You are on page 1of 28

Version of Record: https://www.sciencedirect.

com/science/article/pii/S0360319921040982
Manuscript_4fe5d095b0503d6772841c01e88247b7

Novel structural designs of fin-tube heat exchanger for


PEMFC systems based on wavy-louvered fin and vortex
generator by a 3D model in OpenFOAM
Wei Luoa,1, Zirong Yangb,1, Kui Jiaoa, Yanyi Zhang b,*, Qing Dua,*

1. Equal contribution

a. State Key Laboratory of Engines, Tianjin University, 135 Yaguan Road, Tianjin, 300350, China

b. China Automotive Technology and Research Center Co., Ltd., 68 East Xianfeng Road, Tianjin,

300300, China

* Corresponding author: duqing@tju.edu.cn, zhangyanyi@catarc.ac.cn

Abstract
Heat management is crucial to the stable and high-efficiency operation of proton
exchange membrane fuel cell (PEMFC) system. However, fin-tube heat exchangers
(FTHE) of traditional internal combustion engine vehicles require further optimizations
to be applicable to PEMFC vehicles. In the paper, a three-dimensional steady-state
radiator model is developed in OpenFOAM to investigate three novel structural designs
based on wavy-louvred (WL) fin and vortex generators (VGs). The established model
has been carefully validated against experimental data and correlation reference. To
comprehensively evaluate radiator performances, the air side heat transfer coefficient,
pressure drop, outlet air temperature, heat flux, and JF factor are adopted. It is found
that the FTHE with L-VGs has the highest heat transfer coefficient while the FTHE
with WL-VGs has the highest pressure drop. The temperature, velocity, and pressure
distribution are further demonstrated to reveal performance enhancement mechanisms.
It is seen that the heat exchangers with additional VGs produce two sections of high-
temperature wakes near the wall, which not only promotes the heat convection but also
contributes to the heat exchange in the nearby area. Meanwhile, a low-speed vortex
zone behind VGs appears and generates longitude vortex, making the air stream stay
longer at fin surfaces. The air flow in FTHE with WL is not as much separated as the
conventional FTHE since the zigzag wavy louver restricts flow separation. The paper

© 2021 published by Elsevier. This manuscript is made available under the Elsevier user license
https://www.elsevier.com/open-access/userlicense/1.0/
gives valuable suggestions for cooling capability improvement and radiator volume
diminution.

Keywords: PEMFC system; OpenFOAM; Fin-tube heat exchanger; Wavy-louvered fin;


Vortex generator;

Nomenclature
A0 total heat transfer area (m2)
Ac minimum area of cross section (m2)
Af fin front area (m2)
Cp specific heat capacity (J kg-1 K-1)
Dh hydraulic diameter (m)
E energy (J)
f friction factor
F fin structural size (m)
h heat transfer coefficient (W m-2 K-1)
H height (m)
j Colbum factor
k heat conductivity coefficient (W m-1 K-1)
L louver structural size (m), length (m)
Nu Nusselt number
Pr Prandtl number
p pressure (Pa)
∆ pressure drop (Pa)
Q amount of heat (W)
q mass flow rate (kg s-1)
Re Reynold number
t time (s)
T temperature (K), tube structural size (m)
∆ logarithmic mean temperature difference (K)
u velocity in the direction of x axis (m s-1)
velocity in the minimum cross-sectional flow area (m s-1)
U velocity (m s-1)
v velocity in the direction of y axis (m s-1)
w velocity in the direction of z axis (m s-1)
W wavy fin structural size (m)
x in the direction of x axis
y in the direction of y axis
z in the direction of z axis
Greek letters
attack angle of vortex generators (°)
louver angle (°)
thickness (m)
dynamic viscosity (kg m-1 s-1)
density (kg m-3)
Subscripts and superscripts
a air
co coolant region
d depth
f fin
s solid region
h height
i inner boundary of grid
in inlet
l length
L left side
o outer boundary of grid
out outlet
p pitch
R right side
ref reference data
v vortex generator
w wall
Abbreviations
CFD computational fluid dynamics
FTHE fin-tube heat exchanger
ICE internal combustion engine
L-VG louvered FTHE with vortex generators
PEMFC proton exchange membrane fuel cell
VGs vortex generators
WL wavy-louvred FTHE
WL-VGs wavy-louvred FTHE with vortex generators
1. Introduction
Owing to the advantages of high energy conversion efficiency, zero emissions, and low
operating temperature, proton exchange membrane fuel cell (PEMFC) is considered as
one of the most promising power sources for transportation, stationary, and auxiliary
applications in the future [1]. Although considerable researches have been conducted,
there still exist challenges such as durability and cost, which largely restrict the large-
scale commercialization [2]. To achieve the high energy conversion efficiency, PEMFC
needs to work within the appropriate temperature range, which is typically from 60℃
to 90℃ [3,4]. High operating temperatures (>100℃) can accelerate the degradation of
membrane electrode assembly [5] while low temperatures (<30℃) are not favorable for
electrochemical reaction kinetics [6]. Therefore, heat management is crucial to the
stable and high-efficiency operation of PEMFC system.

According to the stack power and the system complexity, cooling techniques can be
generally divided into four types [7-9], including edge cooling, air cooling, liquid
cooling, and phase change cooling. Edge cooling method depends on heat conduction
in the vertical-plane and in-plane direction to remove heat from the central region to
stack edges, which is usually appropriate for the stack less than 1 kW [10,11]. Air
cooling method makes use of separated channels in bipolar plates or separated cooling
plates for the cooling purpose, and the heat is carried away by the air. The method is
typically suitable for the stack power ranging from 100 W to 2 kW [7]. As for the stack
power over 5 kW, liquid cooling method is considered to be more beneficial than air
cooling method [8]. Owing to the high heat transfer coefficients between coolant and
stack, liquid cooling is widely adopted for PEMFC stacks over 5 kW, especially for
automotive PEMFC stacks with the output power higher than 80 kW [12]. The latent
heat of phase change material is used in the phase change cooling method, which
consists of evaporative cooling and boiling cooling. Among aforementioned cooling
techniques, liquid cooling is mostly applied due to the strong cooling capacity and
flexible control strategy. A typical liquid cooling system consists of coolant pump,
thermostat, fan, and heat exchanger. Heat exchangers can be generally divided into shell
and tube heat exchanger [13], plate and tube heat exchanger [14], and micro heat
exchanger [15]. Shell and tube heat exchangers are surface type heat exchangers
consisting of external shell and multiple tubes, which are widely used in chemical
engineering and processing industry. They are advantageous in versatility, robustness,
and reliability while disadvantageous in heat transfer efficiency, cost, and compactness
[16,17]. Plate and tube heat exchangers are composed of different plates and vertical
tubes, which have the advantages of good compactness and high efficiency. Micro heat
exchangers range from a few microns to hundreds of microns. Flexibility and high heat
transfer coefficient are the biggest advantages, making it widely applied in
semiconductor and microelectronics industries [18]. In comparison, plate and tube heat
exchangers with enhanced fins such as louvered fins and wavy fins are applied widely
in traditional automotive vehicles due to good compactness and performances [19].
However, the heat dissipation capacity of such fin-tube radiator with enhanced fins is
not sufficient for PEMFC vehicles. The engine temperature of traditional internal
combustion engine (ICE) cars usually exceeds 500℃ while PEMFC stack works at
around 80℃. Based on Newton’s law of cooling, the required surface area is given by

= ℎΔ . Q represents the total amount of heat, h represents the heat transfer

coefficient, and Δ represents the temperature difference between coolant and


working core of radiator. To remove the comparable amount of additional heat, the
radiator of fuel cell vehicles requires about 5–10 times larger heat transfer surface area
than that of ICE vehicles. Therefore, current radiators designed for traditional vehicles
require further optimizations to be applicable to PEMFC vehicles.

Many researches about fin-tube heat exchanger (FTHE) have been conducted for
performance enhancement, which can be summarized as active methods and passive
methods. Active methods require additional power consumption such as surface
vibration [20], fluid vibration [21], and fluid injection [23]. Passive methods make use
of special surface or fluid additive. Coiled tubes [24], extended surfaces [25][26], rough
surfaces [27], and swirl flow devices [28] are used as special surface, and nanofluid is
applied as enhanced thermal characteristics fluid [29]. In comparison, passive methods
are widely applied due to the advantages of simplicity, cost, and acceptable pressure
drop loss [30,31]. Based on the above researches, there are three effective passive
methods, including louvered fin type, wavy fin type, and vortex generator type.
Louvered and wavy fin-tube heat exchangers are enhanced fin type methods while
vortex generators are attached inserts. Kim at al.[32] experimentally compared different
types of fin pattern. It was found that herringbone wavy fins showed better heat transfer
characteristics than smooth wavy fins. Feibig et al. [33] found that the averaged heat
transfer performances in laminar channel-flow regime were enhanced by more than
50%, but the pressure drop was increased by 45% with delta and rectangular winglets.
Kim and Bullard [34] studied the air-side heat transfer performances for multi-louvered
fin heat exchangers. Dogan et al. [35] further compared the air-side thermal
characteristics between double-row and triple-row multi-louvered fin exchangers under
identical thermal conditions. Joardar [36] experimentally evaluated the air-side heat
transfer enhancement with winglet type vortex generator arrays. It was found that the
heat transfer coefficient was increased from 16.5% to 44% for the single-row winglet
arrangement with less than 12% pressure drop increment. In addition to experimental
researches, numerous heat exchanger models have also been developed. Perrotin and
Clodic [37] predicted the heat transfer performances based on both two-dimensional
(2D) and three-dimensional (3D) computational fluid dynamics (CFD) models. It was
found that the results of 3D model were much closer to the experimental results
compared with that of the 2D model. Carija Z et al. [38] numerically compared the heat
exchanger performances between flat fins and louvered fins with Reynolds number
ranging from 70 to 350. Based on the commercial fluid flow solver Star-CCM+, it was
found that heat transfer performances were significantly enhanced with a slightly higher
pressure drop by using louvered fins instead of flat fins. The greatest performances
improvement of 58% was obtained with Reynold number equaling 350. Li et al. [39]
and Saleem A [40] studied heat transfer and pressure drop performances of different fin
configurations. The periodic convergent divergent channels by wavy-fins showed the
lowest overall performances but the highest heat transfer rate. Dong [41] numerically
and experimentally investigated heat transfer characteristics of wavy fin heat
exchangers. Standard k-ε model (SST) was applied, and it was shown that waviness
had distinct effects on heat transfer and pressure drop characteristic while wavy fin
profile had little effects. In addition, the combination of different enhancement methods
has also been investigated. The study of Sanders et al.[42] and Ribeiro et al.[43] was
focused on improving the heat transfer along the tube wall by adding winglets on
louvers. Effects of attack angle, aspect ratio, direction, and shape were evaluated.
Huisseune [44] studied the effects of punching delta winglet vortex generators into the
louvered fin surface. The double precision segregated solver was used to solve Navier–
Stokes equations, and the semi-implicit method for pressure-linked equations (SIMPLE)
algorithm was adopted in ANYSYS Fluent. Based on the same heat duty and pumping
power, the louvered fin with delta winglets was more compact than the cases without
delta winglets. Influences of louver angle and fin pitch were further investigated by
Berrin et al. [45]. The Colburn factor, friction factor, Stanton number, and volume
goodness factor were demonstrated. Dezan et al.[46] investigated the heat transfer
enhancement and associated pressure losses with rectangular winglet vortex generators
placed on louvered fins. The screening analysis results indicated that the geometry type
and louvered-vortex-generator type drastically changed the contribution of louvered
angle and vortex generator attack angle. Effects of flow path arrangement [47] and
plate-fin type variation [48] were also studied for the purpose of increasing heat transfer
efficiency and decreasing flow resistance. In summary, the majority of aforementioned
researches are aimed at improving the louvered fin-tube heat exchanger efficiency by
optimizing geometric parameters. However, the combination of different enhanced
structural designs is not thoroughly studied while it is extremely important for heat
transfer augmentation.

In the study, a three-dimensional steady-state radiator model is developed by open


source CFD tool box OpenFOAM. Three novel structural designs of FTHE are
proposed, which combines the advantages of wavy fin and vortex generators. Several
characteristics and non-dimensional parameters are adopted or defined to
comprehensively evaluate radiator performances, including air side heat transfer
coefficient, pressure drop, outlet air temperature, heat flux, and JF factor. Temperature,
velocity, and pressure distribution are further demonstrated to reveal the performance
enhancement mechanisms. The research gives valuable suggestions for cooling
capability improvement as well as radiator volume diminution.

2. Model development
The schematic diagram of a conventional louvered fin-tube heat exchanger (FTHE) is
shown in Fig. 1(a). A single unit structure of louvered FTHE and its cross-sectional
view are shown in Fig. 1(b) and Fig. 1(c). In the channels between adjacent fins, air
flows through the louvered area. Due to periodic and symmetric characteristics of fins
and tubes, the half domain in the z coordinate and two pieces of fins in the y coordinate
is selected as the computational domain, which is shown in Fig. 1(d). Note that the
computational domain is 40 mm extended for the entrance section to ensure the inlet
uniformity, and it is 80 mm extended for the exit section to avoid the circulation effect.

(a) (b)

(c) (d)
Fig. 1. Schematic diagram of a conventional louvered fin-tube heat exchanger. (a) The
overall structure. (b) A single unit. (c) Cross-sectional view of the single unit. (d)
Computational domain.

2.1 Model assumptions


Due to the low Reynolds number (calculated to be 119.1-1270.7 in the presented study),
small differences along the fin side, and other neglectable effects, several assumptions
are made [49,50].
(1) The flow is steady-state and laminar without viscous dissipation.
(2) Air flow is incompressible.
(3) Radiation heat dissipation is neglected.
(4) Natural convection is neglected since it is marginal when compared with the forced
heat convection between coolant and radiator.

2.2 Governing equations


The governing equations of radiator model include continuity, momentum and energy
conservation equations for fluid and solid regions.
Continuity equation:

+ ∇( )=0 (1)

Momentum equation:
( )
+∇⋅( ) = −∇ + ∇ ⋅ ( (∇ + ∇ )) (2)

Energy equation:

( )+∇⋅( ( + )) = ∇ ⋅ ( ∇ ) (3)

2.3 Novel structures with wavy-louvered fin and vortex generator


Three types of novel structural designs are proposed based on the combination of wavy
louver and vortex generator. The single unit structure of wavy-louvered FTHE (WL)
and its cross-sectional view are shown in Fig. 2(a) and Fig. 2(b). The plain louvers are
replaced with folding louvers, which resemble wavy type along the flow path. The
single unit structure of louvered FTHE with vortex generators (L-VGs) and its front
view are shown in Fig. 2(c) and Fig. 2(d). The VGs are implemented in the middle part
of louver surface. The single unit structure of wavy-louvered FTHE with VGs (WL-
VGs) is shown in Fig. 2(e), which combines the wavy louver with the vortex generator.
The detailed view of a louvered fin in WL-VGs is shown in Fig. 2(f). Table 1 lists the
geometric parameters of the above four structural designs.

(a) (b)

(c) (d)

(e) (f)
Fig. 2. Schematic of three novel structural designs. (a) The wavy-louvered unit. (b) The
cross-sectional view of wavy-louvered unit. (c) The unit of FTHE with VGs. (d) The
front sectional view of FTHE unit with VGs. (e) The unit of wavy-louvered FTHE with
VGs. (f) The detailed view of a wavy-louvered fin with VGs.

Table 1. Structural parameters of the proposed heat exchangers.


Parameter Value Parameter Value

Tube length 19 mm Louver pitch 1.7 mm


Tube thickness 0.35 mm Louver height ! 6.4 mm

Tube pitch 10.45 mm Louver angle 23º

Fin height "! 8.15 mm VGs attack angle 60º


Fin thickness # 0.1 mm VGs height $% 0.2 mm

Fin pitch " 1.4 mm VGs length % 0.4 mm

Fin depth "& 20 mm VGs thickness % 0.2 mm


Wavy fin length ' 0.6 mm

2.4 Boundary and initial conditions


The computational domain of fluid and solid regions can be divided into three
subdomains: fin and tube domain, air domain and coolant domain. For the interfaces
between solid domain and air domain as well as the interfaces between solid domain
and coolant domain, the temperature and the heat flux are identical. The top section is
defined as the symmetry boundary condition, and the left and right sides are defined as
the periodic boundary condition due to the continuity of radiator structure. For the air
domain, the inlet temperature is defined as 298 K, and the inlet velocity is defined as
constant. For the fin and tube domain, the initial temperature is defined as 298 K, and
the thermal conductivity is defined as 200 W m-1 K-1 while other surfaces are defined
as the no-slip condition. As for the coolant domain, the inlet temperature is defined as
343 K, and the inlet velocity is defined as constant. The bottom section is defined as
the symmetry boundary condition. Heat transfer surfaces for above three domains are
defined as the temperature coupled boundary condition. A summary of boundary
conditions and constant properties are listed in Table 2 and Table 3.

Table 2. Constant properties for three computational subdomains.


Domain Parameter Value

Thermal conductivity 200 W m-1 K-1


solid
Density 2700 kg m-3
Air Dynamic viscosity 1.91×10-5 kg m-1 s-1
Specific heat 0.24 J kg-1 K-1
Density 1.293 kg m-3
Dynamic viscosity 1×10-3 kg m-1 s-1
Coolant Specific heat 4.2×103 J kg-1 K-1
Density 1000 kg m-3

Table 3. Boundary conditions for three computational subdomains.


Domain Boundary name Boundary conditions

Solid_To_Air ( = ) , +( = +)

Solid Solid_To_Coolant ( = co , +( = +co


. (/ . (1
Top =−
.0 .0
Inlet =constant, 3 = 4 = 0, = in =298 K
7 9 :
Outlet
8
= 8
= 8
= 0, 8
=0

. )/ . )1 .4)/ .4)1
Top =− , =−
.0 .0 .0 .0
Air
) = ) , ) = )
; < ; <

Left and right sides . ); . )< . ); . )<


= , =
.= .= .= .=

Air_To_Solid ) = ( , +) = +(
Inlet 3=constant, = 4 = 0, = in =343 K
. .3 .4 .
Outlet = = = 0, =0
.= .= .= .=
Coolant
. co/ . co1 .4co
/ 1
.4co
Bottom =− , =−
.0 .0 .0 .0
Coolant_To_Solid co = ( , +co = +(

2.5 Numerical procedures


To solve the conjugate heat transfer problem between fluid and solid regions, the
OpenFOAM is adopted in the present study. Preprocessor CATIA V5-6 and ICEM CFD
15.0 are used for establishing the physical model and corresponding meshing.
Considering the mesh equality as well as the calculation efficiency, the hexahedral
structured mesh is created. A multi-phase coupled heat transfer solver called
chtMultiRegionFoam is applied, which simultaneously solves the temperature equation
in the solid domain and the fluid domain. The PIMPLE algorithm is utilized for
coupling velocity and pressure in the solver. Meanwhile, the second-order upwind
scheme is employed for discretizing convection terms while the diffusion term is
discretized with the second-order central difference scheme. The convergence residuals
of continuity and momentum equations are set as 10-6 , and the residual of energy
equation is set as 10-7 .

2.6 Performance evaluation indexes


To comprehensively evaluate radiator performances, several characteristics and non-
dimensional parameters are adopted or defined. The Colbum factor B is calculated as
follows:
H
CD
B= Pr I (4)
7ED

The pressure drop Δ and friction factor J are defined as:


Δ = in − out (5)
MNO ΔP
J= (6)
NQ 7H

The comprehensive performance evaluation factor is defined based on the volume


goodness factor [51].
S
Sref
R" = Y (7)
W I
V X
Wref

3. Results and discussion


3.1 Model validation
The established three-dimensional radiator model in OpenFOAM is firstly validated,
including grid independence study and comparison among experimental data [34,52],
correlation [53], and simulation results. For the experimental data [34], the physical
geometry (fin pitch 1.4 mm, louver angle 23°, fin depth 20 mm, tube pitch 10.15 mm)
is developed. Seven different grid numbers including 2380721, 3823788, 4978875,
6249471, 7058032, 8314923, 9295768 elements are evaluated at the inlet air velocity
being 2 m s-1. Fig. 3(a) shows that the relative differences of air-side heat transfer
coefficient and pressure drop under different mesh numbers. The results show that
relative errors for heat transfer coefficient and pressure drop are decreased to be less
than 0.6% and 0.17% when the grid number reaches 7058032. In addition, Fig. 3(b)
shows the relative differences of heat transfer coefficient while Fig. 3(c) shows the
relative differences of pressure drop between simulation results and experimental data
[34] under different inlet air velocity. Since the experimental data only ranges from 1
m s-1 to 3 m s-1, the comparison between simulation results and correlations [53] is also
shown. For experimental data from Dong J et al. [52], the physical geometry (fin pitch
2 mm, louver angle 28o, fin depth 36.6 mm, fin height 7 mm) is developed. The grid
number from about 2.2 million to about 14.5 million are compared at the inlet air
velocity being 5 m s-1. Fig. 3(d) shows that the relative differences of air-side heat
transfer coefficient and pressure drop become less than 0.15% and 0.36% when the grid
number reaches 10506124. Fig. 3(e) and Fig. 3(f) show that the averaged relative
differences between experimental data [52] and simulation results are 7.89% for heat
transfer coefficient and 5.56% for pressure drop. Based on the above two groups of
comparison, the reliability of established 3D model is well validated.

16
Heat transfer coefficient Heat transfer coefficient
1.0
Pressure drop 14 Pressure drop
Relative difference (%)

Relative difference (%)

0.8 12

10
0.6
8

0.4 6

4
0.2
2

0.0 0
2 4 6 8 10 2 4 6 8 10 12 14 16
Grid number (Million) Grid number (Million)

(a) (d)
220 280

Heat transfer coefficient (W m-2 K-1)


Heat transfer coefficient (W m-2 K-1) Simulation results
Simulation results
Experimental results[48] 260
200 Experimental results[49]
Correlation Ref[50]

180 240

160 220

140 200

120 180

100 160

140
80
3 4 5 6 7 8 9
0 2 4 6 8
-1
Inlet air velocity (m s ) Inlet air velocity (m s-1)

(b) (e)

160 600
Simulation results Simulation results
Experimental results[48]
Experimental results[49]
Correlation Ref[50] 500
120
Pressure drop (Pa)
Pressure drop (Pa)

400
80
300

40
200

0 100
0 2 4 6 8 3 4 5 6 7 8 9
Inlet air velocity (m s-1) Inlet air velocity (m s-1)

(c) (f)
Fig. 3. Model validation. (a) Grid independence study for experimental data [34]. (b)
Comparison of heat transfer coefficient among experimental data [34], correlation [53],
and simulation results under different inlet air velocity. (c) Comparison of pressure drop
under different inlet air velocity. (d) Grid independence study for experimental data
[52]. (e) Comparison of heat transfer coefficient between experimental data [52] and
simulation results under different inlet air velocity. (f) Comparison of pressure drop
under different inlet air velocity.

3.2 Novel structural designs


Based on the specific physical geometry from experimental data [34], three novel
structural designs of fin-tube heat exchanger (FTHE) are proposed based on the
application of wavy-louvered (WL) fin and vortex generator (VG). The proposed three
structural designs are comprehensively compared against the conventional structure
under the inlet air velocity ranging from 0.75 m s-1 to 8 m s-1 (Reynold number 119.1-
1270.7). The air side heat transfer coefficient, pressure drop, outlet air temperature, heat
flux, and JF factor are adopted to evaluate heat transfer performances.

Fig. 4 shows the comparison of FTHE performances among different structural designs.
In Fig. 4(a)(b), it is seen that the FTHE with L-VGs has the highest air side heat transfer
coefficient while the FTHE with WL-VGs has the highest pressure drop. Fig. 4(c)(d)
show that the FTHE with WL-VGs has the highest outlet air temperature and heat flux.
The wavy louver design significantly increases the heat flux, but the heat flux grows
little with the further addition of VGs. For JF factor in Fig. 4(e), the FTHE with L-VGs
shows better performances in the region of low Reynold number, but it decreases
significantly with Re increasing. Meanwhile, the JF factors of FTHE with WL and WL-
VGs rise slightly with higher Re. The reason is that the growing rate of heat transfer
coefficient and the pressure drop varies with Re increasing.

(a) (b)
(c) (d)

(e)
Fig. 4. Comparison of FTHE performances among different structural designs. (a) Air
side heat transfer coefficient. (b) Pressure drop. (c) Outlet air temperature. (d) Heat flux.
(e) JF factor.

Since the heat transfer performances of four structural designs have specific
characteristics, the temperature, velocity, and pressure distribution with inlet air
velocity being 8 m s-1 are further demonstrated to deeply expound the characteristics.
Fig. 5 shows the temperature distribution of fin and tube region. It can be seen from
Fig. 5(a)(b) that the temperature of adjacent louver after the inlet louver decreases
significantly, which is not observed in Fig. 5 (c)(d). The reason is that the air flow in
the near-wall region is characterized by large separation zones, especially in the
adjacent louver after the inlet louver. However, the air flow in FTHE with WL is not as
much separated as the conventional FTHE since the zigzag wavy louver restricts flow
separation. Besides, the low temperature area close to the bottom region increases with
the addition of VGs on both conventional and wavy louver. Since Fig. 4(a) shows that
the L-VGs design has the highest heat transfer coefficient, the reason is that the L-VGs
strongly intensifies the heat transfer between louvers but inapparently affects the flow
conditions on un-louvered fins.

(a) (b)

(c) (d)
Fig. 5. Temperature distribution of fin and tube region (u=8 m s-1). (a) Conventional
FTHE. (b) FTHE with L-VGs. (c) FTHE with WL. (d) FTHE with WL-VGs.

Fig. 6 shows the temperature distribution at the cross-section of air region, which
indicates the heat transfer intensity between air and fin. Compared with the
conventional FTHE, the region behind vortex generator in the air flow direction shows
higher temperature in Fig. 6(b)(d). It is seen in Fig. 6 (a)(c) that zigzag wavy fin
increases the effective heat transfer area in the direction of fins. Fig. 7 further shows
the temperature distribution of near-wall region, and Fig. 7(a) shows the position of
cutting slice. As shown in Fig. 7(c)(e), the heat exchangers with additional vortex
generators produce two sections of high-temperature wakes near the wall. The two
wakes develop along the tail of the vortex generators, which not only promotes the
convection heat exchange, but also contributes to the heat exchange in the nearby area.
Therefore, the air temperature near the entire louver is increased, and the heat exchange
efficiency of heat exchanger is improved.
(a)

(b)

(c)

(d)
Fig. 6. Temperature distribution at the cross-section of air region (u=8 m s-1). (a)
Conventional FTHE. (b) FTHE with L-VGs. (c) FTHE with WL. (d) FTHE with WL-
VGs.

(a)

(b) (c) (d) (e)


Fig. 7. Temperature distribution of the near-wall region (u=8 m s-1). (a) The cutting slice
position. (b) Conventional FTHE. (c) FTHE with L-VGs. (d) FTHE with WL. (e) FTHE
with WL-VGs.

Fig. 8 shows the velocity distribution at the cross-section of air region. It is observed
that L-FTHE and WL-FTHE both separate air flow and induce secondary flow by flat
louver and wavy louver. Although WL-FTHE disturbs flow and generates a specific
low-speed zone, it splits the velocity field and makes it more evenly distributed at
relatively higher Re. Therefore, the heat transfer coefficient of WL-FTHE increases
slightly compared with that of L-FTHE as shown in Fig. 4 (a). Note that the heat transfer
coefficient of WL-FTHE is nearly equal to that of L-FTHE at low Re. The reason is
that the narrow channel between louvers in WL-FTHE makes it hard for air to flow
through, which decreases the influences of longer wavy channel. Owing to the blocking
effect of VGs, the velocity field between louvers is reduced as shown in Fig. 8(b)(d).
Meanwhile, a low-speed vortex zone behind VGs appears and generates longitude
vortex, which makes the air stream stay longer at fin surfaces and strongly promotes
the local heat transfer efficiency. Fig. 9 shows the generated vortex at the tail of vortex
generator for FTHE with L-VGs, revealing the reason for enhanced convective heat
transfer in the part of specific region.

(a)

(b)

(c)

(d)
Fig. 8. Velocity distribution at the cross-section of air region (u=8 m s-1). (a)
Conventional FTHE. (b) FTHE with L-VGs. (c) FTHE with WL. (d) FTHE with WL-
VGs.

Fig. 9. Velocity distribution at the tail of vortex generator for FTHE with L-VGs.

The pressure distribution at the cross-section of air region is shown in Fig. 10. It can be
seen that significant pressure drop losses occur around VGs while wavy louver only
affects the former section of louver. FTHE with WL-VGs has the highest pressure drop
penalty with combination of VGs and wavy louver, which corresponds to Fig. 4(b).
Such significant increasement of pressure drop losses restricts the augment of
comprehensive heat transfer performances as verified in Fig. 4(e). Since the wave
shutter fins reduce the area of air flow path between adjacent fins to a certain extent,
the blocking effect of vortex generating element is enlarged, which limits the
improvement of comprehensive heat transfer performance for FTHE with WL-VGs.

(a)

(b)

(c)

(d)
Fig. 10. Pressure distribution at the cross-section of air region (u=8 m s-1). (a)
Conventional FTHE. (b) FTHE with L-VGs. (c) FTHE with WL. (d) FTHE with WL-
VGs.

4. Conclusions
To investigate three novel structural designs of fin-tube heat exchanger (FTHE) based
on the conventional FTHE, a three-dimensional steady-state radiator model is
developed by open source CFD tool box OpenFOAM. Vortex generators (VGs) and
wavy louvers are utilized to augment heat transfer coefficient and increase effective
heat transfer area. Heat transfer coefficient, pressure drop, outlet air temperature, heat
flux per unit and JF factor are adopted to evaluate overall heat transfer performances.
Numerical simulations are performed based on the solver chtMultiRegionFoam with
PIMPLE algorithm. Grid independence study and comparison among experimental data,
correlation, and simulation results have been rigorously conducted. It is found that the
wavy-louvered FTHE augment heat transfer by generating flow disturbance, making it
advantageous at relatively higher Re. Meanwhile, the wavy-louvered FTHE can
significantly enlarge the effective heat transfer area density, making it possible for
further volume reduction under the same heat dissipation capacity. For FTHE with L-
VGs, near-wall vortexes generated by louvers with vortex generator can intensify the
convective heat transfer because a low-speed vortex zone behind VGs appears, which
generates longitude vortex and makes the air stream stay longer at fin surfaces. At
relatively lower Re, FTHE with VGs show better performances. However, heat transfer
performances of FTHE with VGs decreases rapidly with Re rising, which is caused by
the significant pressure drop penalty. For FTHE with WL-VGs, the wave shutter fins
reduce the area of air flow path between adjacent fins to a certain extent, but the
blocking effect of vortex generating element is enlarged, which limits the further
improvement of comprehensive heat transfer performance.

Acknowledgement
The research is supported by National Engineering Laboratory for Mobile Source
Emission Control Technology (No. NELMS2019A10).

References
[1] O’Hayre R, Cha S-W, Colella W, Prinz FB. Fuel Cell Fundamentals. 2016.
https://doi.org/10.1002/9781119191766.
[2] Wang Y, Chen KS, Mishler J, Cho SC, Adroher XC. A review of polymer
electrolyte membrane fuel cells: Technology, applications, and needs on
fundamental research. Appl Energy 2011.
https://doi.org/10.1016/j.apenergy.2010.09.030.
[3] Wang Y, Ruiz Diaz DF, Chen KS, Wang Z, Adroher XC. Materials, technological
status, and fundamentals of PEM fuel cells – A review. Mater Today 2020.
https://doi.org/10.1016/j.mattod.2019.06.005.
[4] Lochner T, Kluge RM, Fichtner J, El-Sayed HA, Garlyyev B, Bandarenka AS.
Temperature Effects in Polymer Electrolyte Membrane Fuel Cells.
ChemElectroChem 2020. https://doi.org/10.1002/celc.202000588.
[5] Scofield ME, Liu H, Wong SS. A concise guide to sustainable PEMFCs: recent
advances in improving both oxygen reduction catalysts and proton exchange
membranes. Chem Soc Rev 2015. https://doi.org/10.1039/c5cs00302d.
[6] Jiao K, Li X. Water transport in polymer electrolyte membrane fuel cells. Prog
Energy Combust Sci 2011. https://doi.org/10.1016/j.pecs.2010.06.002.
[7] Zhang G, Kandlikar SG. A critical review of cooling techniques in proton exchange
membrane fuel cell stacks. Int J Hydrogen Energy 2012.
https://doi.org/10.1016/j.ijhydene.2011.11.010.
[8] Dicks A, Rand DAJ. Fuel Cell Systems Explained. 2018.
https://doi.org/10.1002/9781118706992.
[9] Song TW, Song TW, Choi KH, Kim JR, Yi JS. Pumpless thermal management of
water-cooled high-temperature proton exchange membrane fuel cells. J Power
Sources 2011. https://doi.org/10.1016/j.jpowsour.2010.12.108.
[10] Vasiliev L. Heat Pipes in Fuel Cell Technology, 2008. https://doi.org/10.1007/978-
1-4020-8295-5_8.
[11] Vasiliev LL. Heat pipes− good tool for fuel cells thermal management, 2011.
https://doi.org/10.1615/ichmt.2009.conv.50.
[12] Fuel Cells 2000. Fuel cell vehicles (from auto manufacturers)
http://www.fuelcells.org/info/charts/carchart.pdf; 2011
[13] Kuchi G, Ponyavin V, Chen Y, Sherman S, Hechanova A. Numerical modeling of
high-temperature shell-and-tube heat exchanger and chemical decomposer for
hydrogen production. Int J Hydrogen Energy 2008.
https://doi.org/10.1016/j.ijhydene.2008.06.072.
[14] Shafiee S, McCay MH. Different reactor and heat exchanger configurations for
metal hydride hydrogen storage systems - A review. Int J Hydrogen Energy 2016.
https://doi.org/10.1016/j.ijhydene.2016.03.133.
[15] Abou Elmaaty TM, Kabeel AE, Mahgoub M. Corrugated plate heat exchanger
review. Renew Sustain Energy Rev 2017.
https://doi.org/10.1016/j.rser.2016.11.266.
[16] Erbay LB, Doğan B, Öztürk MM. Comprehensive Study of Heat Exchangers with
Louvered Fins. Heat Exch. - Adv. Featur. Appl., 2017.
https://doi.org/10.5772/66472.
[17] Lei Y, Li Y, Jing S, Song C, Lyu Y, Wang F. Design and performance analysis of
the novel shell-and-tube heat exchangers with louver baffles. Appl Therm Eng
2017. https://doi.org/10.1016/j.applthermaleng.2017.07.081.
[18] Dixit T, Ghosh I. Review of micro- and mini-channel heat sinks and heat
exchangers for single phase fluids. Renew Sustain Energy Rev 2015.
https://doi.org/10.1016/j.rser.2014.09.024.
[19] Liang YY, Liu CC, Li CZ, Chen JP. Experimental and simulation study on the air
side thermal hydraulic performance of automotive heat exchangers. Appl Therm
Eng 2015. https://doi.org/10.1016/j.applthermaleng.2015.05.018.
[20] Sathyabhama A, Prashanth SP. Enhancement of Boiling Heat Transfer Using
Surface Vibration. Heat Transf - Asian Res 2017. https://doi.org/10.1002/htj.21197.
[21] Cheng L, Luan T, Du W, Xu M. Heat transfer enhancement by flow-induced
vibration in heat exchangers. Int J Heat Mass Transf 2009.
https://doi.org/10.1016/j.ijheatmasstransfer.2008.05.037.
[22] Duan D, Ge P, Bi W, Ji J. Numerical investigation on synthetical performance of
heat transfer of planar elastic tube bundle heat exchanger. Appl Therm Eng 2016.
https://doi.org/10.1016/j.applthermaleng.2016.08.070.
[23] Hussain S, Liu J, Sundén B. Study of effects of axisymmetric endwall contouring
on film cooling/heat transfer and secondary losses in a cascade of first stage nozzle
guide vane. Appl Therm Eng 2020.
https://doi.org/10.1016/j.applthermaleng.2019.114844.
[24] Aly WIA. Numerical study on turbulent heat transfer and pressure drop of
nanofluid in coiled tube-in-tube heat exchangers. Energy Convers Manag 2014.
https://doi.org/10.1016/j.enconman.2013.12.031.
[25] Mohammed HA, Hasan HA, Wahid MA. Heat transfer enhancement of nanofluids
in a double pipe heat exchanger with louvered strip inserts. Int Commun Heat Mass
Transf 2013. https://doi.org/10.1016/j.icheatmasstransfer.2012.10.023.
[26] Zhou G, Ye Q. Experimental investigations of thermal and flow characteristics of
curved trapezoidal winglet type vortex generators. Appl Therm Eng 2012.
https://doi.org/10.1016/j.applthermaleng.2011.11.024.
[27] Kathait PS, Patil AK. Thermo-hydraulic performance of a heat exchanger tube with
discrete corrugations. Appl Therm Eng 2014.
https://doi.org/10.1016/j.applthermaleng.2014.01.069.
[28] Esmaeilzadeh E, Almohammadi H, Nokhosteen A, Motezaker A, Omrani AN.
Study on heat transfer and friction factor characteristics of γ-Al2O3/water through
circular tube with twisted tape inserts with different thicknesses. Int J Therm Sci
2014. https://doi.org/10.1016/j.ijthermalsci.2014.03.005.
[29] Sidik NAC, Adamu IM, Jamil MM, Kefayati GHR, Mamat R, Najafi G. Recent
progress on hybrid nanofluids in heat transfer applications: A comprehensive
review. Int Commun Heat Mass Transf 2016.
https://doi.org/10.1016/j.icheatmasstransfer.2016.08.019.
[30] Kareem ZS, Mohd Jaafar MN, Lazim TM, Abdullah S, Abdulwahid AF. Passive
heat transfer enhancement review in corrugation. Exp Therm Fluid Sci 2015.
https://doi.org/10.1016/j.expthermflusci.2015.04.012.
[31] Starace G, Fiorentino M, Longo MP, Carluccio E. A hybrid method for the cross
flow compact heat exchangers design. Appl Therm Eng 2017.
https://doi.org/10.1016/j.applthermaleng.2016.10.018.
[32] Kim NH, Yun JH, Webb RL. Heat transfer and friction correlations for wavy plate
fin-and-tube heat exchangers. J Heat Transfer 1997.
https://doi.org/10.1115/1.2824141.
[33] Fiebig M, Kallweit P, Mitra N, Tiggelbeck S. Heat transfer enhancement and drag
by longitudinal vortex generators in channel flow. Exp Therm Fluid Sci 1991.
https://doi.org/10.1016/0894-1777(91)90024-L.
[34] Kim MH, Bullard CW. Air-side thermal hydraulic performance of multi-louvered
fin aluminum heat exchangers. Int J Refrig 2002. https://doi.org/10.1016/S0140-
7007(01)00025-1.
[35] Dogan BI, Altun Ö, Ugurlubilek N, Tosun M, Sariçay T, Erbay LB. An
experimental comparison of two multi-louvered fin heat exchangers with different
numbers of fin rows. Appl Therm Eng 2015.
https://doi.org/10.1016/j.applthermaleng.2015.07.059.
[36] Joardar A, Jacobi AM. Heat transfer enhancement by winglet-type vortex generator
arrays in compact plain-fin-and-tube heat exchangers. Int J Refrig 2008.
https://doi.org/10.1016/j.ijrefrig.2007.04.011.
[37] Perrotin T, Clodic D. Thermal-hydraulic CFD study in louvered fin-and-flat-tube
heat exchangers. Int J Refrig 2004. https://doi.org/10.1016/j.ijrefrig.2003.11.005.
[38] Carija Z, Franković B, Perčić M, Čavrak M. Heat transfer analysis of fin-and-tube
heat exchangers with flat and louvered fin geometries. Energy Econ 2014.
https://doi.org/10.1016/j.ijrefrig.2014.05.026.
[39] Li L, Du X, Yang L, Xu Y, Yang Y. Numerical simulation on flow and heat transfer
of fin structure in air-cooled heat exchanger. Appl Therm Eng 2013.
https://doi.org/10.1016/j.applthermaleng.2013.05.012.
[40] Saleem A, Kim MH. Air-side thermal hydraulic performance of microchannel heat
exchangers with different fin configurations. Appl Therm Eng 2017.
https://doi.org/10.1016/j.applthermaleng.2017.07.082.
[41] Dong J, Chen J, Zhang W, Hu J. Experimental and numerical investigation of
thermal -hydraulic performance in wavy fin-and-flat tube heat exchangers. Appl
Therm Eng 2010. https://doi.org/10.1016/j.applthermaleng.2010.02.027.
[42] Sanders PA, Thole KA. Effects of winglets to augment tube wall heat transfer in
louvered fin heat exchangers. Int J Heat Mass Transf 2006.
https://doi.org/10.1016/j.ijheatmasstransfer.2006.03.036.
[43] Ribeiro F, de Conde KE, Garcia EC, Nascimento IP. Heat transfer performance
enhancement in compact heat exchangers by the use of turbulators in the inner side.
Appl Therm Eng 2020. https://doi.org/10.1016/j.applthermaleng.2020.115188.
[44] Huisseune H, T’Joen C, Jaeger P De, Ameel B, Schampheleire S De, Paepe M De.
Performance enhancement of a louvered fin heat exchanger by using delta winglet
vortex generators. Int J Heat Mass Transf 2013.
https://doi.org/10.1016/j.ijheatmasstransfer.2012.09.004.
[45] Erbay LB, Uğurlubilek N, Altun Ö, Doğan B. Numerical Investigation of the Air-
Side Thermal Hydraulic Performance of a Louvered-Fin and Flat-Tube Heat
Exchanger at Low Reynolds Numbers. Heat Transf Eng 2017.
https://doi.org/10.1080/01457632.2016.1200382.
[46] Dezan DJ, Yanagihara JI, Jenovencio G, Salviano LO. Parametric investigation of
heat transfer enhancement and pressure loss in louvered fins with longitudinal
vortex generators. Int J Therm Sci 2019.
https://doi.org/10.1016/j.ijthermalsci.2018.09.039.
[47] Wang Y, Lei RW, Wang H, Wang SF, Bai JQ. Thermal performance analysis of fin-
and-tube heat exchanger circuit in supercritical hydrogen refrigeration cycle
system. Int J Hydrogen Energy 2019.
https://doi.org/10.1016/j.ijhydene.2019.04.163.
[48] Ozturk MM. Effect of Fin Type on Performance of Compact Heat Exchangers with
Mini-Channel. Heat Transf Eng 2020.
https://doi.org/10.1080/01457632.2020.1860522.
[49] Park JS, Byun S, Kim DR, Lee KS. Frost behavior of a louvered fin heat exchanger
with vortex-generating fins. Int J Heat Mass Transf 2017.
https://doi.org/10.1016/j.ijheatmasstransfer.2017.06.090.
[50] Välikangas T, Singh S, Sørensen K, Condra T. Fin-and-tube heat exchanger
enhancement with a combined herringbone and vortex generator design. Int J Heat
Mass Transf 2018. https://doi.org/10.1016/j.ijheatmasstransfer.2017.11.006.
[51] Shah R K, London A L. Laminar flow forced convection heat transfer and flow
friction in straight and curved ducts. A summary of analytical solutions. Stanford
Univ 1971.
[52] Dong J, Chen J, Chen Z, Zhang W, Zhou Y. Heat transfer and pressure drop
correlations for the multi-louvered fin compact heat exchangers. Energy Convers
Manag 2007. https://doi.org/10.1016/j.enconman.2006.11.023.
[53] Wang CC, Chang CT. Heat and mass transfer for plate fin-and-tube heat exchangers,
with and without hydrophilic coating. Int J Heat Mass Transf 1998.
https://doi.org/10.1016/S0017-9310(98)00060-X.

You might also like