Next Article in Journal
Influence of Intake Port Structure on the Performance of a Spark-Ignited Natural Gas Engine
Previous Article in Journal
Prediction of Building Electricity Consumption Based on Joinpoint−Multiple Linear Regression
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of the Parameters Affecting a Heat Pipe Thermal Management System for Lithium-Ion Batteries

by
Kittinan Boonma
1,
Napol Patimaporntap
1,
Hussein Mbulu
1,
Piyatida Trinuruk
1,
Kitchanon Ruangjirakit
1,
Yossapong Laoonual
1 and
Somchai Wongwises
1,2,*
1
Fluid Mechanics, Thermal Engineering and Multiphase Flow Research Lab (FUTURE), Department of Mechanical Engineering, Faculty of Engineering, King Mongkut’s University of Technology Thonburi (KMUTT), Bangmod, Bangkok 10140, Thailand
2
National Science and Technology Development Agency (NSTDA), Pathum Thani 12120, Thailand
*
Author to whom correspondence should be addressed.
Submission received: 7 August 2022 / Revised: 5 October 2022 / Accepted: 30 October 2022 / Published: 15 November 2022
(This article belongs to the Section E: Electric Vehicles)

Abstract

:
The thermal management system of batteries plays a significant role in the operation of electric vehicles (EVs). The purpose of this study is to survey various parameters enhancing the performance of a heat pipe-based battery thermal management system (HP-BTMS) for cooling the lithium-ion batteries (LIBs), including the ambient temperature, coolant temperature, coolant flow rate, heat generation rate, start-up time, inclination angle of the heat pipe, and length of the condenser/evaporator section. This review provides knowledge on the HP-BTMS that can guarantee achievement of the optimum performance of an EV LIB at a high charge/discharge rate.

Graphical Abstract

1. Introduction

Air pollution is a result of the heavy traffic of internal combustion engine vehicles. Electric vehicles (EVs) do not emit air pollution during operation and are one of the alternative modes of transportation that help reduce air pollution.
In EVs, the battery is the main power source. Lithium-ion is one of the key components in EV batteries. The advantages of these batteries are that they have high energy density, high power density, lighter weight, better recyclability, and a long life-cycle compared to other types of batteries [1,2,3,4,5]. Furthermore, they have no effect on memory [1,2,4,5]. However, the main problem with lithium-ion batteries (LIBs) is the temperature increase during the charge/discharge process. Therefore, an appropriate battery thermal management system (BTMS) must be provided to maintain a uniform temperature distribution and optimal temperature range in an EV battery.
Normally, the maximum temperature (Tmax) and temperature difference (ΔT) on the battery surface increase due to the heat generation in a LIB. The heat generated during battery operation originates from reversible and irreversible heat sources. The former includes entropy changes during electrochemical reactions. The latter includes internal resistance, concentration, activation, and ohmic polarization [6].
The accumulation of heat accelerates aging and shortens the battery’s life span, leading to a lower charge/discharge efficiency and reduced battery performance. Extremely high temperatures can lead to thermal runaway [7]. Inversely, low temperatures can result in a higher internal resistance, which also reduces battery performance [8]. Therefore, a battery should be operated under an optimal temperature range. The optimum maximum temperature (Tmax) for LIBs is 25–40 °C [9,10,11], and the temperature difference (ΔT) of the batteries must be maintained at below 5 °C [9,10,11,12].
A well-designed BTMS is necessary to ensure that batteries are working in suitable operating conditions [13]. In the design of a BTMS, safety, ease of maintenance, size, weight, and potential applications have to be taken into account [14]. BTMSs have been categorized into four main groups: air cooling, liquid cooling, phase change material cooling, and heat pipe cooling, as shown in Figure 1.
Air cooling BTMSs are used mainly because they have uncomplicated structures and are inexpensive and lightweight. However, despite their advantages, air cooling BTMSs fail severely in stressful conditions such as in high ambient temperatures and high charge/discharge rates, leading to uneven temperature distributions and high temperature differences [15,16,17,18]. Researchers have tried to improve the performance of air cooling BTMSs by considering structural designs, battery layout optimization, and control strategies, including the use of forced convection systems [17,19,20].
Due to their high heat transfer coefficients, liquid cooling BTMSs are more efficient for LIBs [21]. They can be divided into two types—direct and indirect contacts [22]—corresponding to whether the liquid is in direct contact with the battery. However, if an EV is involved in an accident, the possibility of the liquid leaking into the battery is high [23], resulting in the short circuiting of the battery pack.
Phase change materials (PCMs) also can be used in BTMSs and work well for maintaining temperature uniformity in a battery module/pack [24]. However, a drawback is that PCMs have low thermal conductivity [25]. Researchers have proposed various methods for enhancing PCMS’s heat transfer capacities, such as adding fins [26], metal mesh [27], graphite [28], or metal foam [24]. When a PCM absorbs high heat from a battery surface, the PCM melts and its volume expands; therefore, a leak-proof design is needed [21].
The application of a heat pipe for cooling is an emergent passive cooling technique. Heat pipes are often called superconductors of heat due to their heat transfer rates and minimal heat loss. A heat pipe consists of a condenser section, an adiabatic section, and an evaporator section, as shown in Figure 2. Liquid in the evaporator section absorbs heat and vaporizes. The vapor then flows to the condenser section, where it condenses. The condensed liquid then returns to the evaporator section thought either capillary force, centrifugal force, or gravitational force, and it re-vaporizes to repeat the cycle [29]. Heat pipes have been widely used in cooling applications where their evaporator sections are attached to heat sources.
A heat pipe-based battery thermal management system (HP-BTMS) is one of the technologies that meets almost all the required conditions for a BTMS, ensuring that the battery operates at an optimal temperature. The advantages of such a system are its high heat transfer rate, low weight, compact size, low cost, long life, and high capacity for moving heat from the inside of the battery pack to the outside [30]. Moreover, the heat pipe structure in the condenser section can be coupled with forced air [31] or a water spray [32] to ensure maximum performance. Some investigators have coupled an aluminum fin to the condenser to enhance the heat transfer rate [33]. However, with an HP-BTMS, a problem still exists with keeping the battery temperature within the optimum range, especially when the charge/discharge rate is high. Some researchers have used plate flat heat pipes for BTMSs to obtain temperature uniformity of the battery pack [34,35,36].
To solve this problem, a PCM is used in conjunction with a heat pipe to reduce the temperature difference and maximum temperature [37]. In addition, the thermal behaviour of a constructal theory-based heat pipe was studied by Boonma et al. [38]. The results of their study provided information on adaptable thermal management systems for the development of compact battery packs. By increasing the condensing areas of heat pipes, the size and number of heat pipes needed for cooling could be reduced.
Table 1 summarizes some related research on heat pipes, categorized as air cooling and liquid cooling depending on the coolant used in the condenser end. These previous studies show that a large number of researchers have investigated the effects of the variables related to the thermal performance of HP-BTMSs. Nevertheless, no review paper has put all of that information together. Thus, the objective of this paper is to report the effects of the relevant parameters (heat generation rate, ambient temperature, coolant temperature, coolant mass flow rate/air velocity, condenser/evaporator section length, inclination angle, and start-up time) on the performance of HP-BTMSs.

2. Parameters Affecting the Performance of a HP-BTMS

2.1. Effects of the Heat Generation Rate

Heat generation from batteries is unavoidable. Therefore, the BTMS design is essential to keep the battery temperature within the proper values.
The heat generated is related to the high current rate (C rate) or the charge/discharge rate. The total heat generated ( Q · total) is calculated as [6]:
Q · total = Q · rev + Q · irr = I · T   dU OC dT + I 2 · R ,
where I, T, dU OC dT , and R are the current, temperature, entropy coefficient, and internal resistance of the cell, respectively.
Hu et al. [6] measured the heat generation of LIBs during operation to analyze the BTMS. Their tests on the effects of C rates on heat generation were performed at Tamb = 25 °C. The C rate values considered for charging/discharging were C/3, 0.5 C, 1 C, 2 C, 3 C, and 4 C, as shown in Figure 3.
Zeng et al. [43] applied a hybrid micro heat pipe array (MHPA) to the BTMS of cylindrical batteries. Compared to a system without the MHPA, for a 1 C discharging process, the Tmax was reduced by 34.11% and the ΔT decreased from 3.66 to 0.66 °C. For the 3 C discharging process, the Tmax and ΔT decreased to 41.03 and 2.16 °C, respectively.
Some researchers have used a battery surrogate to represent a real battery. The heat input power (Q) is analogous to the heat generated by an actual LIB during acceleration and high charging or discharging conditions. The heat generated might range from 10 to 40 W under normal working conditions. In contrast, in severe working conditions, it might range from 50 to 70 W. Normally, when the heat generated, Tmax, and ΔT increase, the extensive heat produced can affect the LIB severely to the point where it is beyond recovery. Therefore, the heat transfer rate (QHP) of a heat pipe depends on the input power (Q), which can be calculated as [44]:
Q HP = Q ( T sat , e T sat , c ) kA k L ce ,
where T sat , e and T sat , c are the evaporator and condenser saturation temperatures, respectively. The variables k and A k are the heat pipe’s thermal conductivity and the cross-section area of the heat pipe, respectively, and L ce is the length between the points where the T sat , e and T sat , c are measured. Equation (2) shows that when the input power ( Q ) increases, the Q HP increases and, therefore, the battery surface temperature ( T H ) increases, which leads to an increase in Tmax and ΔT.
Rao et al. [45] conducted an experiment to investigate the temperature distribution of a battery when the input power varied from 10 to 60 W. The results showed that the Tmax could be kept below 50 °C when the heat generation rate was below 50 W and the ΔT was less than 5 °C when the heat generation rate was 30 W or less.
Wang et al. [46] investigated the effectiveness of heat pipes by constructing a simulated heat source in the range of 2.5 to 40 W/cell. They found that the heat pipe could regulate the temperature of a battery to below 40 °C when the heat generated by the battery was less than 10 W/cell. Additionally, the analysis showed that even when the battery generated heat in the range of 20–50 W/cell, which was considered thermal abuse, the heat pipe could keep the temperature of the battery below 70 °C, preventing thermal runaway.
Mbulu et al. [5] reported that their HP-BTMS with sandwiched L- and I-shaped heat pipes was able to keep a battery’s temperature under 55 °C, even when the heat produced from the battery was 60 W. When the heat input increased to 30, 40, 50, and 60 W, the Tmax on the battery surface was 43.82, 46.59, 50.96, and 54.38 °C, respectively, as shown in Figure 4.
According to the information gathered above, the heat generation rate of a battery is the result of the current rate or heat input, which affects the Tmax, ΔT, and battery performance. This could lead to an internal short circuit and thermal runaway [47]. Therefore, it is important to consider heat generation and the related parameters when designing an HP-BTMS. Table 2 summarizes investigations of the heat pipe cooling system at different heat generation rates.

2.2. Effects of the Ambient Temperature

The ambient temperature is a factor that contributes to either a decrease or increase of the temperature surrounding the LIB’s position. The ambient temperature (Tamb) contributes to an increase in the LIB’s heat generation rate if the Tamb is above 15 °C. However, when the Tamb is below 15 °C, the chemical reaction will be impeded; thus, the output power generated will deteriorate [39]. The increased heat generation rate results in increases in the Tmax and ΔT.
Behi et al. [52] carried out an experimental study to analyze the impact of the Tamb of the air on a battery module’s temperature distribution, as shown in Figure 5. The researchers investigated three ambient temperatures: 10, 26, and 35 °C. The experiment clearly proved that the Tamb of the air has a critical effect on the temperature distribution of a battery. When the Tamb is low, the temperature distribution on the battery surface decreases.
Alihosseini and Shafaee [53] analyzed a battery’s thermal performance based on winter and summer temperatures of 18, 28, and 33 °C, as shown in Figure 6. In this experiment, the Tamb was adjusted and controlled in a test chamber. Li et al. [54] studied the effects of the Tamb on the Tmax and ΔT. They found that a high Tamb will cause the Tmax to rise above the optimal temperature and cause the HP-BTMS’s thermal performance to decrease. Meanwhile, the ΔT of the battery decreases when the Tamb increases. This result implies that a low Tamb will elevate the ΔT.
After reviewing the literature on the ambient temperature’s effects on heat pipe performance, we concluded that an HP-BTMS should be operated in a suitable environment to achieve optimal thermal performance for the LIB.

2.3. Effects of the Coolant Temperature

To keep the temperature of the battery within the optimal range, the coolant Tmax must be controlled. A low coolant temperature results in a decrease in the maximum temperature. Too low of a coolant temperature results in a high temperature difference, which leads to a reduction in battery capacity. Therefore, it is necessary to use the optimum coolant temperature.
Ye et al. [10] studied the effects of coolant temperature at a fixed flow rate. They chose coolant temperature values of 15, 20, and 25 °C and found that with a battery heating load of 100 W, the temperature could be kept below 40 °C if a coolant temperature of 15 °C was used. Additionally, they also discovered that although the use of a low coolant temperature increased thermal performance, if the temperature was too low, it led to a significant increase in ΔT.
Liang et al. [39] analyzed the effects of coolant temperature on the Tmax and ΔT of a battery at ambient temperatures of 25 and 35 °C. A lower coolant temperature lowered the Tmax. In this study, the Tmax of the battery was found to be above 40 °C when the coolant temperature was approximately 20 °C. On the other hand, if the coolant temperature was reduced too much, the temperature was more unevenly distributed, as shown in Figure 7.
Li et al. [54] investigated the effects of coolant temperature on battery cooling capacity under an ambient temperature of 40 °C. The study was conducted using batteries with a discharge rate of 4 C. The results showed that the battery temperature could be kept below 40 °C when the coolant temperature was 20 °C, as shown in Figure 8.
Based on the above studies on the effects of coolant temperature, we concluded that reducing coolant temperature is one method for improving the performance of an HP-BTMS. However, the disadvantage is that if the coolant temperature is reduced too much, the temperature of the battery surface will be unevenly distributed. Therefore, the coolant temperature should be within the appropriate range.

2.4. Effects of the Coolant Flow Rate

The mass flow rates of the coolant in the condenser can affect the Tmax and ΔT. When the coolant mass flow rates increase, the Tmax and ΔT decrease. Nevertheless, when the mass flow rates increase beyond the optimal value, it has no significant impact on either the Tmax or ΔT.
Researchers have also found that pressure loss in the cooling channel significantly increases as the coolant mass flow rate increases, which signifies that only increasing the flow rate of the coolant is not applicable for enhancing the performance of an HP-BTMS [39,40,55].
Ye et al. [10] investigated the mass flow rate’s effects on the Tmax and ΔT. They found that the Tmax and ΔT decreased greatly when the flow rate increased from 1 to 2 LPM. However, when the flow rate increased from 2 to 3 LPM, the Tmax and ΔT decreased only slightly.
Wu et al. [56] investigated the effects of air velocities between 0 and 3 ms−1 with a battery discharge rate of 5 C, as shown in Figure 9. They found that the Tmax decreased with the rising air velocity. However, when the air velocity increased further, the Tmax decreased at a lower rate due to the limitation of the phase transition process.
Wan [57] found that an increase in the coolant flow rate from 0.5 to 2 LPM resulted in a decrease in the Tmax of approximately 8.9 °C (16.9%). In addition, when the coolant flow rate increased from 2 to 8 LPM, the Tmax decreased by approximately 4.07 °C (approximately 9.3%). It should be noted that when the coolant flow rate increased from 0.5 to 8 LPM, the pressure drop (ΔP) increased remarkably from 1.35 Pa to 77.76 Pa, as shown in Figure 10.
Gan et al. [41] confirmed the effects of the coolant’s mass flow rates on drops in pressure. An increase in the coolant flow rate resulted in an increased pressure drop in the coolant circulation loop system. The authors found that when the coolant mass flow rate increased from 0.5 to 2 LPM, the pressure drop increased from 872.5 to 7995.3 Pa, as shown in Figure 11. The authors suggested the optimum flow rate in their study was 0.5 LPM because this flow rate provided a relatively low pressure drop.
Yao et al. [48] investigated the performance of heat pipes at air velocities ranging from 0 to 3.5 ms−1 with a heat generation rate of 40 W. They found that at an air velocity of 3.5 ms−1, the battery temperature could be maintained at 40 °C, and at an air velocity of 0 ms−1 (natural cooling), the battery temperature was 75 °C. They also found that at air velocities of between 2.5 and 3.5 ms−1, there were decreases in temperature. However, the rate of the temperature drop was noticeably reduced. In terms of the temperature difference, the mean value for all air velocities applied was 2 °C. Therefore, they concluded that the air velocity had no pronounced effect on the temperature distribution.
Wan [58] investigated the effects of coolant flow rates of between 1 and 3 LPM on the cooling performance of an HP-BTMS. Both the Tmax and ΔT were reduced as the coolant flow rate decreased. Nevertheless, for coolant flow rates ranging from 2 to 3 LPM, the Tmax and ΔT decreased to lesser extents.
Concerning the coolant flow rate’s effects, many researchers have found that increasing the coolant flow rate is a way to increase the thermal performance of an HP-BTMS. However, using a very high coolant flow rate is not necessary because at a very high coolant flow rate, the battery temperature changes very little. In addition, to achieve a higher coolant flow rate, more energy is needed to deliver the fluid. Therefore, an appropriate coolant flow rate is needed to increase the thermal performance of an HP-BTMS.
Furthermore, researchers have studied the effects of the mass flow rate on the heat pipe’s thermal resistance and found that an increase in the mass flow rate led to a decrease in the heat pipe’s thermal resistance [40]. The main cause of this was the increase in heat transfer dissipation caused by the increased water motion [40].

2.5. Effects of the Start-Up Time

The start-up time is the time that a BTMS takes to start after the battery is initiated. It is not worth allowing the BTMS to cool the battery immediately after it starts up because the heat has not yet accumulated in the battery. Synchronized cooling, in which the BTMS works immediately, should not be used due to the overcooling of the battery [39]. This overcooling decreases the Li-ion diffusivity and the rate of the chemical reaction under a low Tamb, which results in a higher over-potential and a lower charge (discharge) capacity [39].
To start a BTMS, one should allow the battery to run and heat up for a while and then start the cooling process (unsynchronized cooling). In short, this kind of cooling prevents the battery from overcooling because the delay allows the battery to reach a higher temperature [39].
Liang et al. [39] found that cooling strategies with a lag time between the start-up of the HP-BTMS and the battery could enhance performance. However, adjusting the lag time was difficult because it depended on the heat generation rate. The researchers suggested the use of a temperature difference between the inlet coolant temperature and the Tmax (Tbf) because these temperatures were determinable. The Tbf was applied to control the start-up time. The Tbf determined the start-up of the BTMS, but it also depended on the Tab. At Tab = 15 °C, the BTMS could be started at Tbf = 15 °C. However, at Tab = 25 °C, the BTMS could be started at Tbf = 10 °C.
Therefore, controlling the start-up time could help conserve power to turn on either the fan or the pump. Wan [59] studied the effects of the BTMS’s start-up time and obtained the same results as Liang et al. [39].

2.6. Effects of the Inclination Angle of the Heat Pipe

The angle between the horizontal plane and the vertical axis (inclination angle) has a significant effect on the performance of a heat pipe [60].
Ye et al. [10] studied the effects of the orientation of the condensation section in three configurations: (1) the condenser on top and the evaporator on the bottom, (2) horizontal, and (3) the condenser on the bottom and the evaporator on top. Their results showed that the optimal performance was obtained when the condenser was on top.
Behi et al. [61] examined the inclination angle’s effects on the performance of heat pipe-based air cooling for a lithium-titanate-oxide (LTO) cell by measuring the thermal resistance. They found that the temperature on the backside of the LTO cell and the heat pipe’s thermal resistance decreased as the inclination angle increased (0° to 90°). In addition, it was obvious that the temperature difference between the condenser and evaporator sections decreased (from 5.26 to 5.04 °C) as the inclination angle increased (0° to 90°). The condensed water could return faster to the evaporator due to gravity.
Yao et al. [48] analyzed the effect of a heat pipe’s tilt angle. They conducted the experiment under a heat generation of 10 W with natural air convection for tilt angles of −75°, −30°, 0°, 15°, 45°, and 90°. At a tilt angle of 90°, the Tmax and ΔT were 60 and 3 °C, respectively. When the tilt angle was −75°, the Tmax and ΔT increased to 88 and 16 °C, respectively. The authors showed that the effective thermal conductivity decreased with decreases in the tilt angles; for example, the effective thermal conductivity was 5820 W·m−1·K−1 at 90° and 295 W·m−1·K−1 at −75° [48].
In summary, the performance obtained from an HP-BTMS depends on the orientation angle of the unit. An increase in the inclination angle caused the condensed working fluid in the heat pipe to return to the evaporation section faster, which resulted in less flow resistance [61].

2.7. Effects of the Condenser and Evaporator Section Lengths

The heat pipe’s total thermal resistance is the sum of the thermal resistance of the evaporator ( R e ) , adiabatic ( R a ) , and condenser ( R c ) , which can be expressed as
Rhp = Re + Ra + Rc,
For the evaporator section, the thermal resistance is calculated as
R e = 1 h e A e δ k l A e ,
where h e , δ , A e , and k l are the evaporation heat transfer coefficient, liquid film thickness, heating area, and thermal conductivity of the liquid, respectively [44]. The equation shows that the evaporator’s thermal resistance is inversely proportional to its heating area ( A e ), which is related to the heating length.
Nasir et al. [40] studied the effects of the condenser and evaporator sections’ lengths on the performance of a heat pipe. Condenser section lengths of 100 and 150 mm were studied. A slight decrease and an uneven distribution of the battery’s surface temperature were found. The authors also found that the surface temperature decreased in the range of 0.949–2.96%. When the condenser section’s length increased, the amount of condensed liquid would increase. This condensed liquid would be transported to the evaporator section and turned into vapors in the evaporator section. This implied more heat would be absorbed from the battery’s surface; therefore, the battery’s surface temperature decreased [40].
Nevertheless, the decrease in the Tmax, (1.76%) was small. Moreover, with the increase in condenser length, the ΔT on the battery’s surface increased moderately, from 0.142 to 5.34%. The researchers also observed a large drop in the total thermal resistance, which decreased by 5.12%. Moreover, the Nusselt number increased from 224 to 400 and from 205 to 319 when the battery heat generation rates were 10 and 15 W, respectively. The average increase was 45.1% when the condenser length increased by 50 mm, which implied that the condenser length had a vital effect on convective heat transfer to the condenser’s external side.
Gan et al. [41] investigated the condenser length’s effects on a BTMS’s cooling performance. When the condenser length was increased from 6 to 42 mm, the Tmax and ΔT decreased by 6.97 and 0.79 °C, respectively, as shown in Figure 12. Although the heat transfer area increased with an increased condenser length, an increase in length might cause the HP-BTMS’s weight to increase. Therefore, selecting the appropriate condenser length enhances the BTMS’s cooling performance.
We found that increasing either the condenser or evaporator length enhanced the thermal performance of a heat pipe. However, the length of either the evaporator or condenser section should be carefully investigated to avoid an increase in the HP-BTMS’s weight and size.

3. Conclusions

In this paper, the effects of ambient temperature, coolant temperature, coolant flow rate, heat generation rate, start-up time, inclination angle of a heat pipe, and cooling/heating length on the performance of HP-BTMSs are reviewed. The results are useful for practical applications in the design of BTMSs.

Author Contributions

Writing—review and editing, K.B. and N.P.; writing—original draft preparation, H.M. Conceptualization, P.T., K.R. and Y.L.; review—editing and project administration, S.W. All authors have read and agreed to the published version of the manuscript.

Funding

The NSTDA Research Chair Grant, and the TSRI Fundamental Fund 2023.

Data Availability Statement

Not applicable.

Acknowledgments

The authors acknowledge NSTDA for the Research Chair Grant, and Thailand Science Research and Innovation (TSRI) for Fundamental Fund 2023.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gou, J.; Liu, W. Feasibility study on a novel 3D vapor chamber used for Li-ion battery thermal management system of electric vehicle. Appl. Therm. Eng. 2019, 152, 362–369. [Google Scholar] [CrossRef]
  2. Guo, S.; Xiong, R.; Wang, K.; Sun, F. A novel echelon internal heating strategy of cold batteries for all-climate electric vehicles application. Appl. Energy 2018, 219, 256–263. [Google Scholar] [CrossRef]
  3. Rajan, A.; Vijayaraghavan, V.; Ooi, M.P.-L.; Garg, A.; Kuang, Y.C. A simulation-based probabilistic framework for lithium-ion battery modelling. Measurement 2018, 115, 87–94. [Google Scholar] [CrossRef]
  4. Zhang, W.; Qiu, J.; Yin, X.; Wang, D. A novel heat pipe assisted separation type battery thermal management system based on phase change material. Appl. Therm. Eng. 2020, 165, 114571. [Google Scholar] [CrossRef]
  5. Mbulu, H.; Laoonual, Y.; Wongwises, S. Experimental study on the thermal performance of a battery thermal management system using heat pipes. Case Stud. Therm. Eng. 2021, 26, 101029. [Google Scholar] [CrossRef]
  6. Hu, Y.; Choe, S.; Garrick, T. Measurement of heat generation rate and heat sources of pouch type Li-ion cells. Appl. Therm. Eng. 2021, 189, 116709. [Google Scholar] [CrossRef]
  7. Xu, J.; Lan, C.; Qiao, Y.; Ma, Y. Prevent thermal runaway of lithium-ion batteries with minichannel cooling. Appl. Therm. Eng. 2017, 110, 883–890. [Google Scholar] [CrossRef] [Green Version]
  8. Jaguemont, J.; Boulon, L.; Dubé, Y. A comprehensive review of lithium-ion batteries used in hybrid and electric vehicles at cold temperatures. Appl. Energy 2016, 164, 99–114. [Google Scholar] [CrossRef]
  9. Ye, Y.; Saw, L.H.; Shi, Y.; Tay, A.A. Numerical analyses on optimizing a heat pipe thermal management system for lithium-ion batteries during fast charging. Appl. Therm. Eng. 2015, 86, 281–291. [Google Scholar] [CrossRef]
  10. Ye, Y.; Shi, Y.; Saw, L.H.; Tay, A.A. Performance assessment and optimization of a heat pipe thermal management system for fast charging lithium ion battery packs. Int. J. Heat Mass Transf. 2016, 92, 893–903. [Google Scholar] [CrossRef]
  11. Pesaran, A.A. Battery thermal models for hybrid vehicle simulations. J. Power Sources 2002, 110, 377–382. [Google Scholar] [CrossRef]
  12. Zhao, R.; Gu, J.; Liu, J. An experimental study of heat pipe thermal management system with wet cooling method for lithium ion batteries. J. Power Sources 2015, 273, 1089–1097. [Google Scholar] [CrossRef]
  13. Hwang, H.-Y.; Chen, Y.-S.; Chen, J.-S. Optimizing the heat dissipation of an electric vehicle battery pack. Adv. Mech. Eng. 2015, 7, 204131. [Google Scholar] [CrossRef] [Green Version]
  14. Chombo, P.V.; Laoonual, Y. A review of safety strategies of a Li-ion battery. J. Power Sources 2020, 478, 228649. [Google Scholar] [CrossRef]
  15. Roe, C.; Feng, X.; White, G.; Li, R.; Wang, H.; Rui, X.; Li, C.; Zhang, F.; Null, V.; Parkes, M.; et al. Immersion cooling for lithium-ion batteries-A review. J. Power Sources 2022, 525, 231094. [Google Scholar]
  16. Dan, D.; Yao, C.; Zhang, Y.; Zhang, H.; Zeng, Z.; Xu, X. Dynamic thermal behavior of micro heat pipe array-air cooling battery thermal management system based on thermal network model. Appl. Therm. Eng. 2019, 162, 114183. [Google Scholar] [CrossRef]
  17. Yang, W.; Zhou, F.; Zhou, H.; Wang, Q.; Kong, J. Thermal performance of cylindrical lithium-ion battery thermal management system integrated with mini-channel liquid cooling and air cooling. Appl. Therm. Eng. 2020, 175, 115331. [Google Scholar]
  18. Mahamud, R.; Park, C. Reciprocating air flow for Li-ion battery thermal management to improve temperature uniformity. J. Power Sources 2011, 196, 5685–5696. [Google Scholar] [CrossRef]
  19. Park, H. A design of air flow configuration for cooling lithium ion battery in hybrid electric vehicles. J. Power Sources 2013, 239, 30–36. [Google Scholar]
  20. Yang, W.; Zhou, F.; Zhou, H.; Liu, Y. Thermal performance of axial air cooling system with bionic surface structure for cylindrical lithium-ion battery module. Int. J. Heat Mass Transf. 2020, 161, 120307. [Google Scholar]
  21. Buidin, T.I.C.; Mariasiu, F. Battery thermal management systems: Current status and design approach of cooling technologies. Energies 2021, 14, 4879. [Google Scholar] [CrossRef]
  22. Wu, W.; Wang, S.; Wu, W.; Chen, K.; Hong, S.; Lai, Y. A critical review of battery thermal performance and liquid based battery thermal management. Energy Convers. Manag. 2019, 182, 262–281. [Google Scholar] [CrossRef]
  23. Lai, Y.; Wu, W.; Chen, K.; Wang, S.; Xin, C. A compact and lightweight liquid-cooled thermal management solution for cylindrical lithium-ion power battery pack. Int. J. Heat Mass Transf. 2019, 144, 118581. [Google Scholar]
  24. Murali, G.; Sravya, G.S.N.; Jaya, J.; Vamsi, V.N. A review on hybrid thermal management of battery packs and it’s cooling performance by enhanced PCM. Renew. Sustain. Energy Rev. 2021, 150, 111513. [Google Scholar]
  25. Singh, R.; Sadeghi, S.; Shabani, B. Thermal conductivity enhancement of phase change materials for low-temperature thermal energy storage applications. Energies 2019, 12, 75. [Google Scholar] [CrossRef] [Green Version]
  26. Ping, P.; Peng, R.; Kong, D.; Chen, G.; Wen, J. Investigation on thermal management performance of PCM-fin structure for Li-ion battery module in high-temperature environment. Energy Convers. Manag. 2018, 176, 131–146. [Google Scholar]
  27. Wu, W.; Yang, X.; Zhang, G.; Ke, X.; Wang, Z.; Situ, W.; Li, X.; Zhang, J. An experimental study of thermal management system using copper mesh-enhanced composite phase change materials for power battery pack. Energy 2016, 113, 909–916. [Google Scholar]
  28. Zhang, C.W.; Chen, S.R.; Gao, H.B.; Xu, K.J.; Xia, Z.; Li, S.T. Study of thermal management system using composite phase change materials and thermoelectric cooling sheet for power battery pack. Energies 2019, 12, 1937. [Google Scholar] [CrossRef]
  29. Zohuri, B. Heat Pipe Design and Technology Modern Applications for Practical Thermal Management, 2nd ed.; Springer: Cham, Switzerland, 2016. [Google Scholar]
  30. Behi, H.; Ghanbarpour, M.; Behi, M. Investigation of PCM-assisted heat pipe for electronic cooling. Appl. Therm. Eng. 2017, 127, 1132–1142. [Google Scholar] [CrossRef]
  31. Wu, M.-S.; Liu, K.; Wang, Y.-Y.; Wan, C.-C. Heat dissipation design for lithium-ion batteries. J. Power Sources 2002, 109, 160–166. [Google Scholar] [CrossRef]
  32. Lei, S.; Shi, Y.; Chen, G. Heat-pipe based spray-cooling thermal management system for lithium-ion battery: Experimental study and optimization. Int. J. Heat Mass Transf. 2020, 163, 120494. [Google Scholar]
  33. E, J.; Yi, F.; Li, W.; Zhang, B.; Zuo, H.; Wei, K.; Chen, J.; Zhu, H.; Zhu, H.; Deng, Y. Effect analysis on heat dissipation performance enhancement of a lithium-ion-battery pack with heat pipe for central and southern regions in China. Energy 2021, 226, 120336. [Google Scholar] [CrossRef]
  34. Xu, X.; Tang, W.; Fu, J.; Li, R.; Sun, X. Plate flat heat pipe and liquid-cooled coupled multistage heat dissipation system of Li-ion battery. Int. J. Energy Res. 2018, 43, 1133–1141. [Google Scholar] [CrossRef]
  35. Wang, Y.; Dan, D.; Xie, Y.; Li, W.; Guo, H.; Zhang, Y. Study on the influence of flat heat pipe structural parameters in battery thermal management system. Front. Energy Res. 2022, 9, 797664. [Google Scholar]
  36. Wang, Y.; Dan, D.; Zhang, Y.; Qian, Y.; Panchal, S.; Fowler, M.; Li, W.; Tran, M.; Xie, Y. A novel heat dissipation based on flat heat pipe for battery thermal management system. Int. J. Energy Res. 2022, 46, 15961–15980. [Google Scholar] [CrossRef]
  37. Huang, Q.; Li, X.; Zhang, G.; Zhang, J.; He, F.; Li, Y. Experimental investigation of the thermal performance of heat pipe pipe assisted phase change material for battery thermal management system. Appl. Therm. Eng. 2018, 141, 1092–1100. [Google Scholar]
  38. Boonma, K.; Mesgarpour, M.; NajmAbad, J.M.; Alizadeh, R.; Mahian, O.; Dalkılıç, A.S.; Ahn, H.S.; Wongwises, S. Prediction of battery thermal behaviour in the presence of a constructal theory-based heat pipe (CBHP): A multiphysics model and pattern-based machine learning approach. J. Energy Storage 2022, 48, 103963. [Google Scholar]
  39. Liang, J.; Gan, Y.; Li, Y. Investigation on the thermal performance of a battery thermal management system using heat pipe under different ambient temperatures. Energy Convers. Manag. 2018, 155, 1–9. [Google Scholar]
  40. Nasir, F.M.; Abdullah, M.Z.; Ismail, M.A. Experimental investigation of water-cooled heat pipes in the thermal management of lithium-ion EV batteries. Arab. J. Sci. Eng. 2019, 44, 7541–7552. [Google Scholar]
  41. Gan, Y.; He, L.; Liang, J.; Tan, M.; Xiong, T.; Li, Y. A numerical study on the performance of a thermal management system for a battery pack with cylindrical cells based on heat pipes. Appl. Therm. Eng. 2020, 179, 115740. [Google Scholar]
  42. Zhang, Z.; Wei, K. Experimental and numerical study of a passive thermal management system using flat heat pipes for lithium-ion batteries. Appl. Therm. Eng. 2020, 166, 114660. [Google Scholar]
  43. Zeng, W.; Niu, Y.; Li, S.; Hu, S.; Mao, B.; Zhang, Y. Cooling performance and optimization of a new hybrid thermal management system of cylindrical battery. Appl. Therm. Eng. 2022, 217, 119171. [Google Scholar]
  44. Chien, L.-H.; Shih, Y.-C. An experimental study of mesh type flat heat pipes. J. Mech. 2011, 27, 167–176. [Google Scholar]
  45. Rao, Z.; Wang, S.; Wu, M.; Lin, Z.; Li, F. Experimental investigation on thermal management of electric vehicle battery with heat pipe. Energy Convers. Manag. 2013, 65, 92–97. [Google Scholar]
  46. Wang, Q.; Jiang, B.; Xue, Q.; Sun, H.; Li, B.; Zou, H.; Yan, Y. Experimental investigation on EV battery cooling and heating by heat pipes. Appl. Therm. Eng. 2015, 88, 54–60. [Google Scholar] [CrossRef]
  47. Khateeb, S.A.; Amiruddin, S.; Farid, M.; Selman, J.R.; Al-Hallaj, S. Thermal management of Li-ion battery with phase change material for electric scooters: Experimental validation. J. Power Sources 2005, 142, 345–353. [Google Scholar]
  48. Yao, C.; Dan, D.; Zhang, Y.; Wang, Y.; Qian, Y.; Yan, Y.; Zhuge, W. Thermal Performance of a Micro Heat Pipe Array for Battery Thermal Management Under Special Vehicle-Operating Conditions. Automot. Innov. 2020, 3, 317–327. [Google Scholar]
  49. Wang, J.; Gan, Y.; Liang, J.; Tan, M.; Li, Y. Sensitivity analysis of factors influencing a heat pipe-based thermal management system for a battery module with cylindrical cells. Appl. Therm. Eng. 2019, 151, 475–485. [Google Scholar]
  50. Chen, M.; Li, J. Nanofluid-based pulsating heat pipe for thermal management of lithium-ion batteries for electric vehicles. J. Energy Storage 2020, 32, 101715. [Google Scholar]
  51. Jouhara, H.; Serey, N.; Khordehgah, N.; Bennett, R.; Almahmoud, S.; Lester, S.P. Investigation, development and experimental analyses of a heat pipe based battery thermal management system. Int. J. 2020, 1, 100004. [Google Scholar]
  52. Behi, H.; Karimi, D.; Behi, M.; Ghanbarpour, M.; Jaguemont, J.; Sokkeh, M.A.; Gandoman, F.H.; Berecibar, M.; van Mierlo, J. A new concept of thermal management system in Li-ion battery using Fair cooling and heat pipe for electric vehicles. Appl. Therm. Eng. 2020, 174, 115280. [Google Scholar]
  53. Alihosseini, A.; Shafaee, M. Experimental study and numerical simulation of a Lithium-ion battery thermal management system using a heat pipe. J. Energy Storage 2021, 39, 102616. [Google Scholar] [CrossRef]
  54. Li, Y.; Guo, H.; Qi, F.; Guo, Z.; Li, M.; Tjernberg, L.B. Investigation on liquid cold plate thermal management system with heat pipes for LiFePO4 battery pack in electric vehicles. Appl. Therm. Eng. 2021, 185, 116382. [Google Scholar]
  55. Behi, H.; Karimi, D.; Behi, M.; Jaguemont, J.; Ghanbarpour, M.; Behnia, M.; Berecibar, M.; Mierlo, J.V. Thermal management analysis using heat pipe in the high current discharging of lithium-ion battery in electric vehicles. J. Energy Storage 2020, 32, 101893. [Google Scholar] [CrossRef]
  56. Wu, W.; Yang, X.; Zhang, G.; Chen, K.; Wang, S. Experimental investigation on the thermal performance of heat pipe-assisted phase change material based battery thermal management system. Energy Convers. Manag. 2017, 138, 486–492. [Google Scholar]
  57. Wan, C. A Numerical Investigation of the Thermal Performances of an Array of Heat Pipes for Battery Thermal Management. Fluid Dyn. Mater. Processing 2019, 15, 343–356. [Google Scholar] [CrossRef]
  58. Wan, C. Prediction on Transient Thermal Performance of Heat Pipe Array for a Battery Thermal Management. Acad. J. Eng. Technol. Sci. 2020, 3, 199–208. [Google Scholar]
  59. Wan, C. Effect of Start-Up Time and Cycle Characteristic of Heat Pipe Array for a Battery Thermal Management. Acad. J. Eng. Technol. Sci. 2020, 3, 121–130. [Google Scholar]
  60. Wu, Y.; Zhang, Z.; Li, W.; Xu, D. Effect of the inclination angle on the steady-state heat transfer performance of a thermosyphon. Appl. Sci. 2019, 9, 3324. [Google Scholar] [CrossRef]
  61. Behi, H.; Karimi, D.; Jaguemont, J.; Berecibar, M.; van Mierlo, J. Experimental study on cooling performance of flat heat pipe for lithium-ion battery at various inclination angels. Energy Perspect. 2020, 1, 77–92. [Google Scholar]
Figure 1. Classification of battery cooling systems.
Figure 1. Classification of battery cooling systems.
Energies 15 08534 g001
Figure 2. Schematic diagram of a heat pipe.
Figure 2. Schematic diagram of a heat pipe.
Energies 15 08534 g002
Figure 3. Measurement results of heat generation rates at different C rates during (a) charging rates of 1 C, 1.5 C, and 2 C at Tamb 25 °C; (b) discharging rates of C/3, 0.5 C, and 1 C at Tamb 25 °C; (c) discharging rates of 1 C, 2 C, 3 C, and 4 C at Tamb 25 °C; (d) charging rate of 1 C at Tambs of 0 °C, 15 °C, 25 °C, and 35 °C; and (e) discharging rate of 1 C at Tambs of 30 °C, 15 °C, 0 °C, 15 °C, 25 °C, and 35 °C (from Hu et al. [6], with permission from Elsevier).
Figure 3. Measurement results of heat generation rates at different C rates during (a) charging rates of 1 C, 1.5 C, and 2 C at Tamb 25 °C; (b) discharging rates of C/3, 0.5 C, and 1 C at Tamb 25 °C; (c) discharging rates of 1 C, 2 C, 3 C, and 4 C at Tamb 25 °C; (d) charging rate of 1 C at Tambs of 0 °C, 15 °C, 25 °C, and 35 °C; and (e) discharging rate of 1 C at Tambs of 30 °C, 15 °C, 0 °C, 15 °C, 25 °C, and 35 °C (from Hu et al. [6], with permission from Elsevier).
Energies 15 08534 g003
Figure 4. The variation of (a) Tmax and (b) ΔT at different input powers and a mass flow rate of 0.0167 kg/s (from Mbulu et al. [5], with permission from Elsevier).
Figure 4. The variation of (a) Tmax and (b) ΔT at different input powers and a mass flow rate of 0.0167 kg/s (from Mbulu et al. [5], with permission from Elsevier).
Energies 15 08534 g004
Figure 5. Battery surface temperatures at Tamb of (a) 10 °C, (b) 26 °C, and (c) 35 °C (from Behi et al. [52], with permission from Elsevier).
Figure 5. Battery surface temperatures at Tamb of (a) 10 °C, (b) 26 °C, and (c) 35 °C (from Behi et al. [52], with permission from Elsevier).
Energies 15 08534 g005
Figure 6. Variation in battery surface temperatures with different ambient temperatures (from Alihosseini and Shafaee [53], with permission from Elsevier).
Figure 6. Variation in battery surface temperatures with different ambient temperatures (from Alihosseini and Shafaee [53], with permission from Elsevier).
Energies 15 08534 g006
Figure 7. (a) Tmax and (b) ΔT of a battery under different inlet coolant temperatures with an input power of 40 W and flow rate of 2 LPM (from Liang et al. [39], with permission from Elsevier).
Figure 7. (a) Tmax and (b) ΔT of a battery under different inlet coolant temperatures with an input power of 40 W and flow rate of 2 LPM (from Liang et al. [39], with permission from Elsevier).
Energies 15 08534 g007
Figure 8. Variation of Tmax at a 4 C discharge rate with different coolant temperatures (from Li et al. [54], with permission from Elsevier).
Figure 8. Variation of Tmax at a 4 C discharge rate with different coolant temperatures (from Li et al. [54], with permission from Elsevier).
Energies 15 08534 g008
Figure 9. Transient temperatures under different air velocities at a 5 C discharge rate (from Wu et al. [56], with permission from Elsevier).
Figure 9. Transient temperatures under different air velocities at a 5 C discharge rate (from Wu et al. [56], with permission from Elsevier).
Energies 15 08534 g009
Figure 10. Effects of coolant flow rates on pressure drops [57].
Figure 10. Effects of coolant flow rates on pressure drops [57].
Energies 15 08534 g010
Figure 11. Effects of coolant flow rates on pressure drops (Pa) (from Gan et al. [41], with permission from Elsevier).
Figure 11. Effects of coolant flow rates on pressure drops (Pa) (from Gan et al. [41], with permission from Elsevier).
Energies 15 08534 g011
Figure 12. Tmax and ΔT of the battery pack at a 2 C discharging rate with different condenser lengths (from Gan et al. [41], with permission from Elsevier).
Figure 12. Tmax and ΔT of the battery pack at a 2 C discharging rate with different condenser lengths (from Gan et al. [41], with permission from Elsevier).
Energies 15 08534 g012
Table 1. Summary of the heat pipe cooling systems reported in the literature.
Table 1. Summary of the heat pipe cooling systems reported in the literature.
Fluid Used for Cooling at Condenser EndAuthorsHeat SourceHeat PipeHeat
Generation Rate
TambCoolant Inlet Temp.TmaxΔT
WaterMbulu et al. [5]
(Experiment)
Battery
surrogate
16 heat pipes (copper)30–60 W-30 °C<55 °C<5 °C
WaterLiang et al. [39]
(Experiment)
2 simulated
batteries
(battery
surrogate)
4 heat pipes
(copper)
20–50 W15–35 °C15–35 °C<40 °C<5 °C
Air and waterNasir et al. [40]
(Experiment)
Proxy cells
(battery
surrogate)
2 heat pipes
(aluminium plate)
10–35 W--<50 °C <5 °C
WaterGan et al. [41]
(Experiment)
24 cylindrical cells15 heat pipes (copper)1 and 2 C
discharge
25 °C20 °C<40 °C<5 °C
Natural convectionZhang and Wei [42]
(Experiment)
5 prismatic cells (16 V,
8.5 Ah)
6 flat heat pipes (aluminium plate)4, 6, and
8 C rate
25 °C-<40 °C<5 °C
Table 2. Summary of studies of the heat pipe cooling system at different heat generation rates and coolant temperatures.
Table 2. Summary of studies of the heat pipe cooling system at different heat generation rates and coolant temperatures.
Fluid Used for Cooling at Condenser EndAuthorsHeat SourceHeat PipeHeat Generation RateTambCoolant Flow RateCoolant Inlet Temp.TmaxΔT
WaterMbulu et al. [5]
(Experiment)
Battery
surrogate
16 heat pipes
(copper)
30–60 W-1 LPM30 °C<55 °C<5 °C
WaterRao et al. [45]
(Experiment)
Aluminium rectangular heater (battery
surrogate)
4 heat pipes
(copper)
10–60 W--25 ± 0.5 °C<50 °C<5 °C
WaterLiang et al. [39]
(Experiment)
Simulated
battery (battery
surrogate)
4 heat pipes
(copper)
20–50 W15, 25, 30, and 35 °C1, 2, and 3 LPM15–35 °C<40 °C<5 °C
AirYao et al. [48]
(Experiment)
Heater (battery surrogate)MHPA10–50 W24.8 °C3.3 ms−124.8 °C<40 °C3.44 °C
AirZhang and Wei [42]
(Experiment and
simulation)
5 prismatic cells (3.65 V, 8.5 Ah)6 flat heat pipes
(aluminium)
2, 4, 6, and
8 C rate
25 ± 2 °C-25 ± 2 °C<40 °C<5 °C
WaterWang et al. [49]
(Experiment and simulation)
12 cylindrical cells
(3.7 V,
1.96 Ah)
5 heat pipes
(copper)
3 and 5 C discharge 25 °C2 LPM25 °C27.62 °C 1.08 °C
TiO2 nano-fluidChen and Li [50]
(Experiment)
Prismatic cells (3.2 V, 68 Ah)PHP 0.5, 1, and 1.5 C discharge25, 30, and
35 °C
-25 °C<42.22 °C<2 °C
Water and refrigerantJouhara et al. [51]
(Experiment)
16 prismatic cells (2.3 V,
23 Ah)
Flat heat pipe (heat mat)4 C-9.5 LPM-<28 °C±1 °C
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Boonma, K.; Patimaporntap, N.; Mbulu, H.; Trinuruk, P.; Ruangjirakit, K.; Laoonual, Y.; Wongwises, S. A Review of the Parameters Affecting a Heat Pipe Thermal Management System for Lithium-Ion Batteries. Energies 2022, 15, 8534. https://0-doi-org.brum.beds.ac.uk/10.3390/en15228534

AMA Style

Boonma K, Patimaporntap N, Mbulu H, Trinuruk P, Ruangjirakit K, Laoonual Y, Wongwises S. A Review of the Parameters Affecting a Heat Pipe Thermal Management System for Lithium-Ion Batteries. Energies. 2022; 15(22):8534. https://0-doi-org.brum.beds.ac.uk/10.3390/en15228534

Chicago/Turabian Style

Boonma, Kittinan, Napol Patimaporntap, Hussein Mbulu, Piyatida Trinuruk, Kitchanon Ruangjirakit, Yossapong Laoonual, and Somchai Wongwises. 2022. "A Review of the Parameters Affecting a Heat Pipe Thermal Management System for Lithium-Ion Batteries" Energies 15, no. 22: 8534. https://0-doi-org.brum.beds.ac.uk/10.3390/en15228534

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop