4 Copyright © 2008 by ASME


values of pressure drop coefficients depend on surface
geometry, porosity and Reynolds number.

 p  G 2h  Li     (29)
   (1    K c )  f
2
 2 i  1  (1  2  K e ) i 
 i h
p 2 ( p 
in in h 
)  r 
h m 
 o   h
o

From the input data, an allowed pressure drop of 5 kPa on


the cold side and 4.2 kPa on the hot side is assumed (Sekulic,
2005). If the desired conditions are not satisfied on either side,
this prompts a need to reiterate the calculation. Anew value of
the mass velocity is used. In the first iteration, the mass
velocity was calculated by using the first approximation based
on a weak dependence of j/f on the Reynolds number.

1
  PConstraint  2
2(Pin  in  
Figure 2. Total Area Comparison of Different Surfaces
  Pi  c (30)
Gc  1
 L i     2 Table 1 illustrates the best surface in each category used in
(1    K c )  f
2
 2 i 1  (1   2  K e ) i  this paper For example, surface 46.45T has the highest
 rh  m   o   o  c
compactness (largest area in small volume) and it has also the
1 highest overall heat transfer coefficient ( U ) for the plain
  PConstraint  2
2(Pin in   surfaces.
  Pi  h (31)
Gh  1 Table 1 High Compactness and Conductance Surfaces
 Li     2
Geometry High High Conductance
(1  Kc )  f
2
 2 i 1  (1 2  Ke ) i 
 rh  m  o  o  h Compactness
Plain 46.45T 46.45T

The new iteration loop starts with the determination of the Louvered 1/2-11.1 1/4(b)-11.1
set of new mass velocities. These values will be used to
Strip 1/10-27.03 1/10-27.03
calculate the refined values of Reynolds numbers (Eq.4, 5). The
new mass velocities G should be calculated from the exact Wavy 17.8-3/8 11.5-3/8
expression for the pressure drop (Eq. 29, 30), then repeating the
procedures from equation 7 to 31. The pressure drop constraint Pin PF-10(F) PF-10(F)
should be satisfied in order to calculate the final dimensions of
Perforated 13.95(P) 13.95(P)
the core.
Vortex Vortex Vortex
SIZING RESULTS
Figure 2 illustrates the various types of surfaces
considered. Type A is 11.11a plain, type B is 11.94T plain, type Table 2 illustrates the ranking of best surfaces in each
C is 3/32-12.22 strip, type D is 3/16-11.1 louvered, and type E category It is clear that the vortex generator geometry is the
is the vortex generator geometry. The results represented in best option because it gives highly compact cores with high
Figure 1 (obtained from this study) match very well with the conductance for the same heat duty.
results obtained by Brockmeier (1993), a more limited study
than that presented here using the same designation of the Table 2 High Compactness and Conductance Ranking
surfaces studied High Compactness Ranking High Conductance Ranking
1-Vortex 1-Vortex
2-46.45T Plain 2-PF-10(F) Pin
3-1/10-27.03 Strip 3-1/10-27.03 Strip
4-17.8-3/8 Wavy 4-11.5-3/8 Wavy
5-13.95(P) Perforated 5-1/4(b)-11.1Louvered

5 Copyright © 2008 by ASME


6-1/2-11.1 Louvered 6-13.95(P) Perforated ENTROPY GENERATION ANALYSIS
Entropy generation (irreversibility) can be used as the
7-PF-10(F) Perforated 7-46.45T Plain quantitative measure of the quality of energy transformation in
the heat exchangers (Sekulic, 1990). The following gives the
main equations used in the entropy generation comparison
(Tagliafico, 1996):
PUMPING POWER CALCULATIONS
Most of surfaces studied satisfied the pressure drop
SP   Q Dh 
2 2
1  2L 3 A 
requirements on both sides of the heat exchanger so the value  c 
f Re3 (33)
 4kTi2 Ac L  j RePr1 / 3  Ti  2 D 4 
of pumping power required to push the fluid may be considered    h 
one of the valuable criteria for judging the selection of compact
N  S P / S P, o (34)
heat exchangers. Of course, the lower pumping power required
to push the same amount of fluid, the better the surface is.
The following equation can be used to calculate the The dimensionless variable B:
pumping power on each side:
 3 kTi V 2  (35)
W  P B  6 
P (32) 8 2 Q 2 b 

The pumping power on the hot side will always be higher The entropy factor can be calculated from:
than the cold side because of the effect on temperature on
 1  1
density.  2   16 4 B  Re 3
 4   j Re Pr 1 / 3
Table 3 illustrates the best surfaces in each category that N (36)
1
require the lowest pumping power to push the same amount of  B  o Re 3o
jo Re o Pr 1 / 3
fluid on both sides (the mass flow rate is the same for the both
sides) for the same heat duty.
Where
Table 3 Lowest Pumping Power Consumption Surfaces   b / Dh
Geometry Type Cold Side Hot Side The constraints considered (Tagliafico, 1996) involve
keeping the heat exchanger plate spacing b and volume V fixed
Plain 3.01 3.01 and assuming the same mass flow rate in and heat transfer Q
Louvered 3/8-6.06 1/2-6.06 for both the compact and reference configurations. The
Reynolds numbers Re and Reo are linked through the
Strip 1/2-11.94(D) 1/2-11.94(D) relationship:
Wavy 17.8-3/8 11.5-3/8 Re  Reo / 2 (37)
Perforated 13.95(P) 13.95(P)
The above method of comparing entropy generation was
Vortex Vortex Vortex applied to more than sixty five surfaces in this study. Figures
3-7 show the resulting variations.

From the results obtained, generally, it is very clear that the


Entropy Factor of Strip Surfaces 1
plain fin is the best geometry considering the pressure drop and
required pumping power On the other hand the strip fin 0.3
1/4(s)-11.1S
requires the largest pumping power. The main factor that 0.25 1/8-15.2 S
influences the value of pumping power is the hydraulic 0.2 1/8-13.95 S
diameter. Although plain surfaces require smaller pumping 1/8-15.61 S
0.15
N

power than other surfaces, the surface 46.45T plain 1/8-19.86 S


(Dh=0.805mm) requires more pumping power than most of 0.1 1/9-22.68 S
other surfaces. We conclude that the lower value of hydraulic 0.05 1/9-25.01 S
diameter, the higher will be the expected pumping power. The 0
1/9-24.12 S
calculated values for pin surfaces were excluded from this 0 5000 10000 15000
1/10-27.03 S
study because most of these surfaces failed to satisfy the 1/10-19.35 S
Reo
pressure drop requirements. Figure 3. Entropy Generation Factor for the First
Group of Strip Surfaces

6 Copyright © 2008 by ASME


0.18
Figure 7 shows that the entropy generation rate will
0.16 increase rapidly when Reynolds number increases so using
3/32-12.22
vortex generator geometry is not recommended in the case of
0.14 1/2-11.94(D) very high Reynolds numbers.
1/4-15.4(D)
0.12 1/6-12.18(D)
1/7-15.75(D)
N

0.3
0.1 1/8-16.00(D)
1/8-16.12(D) 0.25
0.08 1/8-19.82(D)
0.2
1/8-20.06(D)
0.15

N
0.06
0.1
0.04
0.05
0 2000 4000 6000 8000 10000 12000
Reo 0
1709 2051 2735 3418 5128 6837 8546 10260
Figure 4. Entropy Generation Factor for the Reo
Second Group of Strip Surfaces
Figure 7. Entropy Generation Factor for Vortex Surface
The strip fin category mentioned in Kays and London
(1984) contains 21 surfaces so these surfaces were divided into Table 4 shows the lower and the higher entropy generation
two groups and represented in Figures 3 and 4. extremes for different surfaces. Pin surfaces were excluded
from this table because they cause high pressure drop and most
0.24
of that type do not have sufficient data for comparative
purposes.
3/8-6.06L
0.22 3/8-6.06(a)L
1/2-6.06L Table 4 Entropy Generation Classification
0.2 3/8-8.7L Min
3/8-8.7(a) Average Max Entropy
0.18 3/16-11.1L Geometry Entropy
Entropy Surface
Surface
N

1/4-11.1L
0.16 1/4(b)-11.1L
3/8-11.1 Plain 0.191978 30.33T 6.2
0.14 3/8(b)-11.1 Strip 0.096277 1/8-13.95 1/10-19.35
1/2-11.1L

0.12 3/4-11.1L Wavy 0.055898 17.8-3/8 11.5-3/8


3/4(b)-11.1 Louvered 0.160464 3/16-11.1 1/2-6.06 L
1/2-6.06(a)
0.1 Vortex 0.083625 V.G V.G
0 2000 4000 6000
Perforated 0.1795 13.95(P) 13.95(P)
Reo

Figure 5 Entropy Generation Factor for Louvered Surfaces


COMBINATION OF TWO DIFFERENT SURFACES
The combination between two different surfaces inside the
0.09 compact heat exchanger core is a promising technique that can
be used to reduce the total required volume and cost for the
0.08
same heat transfer load. However at the same time, there will
be a significant increase in the pressure drop and consequently
0.07
11.44-3/8
the required pumping power as shown in Table 5. In addition,
0.06 11.5-3/8 there will probably be an increased manufacturing cost. The
N

17.8-3/8 table illustrates the value of using different surfaces on both


0.05 sides of the heat exchanger instead of using the same geometry
on the both sides. Surfaces 3/4-11.1 (louvered) and 1/4(s)-11.1
0.04 (strip) were selected for this comparison because both of them
have the same hydraulic diameter, plate spacing, fin pitch, area
0.03
density, and the ratio of fin area to total area. Although this
0 5000 10000 15000
technique is promising, there is 30% increase in pressure drop
Reo
on the hot side.
Figure 6. Entropy Generation Factor for Wavy Surfaces

7 Copyright © 2008 by ASME


Table 5 Effect of Combination of Two Different Surfaces Enhancement through the use of Longitudinal Vortices: A
Geometry Type Volume Pc Ph Lstack Review of Recent Progress. Experimental Thermal and
Fluid Science, Vol.11, pp. 295-309
3/4-11.1 L (Both) 3 4.39 kPa 4.11 kPa 4.65 m
0.665 m Fiebig, M., (1995). Vortex Generators for Compact Heat
1/4(s)-11.1 S (Both) 3 4.38 kPa 4.14 kPa 5.6 m
Exchangers. Journal of Enhanced Heat Transfer, Vol.2,
0.576 m
pp.43-61.
3/4-11.1L and 1/4(s)-11.1 3 4.99 kPa 6.07 kPa 1.19 m
0.232 m Hall, G., Marthinuss, J., (2003). Air Cooled Compact Heat
Exchanger Design for Avionics Thermal Management.
Proceedings of interpack03, International Electronic
CONCLUSIONS AND RECOMMENDATIONS
Packing Technical conference. Pp. 705-711.
The main purpose of sizing process is to obtain an accurate
Hesselgreaves, J. E. (2001). Compact Heat Exchangers
assessment of the surfaces used in compact heat exchangers.
Selection, Design and Operation (1st Ed.). Pergamon.
Some surfaces fail to satisfy the pressure drop on one side so
Jacobi, A. M., Shah ,R. K. (1995). Heat Transfer Surface
these surfaces are excluded from the end selection; also the
Enhancement through the Use ofLongitudinal Vortices: A
purpose of studying the entropy generation and pumping power
Review of Recent Progress,” Experimental Thermal and
is to refine the selection of the best surface. The best
Fluid Science, Vol. 11, pp. 295-309.
performing surface may not be an optimum heat exchanger
Kays, W. M., and London, A. L. (1984). Compact Heat
surface for a given application. Hence, there is no need to “fine
Exchangers (3rd Ed). McGraw Hill
tune” the selection of a surface individually.
Picon-Nunez, M., Polley, G. T., Torres-Reyes, E., Gallegos-
The main result of this paper is to report a study of the
Munoz, A. (1999). Surface Selection and Design of Plate
performance of heat exchanger surfaces on both sides. In all of
Fin Heat Exchangers. Journal of Applied Thermal
previous comparison methods reviewed, the surface on only
Engineering, Vol.19, pp. 917-931.
side of the exchanger is considered
Sekulic, D. P., (1990). The Second Law Quality of Energy
The sizing results show that the vortex generator geometry
Transformation in a Heat Exchanger. ASME Journal of
is a promising surface to be selected as a core surface because it
Heat Transfer, Vol.112, pp.295-300.
can transfer the required heat rate with very small volume. It
Sekulic, D., (1990). A Reconsideration of the Definition of
also satisfies the pressure drop requirements. Although this
Heat Exchangers. Int. Journal of Heat Mass Transfer,
geometry requires significant pumping power due to boundary
Vol.33, No.12, pp2748-2750.
layer separation, it remains the best selection for the compact
Sekulic, D. P.(2005).Heat Transfer Calculations, ed. Kurtz, M.,
heat exchanger designer.
Chapter 29, McGraw Hill.
The sizing results obtained are matched with all former
Shah, R., Sekulic, D. P.(2003). Fundamentals of Heat
comparative research for the selection of the optimum surface.
Exchangers Design. John Wiley & Sons, Inc.
These results give clearer vision of these surfaces, their sizes,
Shah, R. K., (1978). Compact Heat Exchangers Surface
pumping power requirements and entropy generation.
Selection Methods. Int .Heat Transfer Conf., Toronto,
Vol.4, pp.193-199.
ACKNOWLEDGMENTS
Shah, R. K., (1983). Heat Exchanger Basic Design Methods. In:
The authors appreciate the partial support of this work by
S. Kakac, R.K. Shah and A.E. Bergles (Eds.), Low
the National Renewable Energy Laboratory.
Reynolds Number Flow Heat Exchangers, Hemisphere,
Washington, DC, pp. 21–72.
REFERENCES
Shah, R. K., (1999). Compact Heat Exchangers and
Brockmeier, U., Fiebig, M., Guntermann, T., and Mitra, N.K.
Enhancement Technology for the Process Industries:
(1989).Heat Transfer Enhancement in Fin-Plate Heat
Proceedings of the International Conference on Compact
Exchangers by Wing Type Vortex Generators. Chemical
Heat Exchangers and Process Industries Held at the Banff
Engineering Technology, Vol.12, pp.288-294.
Center.
Brockmeier, U., Guntermann, T., and Fiebig, M., (1993).
Soland, J. G., Mack, W. M., and Rohsenow, W. M. (1978).
Performance Evaluation of a Vortex Generator Heat
Performance Ranking of Plate Fin Heat Exchangers
Transfer Surfaces. International Journal of Heat and Mass
Surfaces. ASME Journal of Heat Transfer, Vol.100,
Transfer, Vol.36, pp.2575-2587.
pp.514-519.
Brockmeier, U., Guentermann, TH., Fiebig, M.(1993).
Stevens A., Adderley, C. I. and Cool, T., (2001). Using Genetic
Performance Evaluation of Vortex Generator Heat Transfer
Algorithms to Optimize Heat Exchange. Proceedings of
Surface and Comparison with Different High Performance
UKSIM Conference.
Surfaces. Int. Journal of Heat and Mass Transfer, Vol.36
Tagliafico, L., Tanda, G., (1996). A Thermodynamic Method
(10), pp.2575-2587. for the Comparison of Plate Fin Heat Exchangers
Cowell, T. A., (1990). A General Method for the Comparison Performance. Journal of Heat Transfer, Vol.118, pp. 805-
of Compact Heat Transfer Surfaces. Journal of Heat 809.
Transfer, Vol.112, pp. 289-294.

8 Copyright © 2008 by ASME

You might also like