Skip to main content

Open Access 02.12.2024 | Electric, Fuel Cell, and Hybrid Vehicle, Fuels and Lubricants, Heat Transfer, Fluid and Thermal Engineering, Vision and Sensors

Study of Cooling Performance of Liquid-Cooled EV Battery Module According to the TIM Compression Ratio

verfasst von: Jahun Gu, Heung-Kyu Kim, Siyoul Jang

Erschienen in: International Journal of Automotive Technology

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

This study examines the coolant and heat flows in electric vehicle (EV) battery pack that employs a thermal interface material (TIM). The overall temperature distribution of the battery pack that consists of many battery modules is precomputed based on the cooling circuit design, and the battery module that is most strongly influenced by cooling circuit is selected. To investigate the detailed effects of the TIM’s performance, we measure its thermal conductivity based on its compression ratio and consider the detailed shape of the battery cell module for incorporating the TIM’s thermal conductivity in the battery assembly. The TIM, which functions as part of the cooling system within the confined space of the battery cell module, is included in the CFD analysis to assess the effect of thermal conductivity variation resulting from the compression ratio of the TIM on the cooling performance of the battery pack. Various cases of confined spaces in the battery module are compared to optimize the cooling efficiency of the EV battery pack.
Hinweise

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Abkürzungen
BLT
Bond line thickness, m
C P
Specific heat, J/kg k
g
Gravitational acceleration, m/s2
\({H}_{\alpha }\)
Compressed TIM height unaffected by shapes of part A and B, mm
\({H}_{\beta }\)
Compressed TIM height influenced by shapes of part A and B, mm
h
Heat transfer coefficient, w/m2 k
I
Circuit current, A
k
Thermal conductivity, W/m k
\({k}_{\text{TIM}}\)
Thermal conductivity of TIM, W/m k
\({k}_{\alpha }\)
Thermal conductivity of \(\alpha\)
\({k}_{\beta }\)
Thermal conductivity of \(\beta\)
p
Pressure, Pa
Q
Heating value, W
R
Internal equivalent resistance, \(\Omega\)
\({R}_{\text{bulk}}\)
Contact thermal resistance, K/W
T
Temperature, °C
\({T}_{\infty }\)
Surrounding temperature, °C
t
Time, s
U OC
Open circuit voltage, V
\(v\)
Velocity, m/s
ε
Compression ratio
\(\epsilon \left({H}_{\alpha }\right)\)
Compression ratio when TIM is compressed unaffected by shapes of part A and B
\(\epsilon \left({H}_{\beta }\right)\)
Compression ratio when TIM is compressed influenced by shapes of part A and B
μ
Viscosity, Pa s
ρ
Density, kg/m3
Φ
Dissipation function
τ
Deviatoric stress tensor, Pa
δ 0
Initial TIM thickness before compressed, mm
ΔH TIM
Difference between initial TIM thickness and compressed TIM height
\(\sigma\)
Electrical conductivity, S

1 Introduction

Worldwide, the issues of energy shortage and environmental pollution have become prominent, leading the automotive industry to focus extensively on the development of environmentally friendly vehicles such as hybrid electric vehicles (HEVs) and electric vehicles (EVs), instead of traditional internal combustion engine (ICE) vehicles. Consequently, the automobile industry is transitioning from using conventional fossil fuels to batteries as the power source. HEVs and EVs require large amounts of power to drive electric motors during vehicle acceleration, thereby necessitating the use of lithium-ion batteries, which are characterized by high energy density and low self-discharge. In this process, current consumption is significant, resulting in the generation of a substantial amount of heat in the battery pack. Prolonged exposure to heat can degrade the battery’s lifespan and cause swelling, that is, expansion of the battery cells owing to gas generation inside them (Oh et al., 2014). This swelling is associated with uncontrollable escalation of abnormal heat, leading to thermal runaway, which can ultimately cause fires in the entire battery pack. Therefore, designing cooling systems for battery packs that are used as power sources in HEVs (Hybrid Electric Vehicles) and EVs (electric vehicles) is a critical challenge.
The cooling methods for the battery packs used in HEVs and EVs broadly include air cooling, phase change material (PCM)-based cooling, and liquid cooling. First, in air cooling, external ambient air is drawn in through a blower motor to cool the battery. This method is predominantly used in hybrid vehicles because the electric power requirement is relatively low owing to combined use of an engine and an electric power source. Although the density and specific heat of air are lower than those of liquids, the output of the electric power source used in HEVs is relatively modest compared to those used in pure EVs, and therefore, system design for cooling performance in HEVs is comparatively simple (Wang et al., 2016). Second, in the PCM-based cooling method, PCMs such as wax or paraffin are commonly used. A PCM transitions between the solid and liquid states by releasing or accumulating the heat generated by the battery. This facilitates maintenance of the battery pack temperature within the operational temperature range through the absorption and storage of excess heat or release of stored heat, as needed (Li et al., 2020). Finally, liquid cooling, which involves the utilization of an insulated liquid coolant, is used to cool batteries in pure EVs, which have high power demands. Depending on the installation location within the vehicle, liquid cooling systems are classified into two main types. Direct liquid cooling involves circulation of a coolant between battery cells to cool them directly (Larrañaga-Ezeiza et al., 2022). By contrast, in indirect liquid cooling, cooling plates installed beneath the battery cells are utilized to create a network of cooling channels that dissipates heat indirectly (Deng et al., 2018). This form of cooling is widely adopted in view of the specific requirements of EVs with high power output (Anisha & Kumar, 2023; Thakur et al., 2020).
In indirect liquid cooling, a gap exists between the cooling plate and the battery module inside the assembly of the battery pack. This gap can lead to noise generation and discontinuity of heat transfer, which would not warrant enhancement of the battery pack’s cooling performance (Viswanath et al., 2000). To address this issue, a thermal interface material (TIM) is applied between the cooling plate and battery module to improve contact thermal conductivity (Gwinn & Webb, 2003; Kumar Swamy & Satyanarayan, 2019).
TIMs are broadly classified into pad- and gel-type materials. While these materials are prone to deformation under pressure, they increase the contact area under compression, thus effectively filling the gap between the cooling plate and battery to improve heat transfer (Sarvar et al., 2006). Furthermore, contact thermal resistance decreases as the pressure applied at the junction between the two surfaces increases, as in Eq. (1) related to bond line thickness (BLT). According to this equation, as the contact thermal resistance decreases, thermal conductivity increases (Sarvar et al., 2006). Here, BLT represents the contact distance between two solid surfaces, as illustrated in Fig. 1.
$${R}_{\text{bulk}}=\frac{\text{BLT}}{{k}_{\text{TIM}}}.$$
(1)
We utilize compression-induced changes in the thermal conductivity of TIMs as an input parameter in our analysis. Thermal conductivity of TIMs for the computational works in this study is measured by following the ASTM-D5470 standard, wherein the heat generated upon compression of the TIM is measured. Variations in thermal conductivity with changes in the compression ratio of the TIM are observed. Based on these observations, a computational fluid dynamics (CFD) analysis is performed to simulate the changes in the thermal conductivity of the TIM with varying thickness under different compression ratios in a liquid-cooled battery module. The objective is to demonstrate the improvement in cooling performance without increasing the power consumption of the pump to provide additional coolant supply. A detailed explanation of the method used to measure thermal conductivity is provided in Sect. 2.5.4.

2 Heat Flow Modeling for Liquid-Cooled EV Battery Modules

2.1 Components of a Liquid-Cooled EV Battery Module System

The capacity of the liquid-cooled battery pack investigated in this study is approximately 35 kWh, and it is suitable for deployment in compact EV models. This battery pack is composed of multiple battery modules, TIMs, upper cooling plates, coolant, and lower cooling plates, as illustrated in Fig. 2a. Each battery module consists of battery cells, heat sinks, end plates, compression pads, and busbars. The battery modules are electrically connected in the 2P4S arrangement. In the EV, this liquid-cooled battery pack is mounted beneath the vehicle, and the battery modules are connected via a wiring harness, with 21 modules forming one battery pack. The components of the fundamental unit of the battery pack, that is, the battery module, are explained and details of each component of the battery pack are as follows:
  • EV Battery Module
    The battery module, which consists of several battery cells grouped together, is the fundamental unit of the battery pack. In addition, it helps protect the battery cells from external impacts.
  • Thermal Interface Material
    The TIM functions as a heatsink, dissipating the excess heat generated within the battery cells to maintain the battery temperature within its operational range. Positioned between the battery cells and the cooling plate, the TIM minimizes the contact thermal resistance between the cells and the cooling plate, thereby enhancing the efficiency of heat transfer.
  • Cooling Plate
    It is positioned at the bottom of the battery module to facilitate heat exchange between the cells and coolant. This component consists of the upper and lower cooling plates.
  • Coolant
    It is an insulating liquid contained within the cooling plate, and it is responsible for receiving the heat generated by the battery cells and facilitating cooling of the cells.
  • Battery Cell
    The battery cell serves as the power source in HEVs and EVs by charging and discharging electricity during vehicle operation. It utilizes lithium-ion charge carriers to convert chemical energy into electrical energy and provides the power generated in the process to the motor as an energy source.
  • Heat Sink
    Positioned between the battery cell and the compression pad, this aluminum plate facilitates efficient heat transfer from the cells to the coolant located at the bottom.
  • End Plate
    An end plate is positioned at each end of the battery module and it helps secure and immobilize the battery to prevent its movement. In addition, it helps prevent expansion of the battery due to prolonged heat exposure, thereby mitigating the risk of thermal degradation.
  • Compression Pad
    Positioned between the cells, it secures the cells in place during vehicle operation. Additionally, it helps prevent dimensional changes in cells due to prolonged usage-related heat exposure and protects them from internal impacts and vibrations. Furthermore, it acts as an insulator by isolating cell-to-cell interactions.
  • Busbar
    It connects the tabs (electrodes) at both ends of the cell, facilitating electrical connection within the battery system.
Figure 2b depicts an isometric view of the upper cooling plate of the liquid-cooled EV battery pack. Figure 2c and d present simplified views of the geometry of the liquid-cooled battery pack for the thermal and flow analysis. In Fig. 2c, the labels “Mo.N” correspond to module numbers, and they are affixed along the coolant flow path. The results of the thermal analysis indicate that the battery module located near the coolant outlet generally exhibits higher temperatures compared to the other modules. Specifically, Module 21 (Mo.21) exhibits the highest temperature among all the 21 modules constituting the battery pack. Consequently, we aim to conduct a detailed thermal and flow analysis with a focus on Module 21 to enhance its cooling performance.
Figure 3 illustrates the basic unit of the liquid-cooled EV battery pack, namely the battery module. Figure 3a–c depict the isometric, front, and rear views of the battery module, respectively. Particularly, the depiction in (a) reveals the configuration of the cooling circuit located at the bottom of Module 21. Figure 3d presents an isometric view of the assembly at the bottom of Module 21, showing the arrangement of the TIM, upper cooling plate, coolant, and lower cooling plate. Parts A and B, visible in Fig. 3d, represent sections of the cooling plates with a 2.0-mm recess, where the TIM is compressed. This aspect is crucial for examining the variations in thermal conductivity based on the compression ratio of the TIM.

2.2 Heat Dissipation Pathway of Indirect Liquid-Cooling Battery Module

As the battery cells undergo repeated charging and discharging, they generate a significant amount of heat. As mentioned in the Introduction, whether through air cooling or liquid cooling, effective heat dissipation from the battery is essential. In terms of the heat dissipation pathway in the indirect liquid-cooled battery module, the heat generated in the battery cell is initially transferred to the compression pad and heat sink located near the cell. The heat sink, depicted in Fig. 4, is L-shaped, which facilitates efficient heat dissipation in the downward direction.
This heat is then transferred to the TIM, which is located directly beneath the heat sink. Although the TIM introduces thermal resistance, it fills the gap between the battery and the cooling plate, thus reducing contact thermal resistance and facilitating effective heat dissipation. The heat absorbed by the TIM is directed toward the cooling plate and, ultimately into the coolant. Figure 5a depicts the heat dissipation path of the battery module, and Fig. 5b presents a schematic flowchart of this process.

2.3 Structure and Physical Properties of Lithium-Ion Pouch Cell

In this study, pouch-type battery cells are used to form the battery modules. Figure 6 schematically illustrates the shape of a lithium-ion battery cell, and eight of these cells are assembled in the 2P4S configuration to form a single battery module. In this configuration, two cells connected in parallel form one unit, and four such units are connected in series to form one battery module. The shape of the module used in the analysis is depicted in Fig. 3. The rated voltage of the battery cell is 3.6 V with a nominal capacity of 56 Ah. Therefore, the rated energy of one 2P4S battery module is 1.6128 kWh. The properties of the battery cell used in the CFD analysis are summarized in Table 1 of which the information is from the battery pack manufacturer. Furthermore, to calculate the heat generated by each battery cell, separate sets of properties of the cathode and anode are used in the analysis. These properties are summarized in Table 2.
Table 1
Thermal properties of Li-ion pouch battery cell
Properties
Value
Nominal capacity
56 Ah
C-rate
5C
Density
2630.7 \(\text{kg}/{\text{m}}^{3}\)
Specific heat (\({C}_{\text{P}})\)
954.0 J\(/\text{kg K}\)
Thermal conductivity
In plane
28.706 W/m k
Normal plane
1.394 W/m k
Table 2
Properties of positive and negative electrodes of Li-ion pouch battery cell
 
Positive tab
Negative tab
Density \((\rho\))
2719.0  \(\text{kg}/{\text{m}}^{3}\)
8978.0 \(\text{kg}/{\text{m}}^{3}\)
Specific heat (\({C}_{\text{P}})\)
871.0 J/kg K
381.0 J/kg K
Thermal conductivity (k)
202.4 W/m k
387.6 W/m k
Electrical conductivity \((\sigma\))
3.5 \(\times {10}^{7}\) S/m
5.8 \(\times {10}^{7}\) S/m

2.4 Battery Module Structure

The battery module targeted in this study, illustrated in Fig. 3, is designed with compression pads and end plates installed at both ends in a structure in which the heat sink-cell-compression pad is repeated in the + x direction, as depicted in Fig. 3b. This design prevents the battery cells from shaking. Additionally, in the -y direction, the battery module-TIM-cooling plate-coolant are arranged, which is the subject of analysis for indirect liquid cooling; the coolant is positioned beneath the battery cells.

2.5 Design Considering Changes in Thermal Conductivity Based on Compression Ratio

2.5.1 Thermal Interface Material Modeling

As mentioned in the previous description of battery module modeling, there exists a gap between the battery cells and the cooling plate when the battery module is assembled. This gap increases the contact thermal resistance, which hinders the heat dissipation performance of the battery. To fill this gap, a TIM is often used.
When the TIM is compressed, the contact thermal resistance decreases, leading to an increase in the thermal conductivity of the TIM (Hong & Shim, 2010) Therefore, we intend to examine the cooling efficiency of the battery module when the TIM is compressed. TIMs are divided into distinct categories, including grease, PCM, elastomer, and carbon-based material (Sarvar et al., 2006). In this study, we utilized a silicone thermal pad, which is classified as an elastomer, as the TIM to fill the assembly gap when assembling the battery module and cooling plate. The initial thickness of the TIM was set to 4 mm, denoted δ0 during assembly.
The uncompressed dimensions of TIM are 357 × 133.1 × 4 mm, and it forms a pad-shaped rectangular prism that fits the lower part of the battery module. Figure 7a depicts compression of the TIM from an initial thickness of HTIM = δ0 to δ owing to self-weight and assembly force of the battery module. Figure 7b depicts the space in which the TIM and the upper cooling plate are assembled in the battery module. In Fig. 7c, the shape of the TIM is shown and it can be observed that the shape is not uniform. This lack of uniformity can be understood by examining the shape of the upper cooling plate, which is presented in Fig. 2b; the indentations shown in this figure cause the TIM to take on an uneven form during assembly. Similarly, Fig. 7c corresponds to a part of Fig. 2b, and it shows that the TIM deforms during assembly owing to the indentations in Parts A and B. We assume that the compressed TIM volume flows toward the boundary of the battery module. Here, HTIM represents the height of the TIM when it is compressed and deformed.

2.5.2 Deformation Height of TIM

In this study, HTIM represents the height of the TIM when it is compressed. This is the most crucial parameter in our modeling process because heat transfer to the cooling plate changes as HTIM changes. Before assembly of the battery, the initial gap between the battery module and the cooling plate is set to δ0. The TIM is used to fill this gap. Subsequently, when TIM is subjected to pressure due to the weight of the battery and assembly force, its height, that is, HTIM, decreases, and any gap or void between the TIM and the cooling plate is eliminated. Consequently, HTIM decreases gradually, and the contact area with the cooling plate increases, thus effectively filling any gap. During this process, the contact thermal resistance of the TIM decreases, leading to an increase in its thermal conductivity.

2.5.3 Deformed Shape of TIM Under Compression in the Battery Module

Owing to external pressure due to the module weight and assembly force, the TIM undergoes compression deformation, as illustrated in Fig. 7c. Consequently, the TIM may extrude beyond the crevice boundaries of the battery module. We do not consider the volume that oozes outward in the x and z directions owing to the difficulty associated with the analysis of irregular shapes and because heat conduction occurs predominantly in the thickness direction. Furthermore, as discussed later, the deformation along the height direction (y-axis) was analyzed at intervals of 0.2 mm.

2.5.4 Relationship Between Compression Ratio and Thermal Conductivity of TIM

We previously described how thermal conductivity improves when the TIM (model KGP-1-7035 V by KOMOTEC) is compressed (b; Komotech Co. & Ltd, 2023a). The TIM, which is the main interest of thermal flow analysis calculation in this study, is composed of silicone, alumina, aluminum hydroxide (b; Jang et al., 2021; Komotech Co. & Ltd, 2023a; Pathumudy & Prabhu, 2021). By measuring the extent of decrease in HTIM relative to the initial TIM thickness and computing the thermal conductivity at that point, we can derive the relationship between the TIM compression ratio and thermal conductivity. In this study, we measured thermal conductivity data by varying the TIM thickness by applying pressure to the specimens according to the ASTM-D5470 method (Hanson, 2006). Figure 8 presents a schematic diagram of the test setup according to ASTM-D5470. In this experimental method, a TIM specimen is inserted between metal bars equipped with temperature sensors. The system heats up the heat source while applying compression. The temperature difference between the contact surfaces is measured when the TIM (specimen) undergoes deformation. The thermal conductivity is then measured based on the post-deformation thickness of the TIM. Table 3 presents the properties of the thermal pad before compression.
Table 3
Thermal properties of thermal interface material (TIM) (b; Komotech Co. & Ltd, 2023a)
Properties
Value
Density \((\rho\))
3240.0 \(\text{kg}/{\text{m}}^{3}\)
Specific heat (\({C}_{\text{P}})\)
995.0 J/kg K
Thermal conductivity (\(k)\)
4.45 W/m k (initial value)
Additionally, Fig. 9 depicts the variation in thermal conductivity of the TIM for different compression ratios, along with the corresponding relationship between compression ratio and thermal conductivity. This relationship may vary depending on the composition or type of thermal conductive material used as the TIM.

2.5.5 HTIM in Response to Compression of TIM

The TIM used herein is in pad form, and it is inserted between the battery module and the cooling plate during assembly. Due to the module weight and assembly force, the TIM is subjected to pressure, resulting in compression and a change in HTIM. Additionally, the compression rate is calculated by considering the thickness variation of the TIM due to the module weight and assembly force. Notably, extrusion of the TIM beyond the battery boundary, as mentioned in Sect. 2.5.3, is not considered.
Under the same assumptions as those described in Sect. 2.5.1, the cooling plate features ribs that protrude by 2.0 mm. This configuration influences the compression rate of the TIM when it undergoes compression. In Fig. 7c, HTIM, as presented, consists of the height of the flat portion of the TIM, denoted Hα, and the height of the non-flat portions due to the ribs, denoted Hβ, as in Part A and Part B, respectively. Figure 10 illustrates the compression ratio (ε), which shows the degree of compression when the TIM is subjected to assembly forces or other compressive forces in the battery module.
In Fig. 10, Hα represents the compressed height unaffected by the shapes of Part A and B, which are the protruding sections of the upper cooling plate. Hβ signifies the more heavily compressed height influenced by the shapes of Part A and B. Equation (2) describes the relationship between Hα and Hβ. The formula for calculating the crucial parameter, that is, the compression ratio (ε), is given as Eq. (3) and Eq. (4).
$${H}_{\beta }={H}_{\alpha }-2.0\left(\text{mm}\right),$$
(2)
$$\varepsilon \left({H}_{\alpha }\right)=\frac{\Delta {H}_{\text{TIM}}}{{\delta }_{0}}\times 100\left(\%\right)=\frac{{\delta }_{0}-{H}_{\alpha }}{{\delta }_{0}}\times 100\left(\%\right),$$
(3)
$$\varepsilon \left({H}_{\beta }\right)=\frac{\Delta {H}_{\text{TIM}}}{{\delta }_{0}}\times 100\left(\%\right)=\frac{{\delta }_{0}-{H}_{\beta }}{{\delta }_{0}}\times 100\left(\%\right).$$
(4)
At this point, the initial TIM thickness, \({\delta }_{0}\), is 4 mm, and the compression ratio in the section without protrusions can be calculated using Eq. (3). Additionally, the compression ratio in the section with protrusions can be calculated using Eq. (4). Table 4 summarizes the TIM thickness and compression ratios for each case, as calculated using Eqs. (3) and (5), Table 5, in conjunction with Table 4, presents the compression-ratio-based thermal conductivity of the TIM in each of the cases.
Table 4
Compression ratio and thickness according to regions in the battery module (Figs. 7c and 10)
 
\({H}_{\alpha }\) (mm)
\({H}_{\beta }\) (mm)
\(\epsilon ({H}_{\alpha })\) (%)
\(\epsilon ({H}_{\beta })\) (%)
Case 1
4.0
2.0
0
50
Case 2
3.8
1.8
5
55
Case 3
3.6
1.6
10
60
Case 4
3.4
1.4
15
65
Case 5
3.2
1.2
20
70
Case 6
3.0
1.0
25
75
Table 5
Study case (compression ratio vs thermal conductivity)
 
\({k}_{\alpha }\) (W/m k)
\({k}_{\beta }\) (W/m k)
Case 1
4.450
5.724
Case 2
4.627
5.769
Case 3
4.957
5.810
Case 4
5.150
5.849
Case 5
5.287
5.884
Case 6
5.394
5.917

3 Coolant and Heat Flow Modeling of Battery Module

The fluid governing equations used for the thermal and flow analysis of the coolant are based on the assumptions of steady state, incompressibility, and a Newtonian fluid. These equations include the continuity (5), momentum (6), and energy conservation equations (White & Xue, 2021). (7) For simulating the behavior of the coolant. Equation (5) is derived from the law of conservation of mass flow rate, which means that mass is conserved when the coolant flows through a closed curved surface inside of cooling plate. Equation (6) is a Navier–Stokes equation derived from Newton’s second law, which is a momentum equation of a coolant in the form of a nonlinear partial differential. Equation (7) is an energy equation used to simulate in consideration of heat transfer and temperature change of the coolant. Here, the \(\mu \Phi\) term means dissipation of frictional heat inside the fluid due to viscosity of the coolant.
Internal spaces of a battery pack has an air portion and a solid parts that constitutes the battery cell itself and module assembly, but convective heat flow in the air area was excluded from the analysis in this study. However, boundary conditions considering the temperature in the air space were reflected on the solid surface with the boundary condition Eq. (9). Equation (9) is convection heat transfer equation. \({T}_{\text{b}}\) and \({T}_{\infty }\) denote battery surface temperature and surrounding temperature.
$$\frac{\partial \rho }{\partial t}+\nabla \cdot \left(\rho v\right)= 0,$$
(5)
$$\rho \frac{D{v}_{i}}{Dt}=-\frac{\partial P}{\partial {x}_{i}}+\frac{\partial {\tau }_{ji}}{\partial {x}_{j}}+\rho {g}_{i},$$
(6)
$$\rho {C}_{\text{P}}\frac{DT}{Dt}=\nabla \bullet \left(k\nabla T\right)+\mu\Phi ,$$
(7)
$$Q=h\left({T}_{\text{b}}-{T}_{\infty }\right).$$
(8)
To implement the thermal behavior of a battery module composed of lithium-ion battery cells, a heat generation model is required. The commonly used formula for calculating the heat generated by a battery is Bernardi’s heat generation model which incorporates Ohm’s law, and it is expressed as Eq. (9) (Bernardi et al., 1985; Zhu et al., 2020).
$$Q={I}^{2}R+\frac{\text{d}{U}_{\text{OC}}}{\text{d}T}.$$
(9)
Here, I denotes the magnitude of battery current, regardless of whether the battery is charging/discharging, and R represents the equivalent internal resistance of the battery. UOC denotes the open-circuit voltage of the battery circuit, and \(\frac{\text{d}{U}_{\text{OC}}}{\text{d}T}\) represents the entropy coefficient indicating the entropy change of the battery cell (Zhu et al., 2020). Furthermore, the charging and discharging rates of the battery are set to 5C rate, as indicated in Table 1, to observe its behavior under extreme conditions.
In this study, the fluid domain near the cooling plate part, which represents the lowest part of the liquid cooling system in the liquid-cooled battery module, consists of the coolant, while the solid domain consists of the battery cell, compression pad, cooling plate, TIM, heat sink, and busbar. To perform the thermal and fluid analysis, we used the commercial CFD software package Ansys Fluent. The mesh was generated using Fluent Meshing, and it consisted of approximately 5,200,000 elements for one battery module. To ensure stability in the analysis process and reduce computational costs, we adopted a polyhedral mesh configuration (Sosnowski et al., 2018). To enhance the accuracy of heat transfer and fluid flow analysis, particularly in the coolant, three prism layers were stacked, as illustrated in Fig. 11.
The initial battery cell temperature and the coolant inlet temperature were assumed to be room temperature, that is, 25 °C. To closely simulate the actual flow conditions, the SST K-omega turbulence model, a standard turbulence model, was used. Additionally, Table 6 summarizes the properties of the coolant and the flow conditions. Propylene glycol, commonly used for cooling battery packs, was used as the cooling fluid (Kwon & Park, 2016.) Considering the coolant supply rate, viscosity and density of coolant (Table 6), coolant circuit dimension (Fig. 7) and flow velocity inside of coolant circuit (Fig. 2d), \(Re\) is averagely in the range of 1200–2000 under these conditions and in some regions of coolant circuit turbulence might occur.
Table 6
Coolant properties for cooling performance analysis @ 25 °C
Properties
Value
Density
1053.0 kg/m3
Specific heat (\({C}_{\text{P}})\)
3633.0 J/kg K
Viscosity
0.019 Pa s
Thermal conductivity
0.340 W/m k
Mass flow rate (inlet)
15.0 LPM
Pressure (outlet)
1.0 atm
The walls on the coolant side of coolant passage were set to be adiabatic, while the remaining walls, excluding the coolant section, were set to be convective. This is because the parts other than the coolant were exposed to air, and the coolant section was blocked from air exposure by the upper and lower cooling plates. That is to say, the parts of coolant, busbar, battery cell, compression pad, end plate, heat sink, and TIM contact are in the heat conduction mode. In addition, the wall part in contact with air is in convective heat transfer mode.

4 Thermal-Fluid Analysis of Liquid-Cooled Battery Module

We analyzed the cooling performance of a liquid-cooled battery module by considering the change in the thermal conductivity of the TIM owing to changes of its compression ratio. During CFD-based numerical analysis of heat and fluid flows, all boundary conditions were set appropriately as Table 7. The analysis was conducted by considering the compression-ratio-dependent variation in the thermal conductivity of the TIM. After completion of the CFD calculations, the maximum, minimum, and average temperatures in each case were examined, and the differences in the cooling performance of the liquid-cooled battery module were analyzed.
Table 7
Heat flow conditions at the walls of battery module
Interfaces and inside of domain
Heat transfer mode
Coolant
Conductive
Busbar, cell
Compression pad
End plate, heat sink, TIM
Wall adjacent to air
Convective
\(h=5\) W/m2 K
\({T}_{\infty }=25^\circ C\)
Figures 12 and 13 depict the streamlines and pressures of the coolant flowing in the coolant jacket at the bottom of the battery module No.21, respectively. Figure 14 depicts the overall temperatures in each case of battery module No.21 by considering the TIM-compression-dependent cooling performance. Figure 15 shows the probe plane at the center of the yz plane of the battery module No.21. The yz plane was selected made to observe the temperature distributions of all cells (Fig. 16). Moreover, the central region in the vertical direction of the cells was selected to easily observe the cooling effect resulting from coolant flow because this area experiences active coolant flow.
Figure 16 depicts the temperatures of the cells, compression pads, end plates, and heat sinks based on the probe plane selected in Fig. 15. The cooling effect is the most pronounced at the highest compression ratio of HTIM = 3.0 mm, where the compression ratio is 25% in relatively thick regions without raggedness and 75% in the thin regions with raggedness. In all cases presented in Figs. 14 and 16, the asymmetric distribution of cell temperatures can be attributed to the streamline distributions of coolant as shown in Fig. 12, where fluid flow is more vigorous near the coolant inlet side and relatively less effective in areas further from the coolant supply.
Figure 17 illustrates the maximum, minimum, and average temperatures of the liquid-cooled battery cells in each case under TIM compression. In Case 1 (HTIM = 4.0 mm), the maximum battery temperature was 63.35 °C. Comparatively, a cooling effect of approximately 3.20 °C reduction was observed when the TIM was compressed between case 1 (63.35 °C @HTIM = 4.0 mm) and case 6(60.15 °C @HTIM = 3.2 mm). The minimum temperature was observed in Cell 01, which was the closest to the coolant, while the average temperatures in Cases 1 and 6 were 54.725 °C and 51.862 °C, respectively, indicating a reduction of approximately 2.863 °C in Case 6. Furthermore, the pressure drop was consistently calculated as 1.53 kPa in all cases, suggesting the existence of a uniform pressure differential because the experimental setup did not alter the fluid flow structure.
As the compression ratio of the TIM material increases, the thermal conductivity \(({k}_{\alpha }\),\({k}_{\beta })\) increases and the rate of change \(\left(\frac{{\Delta k}_{\alpha }}{{\Delta H}_{\alpha }},\frac{\Delta {k}_{\beta }}{\Delta {H}_{\beta }}\right)\) also increases (Table 4). What is more, the heat conduction length decreases as the compression ratio increases, and this \(\left(\frac{\partial }{\partial {x}_{i}}\left({k}_{i}\frac{\partial {T}_{j}}{\partial {x}_{i}}\right)\right)\) accelerates the heat conduction from cell to coolant. The size variations of the cooling plate located under the bottom of each battery cell vary the heat flow rate. Heat is absorbed to the coolant with some distribution and it does not show a proportional decrease depending on the size change in compressibility of TIM. Therefore, as the compression ratio of the TIM increases, the thickness of TIM decreases and the cooling efficiency relatively increases.

5 Conclusion

In this study, thermal cooling analysis of a liquid-cooled battery module was conducted by considering changes in the thermal conductivity of the TIM depending on its compression ratio due to height variations resulting from assembly of the EV battery module. In addition, we explored the variation in the thermal conductivity of the battery module with the compression ratio of the TIM.
1.
The experimental results confirmed that the thermal conductivity of the TIM increased with compression, which reduced its thickness. This, in turn, enhanced the heat dissipation performance of the battery module.
 
2.
By appropriately leveraging the relationship between the compression ratio of the TIM and its thermal conductivity, we can propose a design that enhances the effective heat dissipation performance of the battery module.
 
In future studies, we aim to perform numerical analysis of the battery module through structural analysis coupled with pressure profiles for verifying the cooling analysis results more precisely. This approach will enhance the accuracy of the cooling analysis by considering the coupled structural and thermal effects.

Acknowledgements

This work was supported by “BK21 Four Program (5199990814084),” and the research project titled “Development of High Thermal Conductivity Thermal Interface Material for Enhancing Cooling in Electric Vehicle Battery Pack Systems (20022377)” funded by KEIT.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Unsere Produktempfehlungen

ATZelectronics worldwide

ATZlectronics worldwide is up-to-speed on new trends and developments in automotive electronics on a scientific level with a high depth of information. 

Order your 30-days-trial for free and without any commitment.

ATZelektronik

Die Fachzeitschrift ATZelektronik bietet für Entwickler und Entscheider in der Automobil- und Zulieferindustrie qualitativ hochwertige und fundierte Informationen aus dem gesamten Spektrum der Pkw- und Nutzfahrzeug-Elektronik. 

Lassen Sie sich jetzt unverbindlich 2 kostenlose Ausgabe zusenden.

Literatur
Zurück zum Zitat Bulinski, Z., Pawlak, S., Krysinski, T., Adamczyk, W., & Biatecki, R. (2019). Application of the ASTM D5470 standard test method for thermal conductivity measurements of high thermal conductive materials. Journal of Achievements in Materials and Manufacturing Engineering, 95(2), 57–63. https://doi.org/10.5604/01.3001.0013.7915CrossRef Bulinski, Z., Pawlak, S., Krysinski, T., Adamczyk, W., & Biatecki, R. (2019). Application of the ASTM D5470 standard test method for thermal conductivity measurements of high thermal conductive materials. Journal of Achievements in Materials and Manufacturing Engineering, 95(2), 57–63. https://​doi.​org/​10.​5604/​01.​3001.​0013.​7915CrossRef
Zurück zum Zitat Hong, J. H., & Shim, S. E. (2010). Trends in development of thermally conductive polymer composites. Applied Chemistry for Engineering, 21(2), 115–128. Hong, J. H., & Shim, S. E. (2010). Trends in development of thermally conductive polymer composites. Applied Chemistry for Engineering, 21(2), 115–128.
Zurück zum Zitat Komotech Co., Ltd. (2023a). Sample technical data sheet—KGP-1-7035 (V). Komotech Co., Ltd. (2023a). Sample technical data sheet—KGP-1-7035 (V).
Zurück zum Zitat Komotech Co., Ltd. (2023b). Silicone TIM educational materials. Komotech Co., Ltd. (2023b). Silicone TIM educational materials.
Zurück zum Zitat Thakur, A. K., Prabakaran, R., Elkadeem, M. R., Sharshir, S. W., Arici, M., Wang, C., Zhao, W., Hwang, J. Y., & Saidur, R. (2020). A state of art review and future viewpoint on advance cooling techniques for lithium-ion battery system of electric vehicles. Journal of Energy Storage, 32, 101771. https://doi.org/10.1016/j.est.2020.101771CrossRef Thakur, A. K., Prabakaran, R., Elkadeem, M. R., Sharshir, S. W., Arici, M., Wang, C., Zhao, W., Hwang, J. Y., & Saidur, R. (2020). A state of art review and future viewpoint on advance cooling techniques for lithium-ion battery system of electric vehicles. Journal of Energy Storage, 32, 101771. https://​doi.​org/​10.​1016/​j.​est.​2020.​101771CrossRef
Zurück zum Zitat Viswanath, R., Wakharkar, V., Watwe, A., & Lebonheur, V. (2000). Thermal performance challenges from silicon to systems. Intel Technology Journal, Q3. Viswanath, R., Wakharkar, V., Watwe, A., & Lebonheur, V. (2000). Thermal performance challenges from silicon to systems. Intel Technology Journal, Q3.
Zurück zum Zitat White, F. M., & Xue, H. (2021). Fluid mechanics (9th ed.). McGrawHill. White, F. M., & Xue, H. (2021). Fluid mechanics (9th ed.). McGrawHill.
Metadaten
Titel
Study of Cooling Performance of Liquid-Cooled EV Battery Module According to the TIM Compression Ratio
verfasst von
Jahun Gu
Heung-Kyu Kim
Siyoul Jang
Publikationsdatum
02.12.2024
Verlag
The Korean Society of Automotive Engineers
Erschienen in
International Journal of Automotive Technology
Print ISSN: 1229-9138
Elektronische ISSN: 1976-3832
DOI
https://doi.org/10.1007/s12239-024-00167-8