You are on page 1of 32

Journal Pre-proofs

A new concept of thermal management system in Li-ion battery using air


cooling and heat pipe for electric vehicles

Hamidreza Behi, Danial Karimi, Mohammadreza Behi, Morteza


Ghanbarpour, Joris Jaguemont, Mohsen Akbarzadeh Sokkeh, Foad Heidari
Gandoman, Maitane Berecibar, Joeri Van Mierlo

PII: S1359-4311(19)37109-1
DOI: https://doi.org/10.1016/j.applthermaleng.2020.115280
Reference: ATE 115280

To appear in: Applied Thermal Engineering

Received Date: 15 November 2019


Revised Date: 22 March 2020
Accepted Date: 2 April 2020

Please cite this article as: H. Behi, D. Karimi, M. Behi, M. Ghanbarpour, J. Jaguemont, M. Akbarzadeh Sokkeh,
F. Heidari Gandoman, M. Berecibar, J. Van Mierlo, A new concept of thermal management system in Li-ion
battery using air cooling and heat pipe for electric vehicles, Applied Thermal Engineering (2020), doi: https://
doi.org/10.1016/j.applthermaleng.2020.115280

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. Please note that, during the production process, errors
may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier Ltd.


A new concept of thermal management system in Li-ion battery using air cooling and heat
pipe for electric vehicles
Hamidreza Behiab*, Danial Karimiab, Mohammadreza Behicd, Morteza Ghanbarpourd, Joris Jaguemontab, Mohsen Akbarzadeh Sokkehab, Foad
Heidari Gandomanab, Maitane Berecibarab, Joeri Van Mierloab

a Research group MOBI – Mobility, Logistics, and Automotive Technology Research Centre, Vrije Universiteit Brussel, Pleinlaan 2, Brussels 1050, Belgium

b Flanders Make, Heverlee 3001, Belgium

c The University of Sydney, School of Chemical and Biomolecular Engineering, NSW 2006, Australia

d Department of Energy Technology, KTH Royal Institute of Technology, SE-10044 Stockholm, Sweden

Abstract
This paper presents the concept of a hybrid thermal management system (TMS), including air
cooling and heat pipe for electric vehicles (EVs). Mathematical and thermal models are described
to predict the thermal behavior of a battery module consisting of 24 cylindrical cells. Details of
various thermal management techniques, especially natural air cooling and forced-air cooling TMS
are discussed and compared. Moreover, several optimizations comprising the effect of cell spacing,
air velocity, different ambient temperatures, and adding a heat pipe with copper sheets (HPCS) are
proposed. The mathematical models are solved by COMSOL Multiphysics®, the commercial
computational fluid dynamics (CFD) software. The simulation results are validated against
experimental data indicating that the proposed cooling method is robust to optimize the TMS with
HPCS, which provides guidelines for further design optimization for similar systems. Results
indicate that the maximum module temperature for the cooling strategy using forced-air cooling,
heat pipe, and HPCS reaches 42.4 °C, 37.5 °C, and 37.1 °C which can reduce the module
temperature compared with natural air cooling by up to 34.5%, 42.1%, and 42.7% respectively.
Furthermore, there is 39.2%, 66.5%, and 73.4% improvement in the temperature uniformity of the
battery module for forced-air cooling, heat pipe, and HPCS respectively.
Keywords
Lithium-ion battery; Battery thermal management; Air cooling system; Heat pipe; Computational
Fluid Dynamics (CFD)

Nomenclature
t Time Interval (t)

T Battery Temperature (K)

I Discharge Current (Ah)

OCV Open-circuit Voltage (V)

V Operating Voltage (V)

v Velocity (m.s-1)

S Spacing (m)

cp Specific Heat capacity (J/kg.K)

Tamb Temperature of ambient (K)

Rt Total Resistance of Heat pipe (K/W)

k Thermal Conductivity (W/m.K)

kr Radial Thermal Conductivity (W/m.K)

kz Axial Thermal Conductivity (W/m.K)

p Pressure (Pa)

R Resistance (K/W)

A Surface Area (m2)

h Heat Transfer Coefficient (W/m2.K)

𝑞 Volumetric Heat Generation (W/m3)

Q loss Power Loss of Battery (W)

gi Gravity (m.s-2)

L Thickness of Wall or Wick (m)

La Adiabatic Length of Heat Pipe (m)

Lc Condenser length of Heat Pipe (m)

Le Evaporator Length of Heat Pipe (m)


Partial Error
𝑈𝑉𝐼
𝑈𝑅 Total Error
Greek

∀ Volume of Battery Cell (m2 /s)

𝜇 Dynamic Viscosity (Pa.s)

𝜌 Density (kg/m3)

𝜆 Heat Generation (W)

Acronyms

CFD
Computational Fluid Dynamics
TMS
Thermal Management System
HPCS
Heat Pipe Copper Sheet
NC
Natural Convection
HP
Heat pipe
FC
Forced Convection
EV
Electric Vehicle
OEM
Original Equipment Manufacturer
PCM Phase Change Material

CO2 Carbon Dioxide


1. Introduction

Nowadays, climate change is one of the most critical issues worldwide [1]. The transportation
industry is responsible for 24% of the direct carbon dioxide (CO2) emission from fuel combustion
in 2018 [2]. A possible solution is the usage of electric vehicles (EVs) as an alternative to reduce
some parts of the air pollution problem through original equipment manufacturers (OEMs) [3].
Lithium-ion (Li-ion) batteries are introduced to feed electric motors in the EVs as a source. Li-ion
batteries usage has been developed rapidly because of the advantages of acceptable recyclability,
high energy, and power density. They have been widely applied in different fields comprising of
electronic products and EVs [4]. Despite positive features, using Li-ion batteries in high-current
applications and different ambient temperatures [5,6] lead to overheating because of excessive
heat generation [7]. Li-ion batteries should work within a specific temperature and thermal
specifications. The most effective temperature in which Li-ion batteries should operate is in the
range of working temperatures between 15 ℃ to 40 ℃ [8]. An inappropriate thermal management
system (TMS) for battery systems leads to heat accumulation, which could overheat the battery
module. If this excessive heat generation cannot be controlled, the lifetime and efficiency reduction
will be inevitable, and in the worst case, more severe safety issues may lead to the explosion of
the battery module [9]. Additionally, the temperature uniformity among all the cells in the module
is essential. Non-uniformity of temperature distribution within battery modules may cause
electrical imbalance over time, which causes the mismatch between the state of charge of cells
[10] and the reduction of battery module performance. For example, a 10 ℃ to 15 ℃ temperature
variation in the battery module causes a 30% to 50% degradation [11]. Generally, the TMS of
battery systems is divided into active and passive cooling systems. On the one hand, active cooling
comprising of the air and liquid cooling systems that need an external source of energy [12]. On
the other hand, passive cooling, such as phase change materials (PCM) does not consume any
external energy [13]. Air cooling is one of the most suitable TMS because of the low
manufacturing cost, simple layout requirements, and high reliability. Generally, it is divided into
two classes, natural air cooling and forced air cooling. A number of researchers experimentally
and numerically discussed the effect of natural and forced air cooling on the cylindrical cells
[14,15]. Moreover, several studies [14,16–19] have been considered the influence of airflow
pattern, airflow temperature, type of cell arrangement, air velocity, inlet, and outlet position.
Nevertheless, the air-cooling system is not practical in some cases comprising stressful conditions,
high ambient temperatures, and high charge/discharge rates due to the low heat transfer coefficient
of the air [20]. Therefore, using a high heat transfer device to increase the cooling efficiency of the
system in the design of many air cooling applications is necessary. Giuliano et al. [21] built a
metal-foam heat exchanger for lithium titan battery to increase the heat transfer coefficient for air
cooling applications. Mohammadian et al. [22] designed a kind of aluminum pin fin heat sink for
prismatic Li-ion batteries. Li et al. [23] designed experimentally and numerically a battery air
cooling thermal management system using a double silica cooling plate with copper mesh. The
mentioned methods are generally used in many air cooling applications to increase the heat
transfer, however, they increase the volume and weight of the system. Therefore, finding less
volume, weight and more compact ways of cooling for Li-ion batteries is necessary. It is found
that heat pipe can transfer the heat very efficiently [24–27]. In recent years the heat pipe cooling
technology has been used vastly in battery modules [4,20,28–33]. Burban et al. [34] studied
experimentally an unlooped pulsating heat pipe for the electronic thermal management field with
hybrid vehicle applications. Rao et al. [35] used the plate heat pipe that could control the
temperature of LiFePO4 battery within 50 ℃ in the heat generation rate of less than 30 W. Besides,
it maintained the temperature difference of module in 5 ℃. Tran et al. [36] considered and analyzed
the effect and cooling performance of heat pipes and fins in different inclinations on the
evaporation of a Li-ion battery. Wu et al. [37] studied the effect of the natural convection, forced
convection, and tubular heat pipe cooling method. They found that the heat pipe cooling method
performed the best in controlling temperature rise. However, the temperature on the surface of the
cells was undesirable due to the low effective heat transfer area between the heat pipes and the
battery surface. Heat pipe TMS is often used for prismatic and pouch cells as it requires a large
contact area and volume. In fact, for cylindrical shape cells, a new contact design is needed. Feng
et al. [38] experimentally monitor the thermal and strain of cylindrical Li-ion battery pack, using
a new design combined heat pipe and fins to manage it. They found that heat pipes and fins with
the fan system can control the temperature of the whole pack to reach the operation temperature
requirement with low energy consumption. Gan et al. [39] experimentally designed a TMS
embedded with the heat pipe for a battery module with cylindrical cells. They used a conduction
element to increase the contact area between heat pipe and cells. Wang et al. [40] presented a study
using the heat pipe and the conduction element for thermal management of the cylindrical module.
They found the best optimization is achieved by 19 mm battery spacing, 4 mm conduction element
thickness and 120° circumference angle and 60 mm conduction element heigh. Nonetheless, due
to difficulty in the design of heat pipe TMS for cylindrical cells in all the mentioned cooling
systems the condenser needs a separate unit that increases the volume and weight of the system.
In present work, in order to decrease the volume and more compact cooling system, several optimal
designs have been done experimentally and numerically with air cooling, L shape heat pipe, and
heat pipe copper sheet (HPCS). The new design of HPCS is proposed for TMS of a battery module
with cylindrical cells in which the condenser located inside the module. The copper sheets are
designed for thermal contact between flat heat pipe and battery to enhance the heat transfer surface.
CFD model is conducted to analyze the temperature distribution of the battery module. The results
show that all cooling designs not only decreases the maximum temperature of the module but also
increase the uniformity. The paper is organized as follows. Section 2 represents the objective of
the research. In Section 3, the experimental setup is represented. The simulation of the battery
module model is described in Section 4. Results, discussion, and optimization of the air-cooling
system are explained in Section 5. Lastly, a relevant conclusion is drawn in Section 6.

2. Objective of research
The appropriate selection of the design parameters as a TMS and the decision of their values were
the important factors in establishing control over the maximum battery’s module temperature and
performance. The design parameters and their range values were selected for developing a more
successful cooling system based on the review of previous works. In this work, the hybrid cooling
system studied experimentally and numerically. In experimental tests the effect of natural and
forced convection considered on the module temperature and the temperature uniformity. For
further consideration validate the simulation with experimental data. Afterward, various design
parameters, such as the different inlet velocity, different ambient temperatures, and different
spacing between the cells in the module have simulated. Moreover, the influence of adding heat
pipe and HPCS has been considered. All these parameters and designs have a high influence on
controlling the temperature rise and maintaining uniform heat distribution between the cells in the
module. The effect of the above-mentioned parameters on Li-ion battery module temperature was
observed in various studies. However, there is no study on the effect of L shape heat pipe when
the condenser and evaporator located inside the module. Fig.1. shows the whole experimentation
and simulation were carried out as shown in the flowchart.

Fig. 1. Experimental and simulation flowchart

3. Experimental setup
3.1 Concept description of the battery module
Generally, the EVs’ battery module/pack consists of several cells that are connected in series or
parallel. The present work test setup consists of a battery module, a PVC case, a power supply, a
cooling fan, K-type thermocouples, a data logger, and a personal computer. To fulfill the module
model, the hardware-case is designed for 2 mm spacing between each cell. The experimental setup
is built to investigate the performance of the air cooling system for the thermal management of a
commercial 18650 lithium-ion battery. The schematic of the test bench, as well as the dimensions
of the battery module, can be seen in Fig. 2.
a b

Fig. 2. (a) Picture of the battery module and (b) the schematic of battery module dimensions

Moreover, inlet airflow and the outlet suction fan designed for the module are depicted in Fig. 3.

a b

Fig. 3. The picture of the battery module, (a) inlets airflow and (b) outlet suction fan
The battery module used in this work is comprised of 24 commercial cylindrical cells in parallel-
series connection. The electrical and thermophysical properties of the cells utilized in the battery
module are shown in Table 1.

Table 1. The electrical and thermophysical properties of the cell in the battery module

value
Parameter Value Parameter
2.2
Nominal voltage, (V) 3.6 Nominal capacity, (A)
1200
Mass, (kg) 0.045 Specific heat capacity, (J.kg-1.K-1)
18
Length, (mm) 65 Diameter, (mm)
kr=0.2, kz=37.6
Density, (kg.m-3) 2722 Thermal conductivity, (W.m.-1K-1) [41–44]

In the air cooling system, 54 holes (9 rows × 6 columns) were embedded in the inlet of the module
to improve the cooling effect. The diameter of each hole is 5 mm. The outlet is a single suction fan
with a diameter of 50 mm that controlled by an adaptor. According to the datasheet of the fan, each
voltage corresponds to a specific airflow velocity. The air enters through the inlet and flows out
when the fan works. The modeling approach is based on the 3D-temperature distribution model
for a module with an air cooling thermal management system. The initial temperature, inlet
velocity and cell spacing for the experiment are 26 °C, 2 m.s-1 and 2 mm respectively. As it is
mentioned, the forced air cooling is supplied by a suction fan. The specification parameters of the
fan as a cooling system are shown in Table 2.
Table 2. The main properties of the installed fan on the battery module

value
Parameter Value Parameter

Outlet diameter, (mm) 50


Module case size (length × width × 130×90×70
height), (mm)
2
Inlet air temperature, (℃) 26 Inlet air velocity at 12V, (m.s-1)

3.2 Experimental setup of the test-bench


Fig. 4 illustrates the schematic of the experimental setup for the thermal management test. The
position of inlet airflow, suction outflow, and location of both K-type thermocouples are
demonstrated to have a better understanding of the system. A battery tester is applied to the module
to charge/discharge the cells, which should be connected to a computer system. The Pico USB TC-
08 data logger with the accuracy of ±0.025 ℃ is used to record the temperature sensed by
thermocouples. During the charging/discharging process, heat is generated by the battery module.
Two K-type thermocouples are attached to the cells, as can be seen from Fig. 4(b), to measure the
temperature. All the thermocouples were calibrated, and the accuracy of the thermocouples was
founded to be ±0.2 ℃. Schultz and Cole [45,46] method have been used to calculate the
uncertainty.
2 1/2
[ 𝑛 ∂𝑅
𝑈𝑅 = ∑𝑖 = 1 (∂𝑉𝐼𝑈𝑉𝐼) ] (1)

where 𝑈𝑉𝐼 and 𝑈𝑅 are the partial and total errors respectively. The uncertainty, for temperature
recorded by thermocouples, data logger resolution, current, and voltage are ±0.2 ℃, 0.025 ℃,
±1%, and ±1% respectively. For all cases, the maximum absolute uncertainty was less than 2.22%.
The ambient and initial temperature is 26 °C. Different discharge rates, internal electrochemical
reactions, as well as resistances, are the leading cause of the temperature rise in the module. The
generated heat (q) can be derived as follows [14,47]

[
𝑞 = 𝐼 (𝑂𝐶𝑉 ― 𝑉) ― 𝑇
∂𝑂𝐶𝑉
∂𝑇𝑎 ] (2)

where I , OCV , V , and T are the discharge current, open-circuit voltage, operating voltage and
∂𝑂𝐶𝑉
temperature of the battery, respectively. Also, ∂𝑇𝑎 represents the entropy coefficient which is
measured to obtain the heat generation of the battery. In the present study for 1.5C discharge rate,
a uniform and constant average heat source of 48750 W.m-3 is selected for the module simulation.

a b

Fig. 4.(a) The picture of the battery module, (b) and the schematic position of thermocouples

In order to build the test bench, the battery module is connected to the power supply. Two K-type
thermocouples, as can be seen in Fig. 4, are attached to the cells and the data logger. Moreover,
the data logger is connected to a personal computer to extract the results. The cooling fan
embedded in the PVC case. Fig. 5 depicts the picture of the experimental test bench.

a b c
Fig. 5. Experimental setup including (a) power supply, (b) battery module, (c) data logger and PC

4. Simulation of the battery module model

4.1 The battery module, heat pipe, and HPCS geometry model
In the present study, the commercial CFD software COMSOL Multiphysics® is employed to
numerically simulate the battery module to observe the temperature distribution across the module
surface. The numerical procedure is validated with the experimental results for both natural and
forced convection cooling. Besides, a detailed numerical analysis is performed to investigate the
effect of cell spacing, inlet temperature, and inlet velocity of the air on the thermal behavior of the
battery module with forced convection cooling. Finally, the potential of adding heat pipes to the
current battery module with forced convection cooling is studied to improve the temperature
uniformity. The INVENTOR® software is used to design the geometry of the battery module
including 24 cylindrical cells and a PVC case. The main properties of the PVC are mentioned in
Table 3. The CAD file is then imported into the COMSOL, and the computational grid is generated
through the domain. Fig. 6 illustrates the geometry of the battery module and the mesh generated
in the computational domain, respectively.

a b

Fig. 6.(a) Schematic of the battery module, (b) mesh distribution in the battery module
Table 3. The main properties of PVC

Value
Parameter value Parameter value Parameter
Thermal conductivity, Specific heat 600
0.1 Density,(kg.m-3) 100
(W.m.-1K-1) capacity, (J.kg-1.K-1)

For the heat pipe-air flow model, L shape flat heat pipe is designed. The evaporation section of the
heat pipe is contacted with the bottom of the cylindrical cells, where the thermal conductivity of
the cells in this direction is 37.6 W.m-1.K-1. Also, the condensation section of the heat pipe is
exposed to the forced-air. Typically, the working fluid in the evaporation section of the heat pipe
absorbs heat when the batteries are operating at 1.5C discharge rate. By getting the heat, the liquid
becomes steam and flows to the condensation section due to the pressure rise. Again, this steam is
cooled to liquid at the condensation section due to the emancipation of the absorbed heat to the air
coming by the fan. During this process, the liquid goes back by the capillary force of wick to the
evaporation section, and this cycle continues repeatedly. The width and thickness of the flat heat
pipe are designed 15.2 mm and 3.4 mm, respectively. According to the literature, the thermal
conductivity and the total resistance of the heat pipe are calculated as 8785 W.m-1.K-1 and 0.202
K.W-1, respectively [48,49]. Due to the high thermal conductivity of the heat pipe, the heat
generated in the battery module will be quickly taken away by the heat pipe. All details and
specifications of the flat heat pipe are shown in Table 4.
For further design, in the optimization section, the HPCS is produced for a better heat transfer in
each cell. Also, this optimization method improves temperature uniformity. Six copper sheets have
been embedded in each heat pipe for which the size of each copper sheet is 120˚ Φ 18.2 mm × 60
mm. The copper sheets are welded on the surface of the heat pipes.

Fig. 7. Schematic illustration and dimension of HPCS


4.2 Governing equations
The governing equations including continuity, momentum, and energy are expressed as
follows: [50]

 Continuity equation:

∂𝑢𝑖
∂𝑥𝑖 =0 (3)

where u is the air velocity.


 Momentum equation:
∂2𝑢𝑖
𝜌 ( ∂𝑢𝑖
∂𝑡
∂𝑢𝑖
) ∂𝑝
+ 𝑢𝑗∂𝑥𝑗 = ― ∂𝑥𝑖 + 𝜇∂𝑥 2 + 𝜌𝑔𝑖
𝑗
(4)

where 𝜌 is the density of air, 𝑝 is pressure, 𝜇 is the air viscosity and 𝑔𝑖 represents the considered
body force which is considered in natural convection.

 Energy equation:
∂2𝑇
𝜌𝐶𝑝( ∂𝑇 ∂𝑇
)
∂𝑡 + 𝑢𝑗∂𝑥𝑗 = 𝜆∂𝑥 2 + 𝑞𝑗
(5)

where 𝐶𝑝 and 𝜆 is the specific heat capacity and heat generation, respectively. The 𝑞 denotes the
volumetric heat generation rate in battery cells which is calculated as follow: [50]
𝑅𝐼2
𝑞= ∀
(6)

where 𝑅 is the total internal resistance of the cell, 𝐼 represents the electrical current and, ∀ is the
volume of the battery cells.

4.3 Thermal analysis of the heat pipe


The total thermal resistance of a heat pipe, 𝑅𝑡, includes several partial resistances arranged between
the heat source and heat sink [51–53]. Fig. 8 displays the equivalent thermal resistance of a heat
pipe.
Axisymmetric

Pipe core R3
R9 R6

Wick R2
R8 R5
R4
Pipe wall R1
R7

Condenser Evaporator

Fig. 8. Thermal resistance network of the flat heat pipe

The process for thermal analysis is as follows. The rate of heat transfer from the heat source to the
surrounding air (Q) is stated as [26]:
∆𝑇
𝑄= 𝑅𝑡
(7)

where the ∆𝑇 and 𝑅𝑡 are the temperature difference of the evaporator and condenser surface of the
heat pipe and the total resistances respectively. The thermal resistance at the pipe wall and wick
(R1, R7, R2, and R8) can be calculated as [26]:
𝐿
𝑅 = A.k (8)

where the 𝐿 , 𝐴, and 𝑘 are the thickness of pipe wall/wick, the area of the heat source/sink and
thermal conductivity of pipe wall/wick respectively. To simplification of the thermal resistance
network of a heat pipe, the axial resistances comprising R4 and R5 can be considered infinite, while
R3, R6, and R9 may be assumed to be negligible due to the comparative magnitudes of the resistance
of the vapor space and axial resistances of the wick and pipe wall [26].
To investigate the ability of the heat pipes to conduct heat, the effective thermal conductivity of
the heat pipe is presented and calculated as [26]:
𝐿𝑒𝑓𝑓
𝑘𝑒𝑓𝑓 = 𝐴.𝑅𝑡 (9)

where A is the cross-section of the heat pipe and 𝐿𝑒𝑓𝑓 is the effective transport length that is
presented as [26]:
𝐿𝑒 + 𝐿𝑐
𝐿𝑒𝑓𝑓 = 2 + 𝐿𝑎 (10)

where Le, La and Lc are evaporator, adiabatic and condenser length of the heat pipe, respectively.

It is noted that the essential specifications of the designed flat heat pipe are presented in Table 4.
Table 4. The main properties of the heat pipe

Value
Parameter
Copper
Heat pipe material
0.5 mm
Pipe wall thickness
0.5 mm
Pipe wick thickness
180.3 mm
Heat pipe length

Flat heat pipe 15.2×3.4 mm (width × thickness)

Water
Working fluid
Sintered copper powder
Wick structure
8785, (W.m-1.K-1)
Thermal conductivity
0.202, (K.W-1)
Total Resistance
3.69, (W.m-1.K-1) [49]
Wick thermal conductivity

4.4 Boundary and initial conditions


According to the ambient experimental temperature, the initial temperature for the module, and
the airflow in the simulation is set to 26 °C. Moreover, the pressure outlet and the inlet air velocity
have been selected as the boundary conditions in the numerical calculations. At the module
surfaces, a no-slip wall has been chosen. The incompressible airflow is assumed. The laminar flow
model is selected for the numerical calculation. The discharge time for natural convection lasts
2740 second whereas the step size is set to 1 second. In the present study, the user-defined function
is used to describe the heat generation of the module. The heat transfer with the surroundings is
determined by following the convection equation:
𝑄𝑙𝑜𝑠𝑠 = 𝐴.ℎ.(𝑇 ― 𝑇𝑎𝑚𝑏) (11)

where 𝑄𝑙𝑜𝑠𝑠, 𝐴, ℎ, 𝑇, and 𝑇𝑎𝑚𝑏 demonstrate the power loss of the battery, the surface area, heat
transfer coefficient, battery temperature, and ambient temperature, respectively.

4.5 Meshing and grid independence analysis


The precision and quickness of calculation are highly dependent on the number of meshes and
solver. According to the high non-linearity of the governing equations and because of the different
geometrical scales in the presented model, the simulation process is very time-consuming. The
temperature of the middle cell specified by T2 is used to measure the independence of the grid
number. The test results are shown in Fig. 9. When the grid number varies from 977,654 to
1,594,076, the results differ from 0.992%. To save the calculation time and increase the efficiency,
the grid number of 977,654 is chosen for the module simulation of the forced cooling.
36

35.5

Temperature(℃)
35

34.5

T2 temperature
34
0 5 10 15 20 25

Grid number ×100000

Fig.9. Grid number independency test

4.6 Validation of the thermal model


With the purpose of validation to show the accuracy of the numerical method, the temperature of
the thermocouples of T1 and T2 during discharging mode (1.5C) for forced and natural convection
are compared with the simulation results. As schematically explained earlier, the presented module
has been simulated in COMSOL Multiphysics® software. The air velocity is set to 2 m.s-1 and cell
spacing of 2 mm used for model validation. The experimental temperature was measured by
attaching two K-type thermocouples in the middle surface of the cells, and the discharge process
finished when the voltage of the module reached the manufacturer's recommended voltage for each
cell. The comparison of simulation and experimental data is shown in Fig. 10(a,b) and illustrates
acceptable agreement with each other during the discharge process. As stated in Fig. 10a, for
natural convection and under constant 1.5C rate discharge, the highest temperature of the battery
occurs at the end of the discharge process and for T1 and T2 reaches 59.2 ℃ and 65.1 ℃,
respectively. Moreover, in Fig. 10b for the forced convection, the highest temperature for T1 and
T2 reaches to 31.4 ℃ and 35.9 ℃, respectively. It is also obvious in Fig.10b and at the end of the
forced convection experimental test, the temperature of T1 and T2 increased a little. As in the last
10% of the discharge, resistance always increases [54], an increase in the temperature can happen
at the last minutes of discharging. Therefore, as the temperature for T1 and T2 increase in the last
minutes is less than 1.5 ℃, the error is relatively acceptable.
a b
70 40

60

Temperature(℃)
50 30
Temperature(℃)

40

30 20
1.5C experimental data(T1) 1.5C experimental data(T1)

1.5C simulation data(T1) 1.5C simulation data(T1)


20
1.5C experimental data(T2) 1.5C simulation data(T2)

1.5C simulation data(T2) 1.5C experimental data(T2)


10 10
0 1000 2000 3000 0 1000 2000 3000

Time(s) Time(s)

Fig. 10. Thermal model validation of (a) natural convection and (b) forced convection

5. Results, discussion, and optimization of air cooling system

5.1 Natural convection cooling performance

Fig. 11 shows the contour of temperature distribution for modules with a natural air-cooling
strategy. It can be seen that with natural convection, the maximum and minimum temperature of
the cells are around 64.8 °C and 57 °C, respectively. In this case, the battery temperature is much
higher than the optimal range for lithium-ion batteries (between 15 °C to 40 °C), which can
significantly affect the lifetime, performance, and safety of the battery cells negatively.
Furthermore, the temperature distribution of the module is non-uniform, where the temperature of
the inner cells is higher than the outer cells located near the module casing. This non-uniformity
of temperature distribution can lead to a performance reduction of the module. Therefore, there is
a crucial need for employing a forced-air cooling strategy to enhance the lifespan and performance
of the module and to satisfy the safety needs.
Fig. 11. Temperature distribution of natural convection

5.2 Forced-air cooling approach with cell spacing

In this section, the effect of cell spacing at 0, 2, and 4 mm in constant boundary conditions is
considered. Moreover, the maximum temperatures of the battery module and cell temperature
uniformity are investigated. Fig. 12 demonstrates the temperature contours of three cell spacing at
the 1.5C rate discharging processes when the inlet velocity is 2 m.s-1, and the initial temperature
is 26 ℃. It can be seen when the cells stick together, the maximum module temperature reaches to
66.7 ℃ due to the low heat transfer coefficient for central cells. Moreover, temperature uniformity
is very low. In 2 mm spacing, the maximum temperature decreases tremendously, and temperature
uniformity reaches to more acceptable manner. By increasing the cell spacing to 4 mm the
maximum temperature increases 1.5 ℃ compared with 2 mm spacing. It means that increasing the
space between the cells to 4 mm only grows the system volume and the following flow channel is
not conducive to the formation of flow turbulence, which also decreases the heat transfer
coefficient. For the 4 × 6 cylindrical battery module, the spacing value of 2 mm is more suitable
considering heat dissipation and volume.
a b c

Fig. 12. Temperature distribution of cell spacing (a) 0 mm, (b) 2 mm, and (c) 4 mm

5.3 Forced-air cooling approach in different ambient temperatures

The influence of the ambient temperature of the air on the temperature distribution of the module
is shown in Fig. 13. For this purpose, three different ambient temperatures, including 10 °C, 26
°C, and 35 °C are considered. It is clear that the ambient temperature of the air significantly affects
the temperature of the cells. A lower ambient temperature results in the lower temperature of the
cells, while the temperature distribution for all three cases remains the same.
a b c

Fig. 13. Temperature distribution of ambient temperature of (a) 10 ℃, (b) 26 ℃, (c) 35 ℃

5.4 Forced-air cooling approach with different inlet velocities

In this section, the effect of different inlet velocities is examined to improve cooling and
temperature uniformity in the module. In this parametric numerical study, different inlet velocities
are entered into the simulation software to conduct comparative analysis on the relationship
between inlet velocity and maximum module temperature. By setting the velocity of the air to 1
m.s-1, 2 m.s-1, and 3 m.s-1, the maximum temperature evolution stands on 51.1 °C, 42.4 °C, and 39
°C, respectively. It can be concluded that by increasing the velocity of the airflow, the local
temperature diminishes. Hence, there is an inverse relationship between the inlet flow velocity
through the entire passage and the maximum temperature of the battery module. The contours of
velocity for the different cases are depicted in Fig. 14. Moreover, it should be added that if the
ambient temperature is higher than the standard range, the inlet air velocity should be enhanced to
achieve effective and uniform cooling.

a b c

Fig. 14. Temperature distribution of air


velocity (a) 1m.s-1, (b) 2m.s-1, (c) 3m.s-1
5.5 Forced air cooling approach with heat pipe and HPCS

A combination of heat pipe with different cooling systems is common for prismatic cells.
However, due to design efforts, the influence of such systems has not been widely used for
cylindrical cells. In this section, the temperature of cylindrical cells is monitored using the new
design of L shape flat heat pipe and HPCS. The stated designs of heat pipe and HPCS are shown
in Fig. 15. As it is mentioned the HPCS consists of one L shape flat heat pipe and six copper sheets.
Copper sheets with size in 120˚ Φ 18.2 × 60 mm are welded on the surface of the heat pipe to
increase the effective contact area of each cell and heat pipe. To prevent the short circuit, thin
thickness (0.2 mm) of thermal interface material [55] is embedded between the cells and the heat
pipe.

a b

Fig. 15. Schematic illustration of battery module with (a) heat pipe and (b,c) HPCS

5.5.1. Validation of the heat pipe cooling system

In order to validate the accuracy of the numerical method, the experimental [49] wall temperature
distribution along the heat pipe is compared with the simulation results. As schematically
explained earlier, the presented module has been simulated in COMSOL Multiphysics® software.
The air velocity is set to 2 m.s-1 and cell spacing of 2 mm used for model validation. It is important
to mention that the parameters of the heat pipe in experimental results are in accordance with the
simulation.

55

50

45
Temperature (℃)

40

35

30

25
experimental data
simulation data
20
0 200 400
Position along heat pipe (mm)

Fig. 16. Thermal model validation of heat pipe cooling system

According to the experimental data the heat pipe wall temperature is recorded along heat pipe
length in 20, 70, 150 and 210 mm respectively. Since the length of the simulated heat pipe is 180.3
mm the wall temperature of the heat pipe in 20, 70, 150 and 180.3 mm compared with experimental
results. As can be seen in Fig. 16 there is an excellent trend agreement between the experimental
data and simulation. It is necessary to mention that the difference in temperature is due to the
different heat input of the evaporator.

5.5.2. Comparison results

To evaluate the effect of the heat pipe and HPCS on the cooling performance, the temperature
distribution contour of the module is shown in Fig.17(a,b). As can be seen, the maximum
temperature of the module compared with forced convection decreases by 4.9 ℃ and 5.3 ℃ for the
heat pipe and HPCS, respectively. Moreover, increasing the effective contact area of each cell and
heat pipe by copper sheet, helps to improve the heat transfer by a heat pipe, decreases the module
temperature, and has better temperature uniformity along with the whole module.
a b

Fig. 17. Temperature distribution of battery module for a)HP and b)HPCS

As expected, in Fig. 18, the maximum temperature of the module decreases visibly with the heat
pipe and HPCS. Moreover, the temperature uniformity in the whole of the module increases
tremendously. In Fig. 18(a) in same boundary conditions(v=2 m.s-1, S=2 mm, T=26 ℃), the
maximum temperature of the battery module in natural convection, forced convection, with the
heat pipe, and HPCS reaches to the 64.8 ℃, 42.4 ℃, 37.5 ℃, and 37.1 ℃ respectively. The
temperature uniformity within the module affects its charging and discharging power. Moreover,
it is vital for optimum performance of the module. In this case, the temperature uniformity is the
temperature difference between T1 and T2. In Fig.18(b) the temperature uniformity for HPCS, heat
pipe and forced convection compare with natural convection improved by 73.4%, 66.5%, and
39.2%, respectively. It is essential to mention that, the size of the module for the heat pipe and
HPCS only grow 7 mm (in the x-axis) next to the inlet that does not affect the cell temperatures in
different boundary conditions.
a b
70 10
NC, 64.8 ℃
9 NC,8.66℃
60
8
50
Temperature(℃)

7
FC, 42.4 ℃

Temperature(℃)
HP, 37.5 ℃ HPCS, 37.1 6 FC, 5.26 ℃
40 ℃
5
30
4
HP, 2.9 ℃
20 3 HPCS, 2.3 ℃

2
10
1

0 0
Tmax T2 – T1
1 1

Fig. 18.(a) Maximum temperature of module and (b) Temperature difference of T1 and T2 for natural convection (NC),
forced convection (FC), heat pipe (HP) and HPCS

Fig. 19 shows the temperature variation in the direction of the cut line in the same boundary
condition. For natural convection, the temperature variation of the cells starts from 55 ℃ for the
first cell and reaches a constant peak of 64.8 ℃ for cell number of 2 to 5. In the end, the temperature
line decreases again to 55 ℃ for cell number 6. As it is evident, the temperature of outer cells is
lower than inner cells because of the effect of natural convection on the PVC case. For the forced
convection, the temperature of cells from inlet to outlet gradually increases and reaches 42.4 ℃.
Because the temperature of air increases as the air from the inlet moves to the outlet. As can be
seen in part c, adding HPCS has a significant effect on controlling the temperature of the cells.
The maximum temperature reaches to the 37.1 ℃. More importantly, it has a marvelous effect on
the uniformity of cell temperatures.
a

b
c

Fig. 19. Temperature uniformity graph of the cut line (H=35 mm) for (a) natural convection,(b) Forced convection, and (c) HPCS

6. Conclusion

In this paper, an optimization study to improve cooling and temperature uniformity in a 18650
battery module was developed and examined. A number of optimization scenarios were developed.
In the first scenario, the temperature of the module considered experimentally in natural and forced
convection. In the second scenario, the effect of cell spacing, ambient temperature and air velocity
numerically considered to improve cooling and temperature uniformity. Finally, in the third
scenario, the effect of adding the heat pipe and HPCS has been considered. According to the
results, the temperature uniformity for HPCS, heat pipe, and forced convection compared with
natural convection improved by 73.4%, 66.5%, and 39.2%, respectively. Moreover, the maximum
temperature of the battery module compared with the natural convection was reduced by 42.7%,
42.1%, and 34.5% for HPCS, heat pipe, and forced convection respectively.

Acknowledgment
I wish to express my sincere gratitude to Flanders Make for support to our research team.

References
[1] M. Behi, S.A. Mirmohammadi, M. Ghanbarpour, H. Behi, B. Palm, Evaluation of a novel
solar driven sorption cooling/heating system integrated with PCM storage compartment,
Energy. 164 (2018) 449–464. doi:10.1016/J.ENERGY.2018.08.166.
[2] J. Teter, P. Le Feuvre, M. Gorner, S. Scheffer, Iea website, Traking Transport, May 2019.
(n.d.). https://www.iea.org/reports/tracking-transport-2019 (accessed January 11, 2020).
[3] F.H. Gandoman, A. Ahmadi, P. Van den Bossche, J. Van Mierlo, N. Omar, A.E. Nezhad,
H. Mavalizadeh, C. Mayet, Status and future perspectives of reliability assessment for
electric vehicles, Reliab. Eng. Syst. Saf. 183 (2019) 1–16.
doi:10.1016/J.RESS.2018.11.013.
[4] A. Wei, J. Qu, H. Qiu, C. Wang, G. Cao, Heat transfer characteristics of plug-in
oscillating heat pipe with binary-fluid mixtures for electric vehicle battery thermal
management, Int. J. Heat Mass Transf. 135 (2019) 746–760.
doi:10.1016/J.IJHEATMASSTRANSFER.2019.02.021.
[5] X. Peng, S. Chen, A. Garg, N. Bao, B. Panda, A review of the estimation and heating
methods for lithium-ion batteries pack at the cold environment, Energy Sci. Eng. 7 (2019)
645–662. doi:10.1002/ese3.279.
[6] T. Gao, Z. Wang, S. Chen, L. Guo, Hazardous characteristics of charge and discharge of
lithium-ion batteries under adiabatic environment and hot environment, Int. J. Heat Mass
Transf. 141 (2019) 419–431. doi:10.1016/J.IJHEATMASSTRANSFER.2019.06.075.
[7] A. Tomaszewska, Z. Chu, X. Feng, S. O’Kane, X. Liu, J. Chen, C. Ji, E. Endler, R. Li, L.
Liu, Y. Li, S. Zheng, S. Vetterlein, M. Gao, J. Du, M. Parkes, M. Ouyang, M. Marinescu,
G. Offer, B. Wu, Lithium-ion battery fast charging: A review, ETransportation. 1 (2019)
100011. doi:10.1016/J.ETRAN.2019.100011.
[8] S. Arora, Selection of thermal management system for modular battery packs of electric
vehicles: A review of existing and emerging technologies, J. Power Sources. 400 (2018)
621–640. doi:10.1016/J.JPOWSOUR.2018.08.020.
[9] T. Yang, N. Yang, X. Zhang, G. Li, Investigation of the thermal performance of axial-
flow air cooling for the lithium-ion battery pack, Int. J. Therm. Sci. 108 (2016) 132–144.
doi:10.1016/J.IJTHERMALSCI.2016.05.009.
[10] F. Bahiraei, A. Fartaj, G.-A. Nazri, Electrochemical-thermal Modeling to Evaluate Active
Thermal Management of a Lithium-ion Battery Module, Electrochim. Acta. 254 (2017)
59–71. doi:10.1016/J.ELECTACTA.2017.09.084.
[11] L.H. Saw, Y. Ye, A.A.O. Tay, W.T. Chong, S.H. Kuan, M.C. Yew, Computational fluid
dynamic and thermal analysis of Lithium-ion battery pack with air cooling, Appl. Energy.
177 (2016) 783–792. doi:10.1016/J.APENERGY.2016.05.122.
[12] R. Sabbah, R. Kizilel, J.R. Selman, S. Al-Hallaj, Active (air-cooled) vs. passive (phase
change material) thermal management of high power lithium-ion packs: Limitation of
temperature rise and uniformity of temperature distribution, J. Power Sources. 182 (2008)
630–638. doi:10.1016/J.JPOWSOUR.2008.03.082.
[13] K. Somasundaram, E. Birgersson, A.S. Mujumdar, Thermal–electrochemical model for
passive thermal management of a spiral-wound lithium-ion battery, J. Power Sources. 203
(2012) 84–96. doi:10.1016/J.JPOWSOUR.2011.11.075.
[14] J. Zhao, Z. Rao, Y. Huo, X. Liu, Y. Li, Thermal management of cylindrical power battery
module for extending the life of new energy electric vehicles, Appl. Therm. Eng. 85
(2015) 33–43. doi:10.1016/J.APPLTHERMALENG.2015.04.012.
[15] X. Li, F. He, L. Ma, Thermal management of cylindrical batteries investigated using wind
tunnel testing and computational fluid dynamics simulation, J. Power Sources. 238 (2013)
395–402. doi:10.1016/J.JPOWSOUR.2013.04.073.
[16] M. Soltani, G. Berckmans, J. Jaguemont, J. Ronsmans, S. Kakihara, O. Hegazy, J. Van
Mierlo, N. Omar, Three dimensional thermal model development and validation for
lithium-ion capacitor module including air-cooling system, Appl. Therm. Eng. 153 (2019)
264–274. doi:10.1016/J.APPLTHERMALENG.2019.03.023.
[17] T. Wang, K.J. Tseng, J. Zhao, Z. Wei, Thermal investigation of lithium-ion battery
module with different cell arrangement structures and forced air-cooling strategies, Appl.
Energy. 134 (2014) 229–238. doi:10.1016/J.APENERGY.2014.08.013.
[18] K. Yu, X. Yang, Y. Cheng, C. Li, Thermal analysis and two-directional air flow thermal
management for lithium-ion battery pack, J. Power Sources. 270 (2014) 193–200.
doi:10.1016/j.jpowsour.2014.07.086.
[19] X. Peng, C. Ma, A. Garg, N. Bao, X. Liao, Thermal performance investigation of an air-
cooled lithium-ion battery pack considering the inconsistency of battery cells, Appl.
Therm. Eng. 153 (2019) 596–603. doi:10.1016/J.APPLTHERMALENG.2019.03.042.
[20] D. Dan, C. Yao, Y. Zhang, H. Zhang, Z. Zeng, X. Xu, Dynamic thermal behavior of micro
heat pipe array-air cooling battery thermal management system based on thermal network
model, Appl. Therm. Eng. 162 (2019) 114183.
doi:10.1016/J.APPLTHERMALENG.2019.114183.
[21] M.R. Giuliano, A.K. Prasad, S.G. Advani, Experimental study of an air-cooled thermal
management system for high capacity lithium-titanate batteries, J. Power Sources. 216
(2012) 345–352. doi:10.1016/j.jpowsour.2012.05.074.
[22] S.K. Mohammadian, Y. Zhang, Thermal management optimization of an air-cooled Li-ion
battery module using pin-fin heat sinks for hybrid electric vehicles, J. Power Sources. 273
(2015) 431–439. doi:10.1016/J.JPOWSOUR.2014.09.110.
[23] X. Li, F. He, G. Zhang, Q. Huang, D. Zhou, Experiment and simulation for pouch battery
with silica cooling plates and copper mesh based air cooling thermal management system,
Appl. Therm. Eng. 146 (2019) 866–880.
doi:10.1016/J.APPLTHERMALENG.2018.10.061.
[24] H. Behi, M. Ghanbarpour, M. Behi, Investigation of PCM-assisted heat pipe for electronic
cooling, Appl. Therm. Eng. 127 (2017) 1132–1142.
doi:10.1016/J.APPLTHERMALENG.2017.08.109.
[25] H. Behi, R.Khodabandeh, M.Ghanbarpour, Experimental and Numerical Study on Heat
Pipe Assisted PCM Storage System, (2014).
[26] M. Ghanbarpour, Investigation of Thermal Performance of Cylindrical Heat pipes
Operated with Nanofluids, 2017.
[27] N.K. Gupta, A.K. Tiwari, S.K. Ghosh, Heat transfer mechanisms in heat pipes using
nanofluids – A review, Exp. Therm. Fluid Sci. 90 (2018) 84–100.
doi:10.1016/J.EXPTHERMFLUSCI.2017.08.013.
[28] Z.Y. Jiang, Z.G. Qu, Lithium–ion battery thermal management using heat pipe and phase
change material during discharge–charge cycle: A comprehensive numerical study, Appl.
Energy. 242 (2019) 378–392. doi:10.1016/J.APENERGY.2019.03.043.
[29] Q. Huang, X. Li, G. Zhang, J. Zhang, F. He, Y. Li, Experimental investigation of the
thermal performance of heat pipe assisted phase change material for battery thermal
management system, Appl. Therm. Eng. 141 (2018) 1092–1100.
doi:10.1016/J.APPLTHERMALENG.2018.06.048.
[30] X. Ye, Y. Zhao, Z. Quan, Experimental study on heat dissipation for lithium-ion battery
based on micro heat pipe array (MHPA), Appl. Therm. Eng. 130 (2018) 74–82.
doi:10.1016/J.APPLTHERMALENG.2017.10.141.
[31] Y. Ye, L.H. Saw, Y. Shi, A.A.O. Tay, Numerical analyses on optimizing a heat pipe
thermal management system for lithium-ion batteries during fast charging, Appl. Therm.
Eng. 86 (2015) 281–291. doi:10.1016/J.APPLTHERMALENG.2015.04.066.
[32] J. Liang, Y. Gan, Y. Li, M. Tan, J. Wang, Thermal and electrochemical performance of a
serially connected battery module using a heat pipe-based thermal management system
under different coolant temperatures, Energy. 189 (2019) 116233.
doi:10.1016/J.ENERGY.2019.116233.
[33] J. Liang, Y. Gan, Y. Li, Investigation on the thermal performance of a battery thermal
management system using heat pipe under different ambient temperatures, Energy
Convers. Manag. 155 (2018) 1–9. doi:10.1016/j.enconman.2017.10.063.
[34] G. Burban, V. Ayel, A. Alexandre, P. Lagonotte, Y. Bertin, C. Romestant, Experimental
investigation of a pulsating heat pipe for hybrid vehicle applications, Appl. Therm. Eng.
50 (2013) 94–103. doi:10.1016/J.APPLTHERMALENG.2012.05.037.
[35] Z. Rao, S. Wang, M. Wu, Z. Lin, F. Li, Experimental investigation on thermal
management of electric vehicle battery with heat pipe, Energy Convers. Manag. 65 (2013)
92–97. doi:10.1016/J.ENCONMAN.2012.08.014.
[36] T.-H. Tran, S. Harmand, B. Desmet, S. Filangi, Experimental investigation on the
feasibility of heat pipe cooling for HEV/EV lithium-ion battery, Appl. Therm. Eng. 63
(2014) 551–558. doi:10.1016/J.APPLTHERMALENG.2013.11.048.
[37] M.-S. Wu, K.H. Liu, Y.-Y. Wang, C.-C. Wan, Heat dissipation design for lithium-ion
batteries, J. Power Sources. 109 (2002) 160–166. doi:10.1016/S0378-7753(02)00048-4.
[38] L. Feng, S. Zhou, Y. Li, Y. Wang, Q. Zhao, C. Luo, G. Wang, K. Yan, Experimental
investigation of thermal and strain management for lithium-ion battery pack in heat pipe
cooling, J. Energy Storage. 16 (2018) 84–92. doi:10.1016/J.EST.2018.01.001.
[39] Y. Gan, J. Wang, J. Liang, Z. Huang, M. Hu, Development of thermal equivalent circuit
model of heat pipe-based thermal management system for a battery module with
cylindrical cells, Appl. Therm. Eng. 164 (2020) 114523.
doi:10.1016/J.APPLTHERMALENG.2019.114523.
[40] J. Wang, Y. Gan, J. Liang, M. Tan, Y. Li, Sensitivity analysis of factors influencing a heat
pipe-based thermal management system for a battery module with cylindrical cells, Appl.
Therm. Eng. 151 (2019) 475–485. doi:10.1016/j.applthermaleng.2019.02.036.
[41] S.J. Drake, D.A. Wetz, J.K. Ostanek, S.P. Miller, J.M. Heinzel, A. Jain, Measurement of
anisotropic thermophysical properties of cylindrical Li-ion cells, J. Power Sources. 252
(2014) 298–304. doi:10.1016/j.jpowsour.2013.11.107.
[42] L.H. Saw, Y. Ye, A.A.O. Tay, Electrochemical-thermal analysis of 18650 Lithium Iron
Phosphate cell, Energy Convers. Manag. 75 (2013) 162–174.
doi:10.1016/j.enconman.2013.05.040.
[43] R. Mazurick, Novel 18650 lithium-ion battery surrogate cell design with anisotropic,
(2016).
[44] E. Jiaqiang, M. Yue, J. Chen, H. Zhu, Y. Deng, Y. Zhu, F. Zhang, M. Wen, B. Zhang, S.
Kang, Effects of the different air cooling strategies on cooling performance of a lithium-
ion battery module with baffle, Appl. Therm. Eng. 144 (2018) 231–241.
doi:10.1016/j.applthermaleng.2018.08.064.
[45] M. Sheikholeslami, D.D. Ganji, Heat transfer improvement in a double pipe heat
exchanger by means of perforated turbulators, Energy Convers. Manag. 127 (2016) 112–
123. doi:10.1016/J.ENCONMAN.2016.08.090.
[46] M. Sheikholeslami, D.D. Ganji, Heat transfer enhancement in an air to water heat
exchanger with discontinuous helical turbulators; experimental and numerical studies,
Energy. 116 (2016) 341–352. doi:10.1016/J.ENERGY.2016.09.120.
[47] R. Mahamud, C. Park, Reciprocating air flow for Li-ion battery thermal management to
improve temperature uniformity, J. Power Sources. 196 (2011) 5685–5696.
doi:10.1016/J.JPOWSOUR.2011.02.076.
[48] M. Ghanbarpour, R. Khodabandeh, Entropy generation analysis of cylindrical heat pipe
using nanofluid, Thermochim. Acta. 610 (2015) 37–46. doi:10.1016/J.TCA.2015.04.028.
[49] M.H.A. Elnaggar, M.Z. Abdullah, S. Raj, R. Munusamy, Experimental and Numerical
Studies of Finned L-Shape Heat Pipe for Notebook-PC Cooling, IEEE Trans.
Components, Packag. Manuf. Technol. 3 (2013) 978–988.
doi:10.1109/TCPMT.2013.2245944.
[50] M. Soltani, J. Ronsmans, S. Kakihara, J. Jaguemont, P. Van den Bossche, J. van Mierlo,
N. Omar, Hybrid battery/lithium-ion capacitor energy storage system for a pure electric
bus for an urban transportation application, Appl. Sci. 8 (2018). doi:10.3390/app8071176.
[51] N. Zhu, K. Vafai, Analysis of cylindrical heat pipes incorporating the effects of liquid–
vapor coupling and non-Darcian transport—a closed form solution, Int. J. Heat Mass
Transf. 42 (1999) 3405–3418. doi:10.1016/S0017-9310(99)00017-4.
[52] Peterson, G. P., (1994). Heat pipes; modeling, testing, and applications, John Wiley &
Sons., (n.d.).
[53] Reay, D. and Kew, P., (2006). Heat Pipes, 5th Edition, Elsevier., (n.d.).
[54] A. Nikolian, Y. Firouz, R. Gopalakrishnan, J.M. Timmermans, N. Omar, P. van den
Bossche, J. van Mierlo, Lithium ion batteries-development of advanced electrical
equivalent circuit models for nickel manganese cobalt lithium-ion, Energies. 9 (2016) 360.
doi:10.3390/en9050360.
[55] D. Mao, J. Chen, L. Ren, K. Zhang, M.M.F. Yuen, X. Zeng, R. Sun, J.-B. Xu, C.-P.
Wong, Spherical core-shell Al@Al2O3 filled epoxy resin composites as high-performance
thermal interface materials, Compos. Part A Appl. Sci. Manuf. 123 (2019) 260–269.
doi:10.1016/J.COMPOSITESA.2019.05.024.
Highlights

 A heat pipe copper sheets (HPCS) for the battery thermal management system is designed.

 A thermal management system (TMS) for cylindrical cells is simulated and validated with
experiments.

 Temperature variations of a cylindrical battery module in different initial conditions are


reported.

 39.2%, 66.5%, and 73.4% improvement in the temperature uniformity of the battery
module for forced-air cooling, heat pipe, and HPCS are reported respectively.
Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

You might also like